Abstract
The development of hole-transport layers (HTLs) that elevate charge extraction, improve perovskite crystallinity, and decrease interfacial recombination is extremely important for enhancing the performance of inverted perovskite solar cells (PSCs). In this work, the nanoporous nickel oxide (NiOx) layer as well as NiOx thin film was prepared via chemical bath deposition as the HTL. The sponge-like structure of the nanoporous NiOx helps to grow a pinhole-free perovskite film with a larger grain size compared to the NiOx thin film. The downshifted valence band of the nanoporous NiOx HTL can improve hole extraction from the perovskite absorbing layer. The device based on the nanoporous NiOx layer showed the highest efficiency of 13.43% and negligible hysteresis that was better than the one using the NiOx thin film as the HTL. Moreover, the PSCs sustained 80% of their initial efficiency after 50 days of storage. This study provides a powerful strategy to design PSCs with high efficiency and long-term stability for future production.
1. Introduction
Perovskite solar cells (PSCs) have made an impressive progress with maximum power conversion efficiency (PCE) from 3.8 to 25.5% within a decade1,2 due to high absorption in the visible region,3 long carrier diffusion length,4 high carrier mobility,5 low exciton binding energy,6 and tunable band gaps by exchanging composition.7,8 Single cationic perovskite materials such as methylammonium lead iodide (MAPbI3) and formamidinium lead iodide (FAPbI3) have been proposed as the absorbing layer in PSCs with PCE values up to 20.8 and 18.94%, respectively.9,10 Compared with single cationic perovskites, multiple-cation perovskite materials solve the disadvantage of instability and achieve higher device performance. Saliba et al. incorporated rubidium cations into PSCs to reveal an optimized open-circuit voltage (VOC) of 1180 mV, a short-circuit current density (JSC) of 22.8 mA/cm2, a fill factor (FF) of 81%, and a certified PCE of 21.8%.11 Furthermore, the device retained 95% of its initial performance after 500 h of aging at 85 °C in a nitrogen-filled glovebox. Jeon et al. demonstrated PSCs using (FAPbI3)0.85(MAPbBr3)0.15 as the active layer, confirming a certified PCE of 23.2% with a slight hysteresis behavior.49 The device maintained 92.6% of its initial PCE value after 310 h of continuous illumination and almost 95% for more than 500 h of thermal annealing at 60 °C. Saliba et al. reported cesium-containing triple cation perovskite CsX(FA0.17MA0.83)1–XPb(I0.83Br0.17)3 as the light absorber.12 The PCE value of the optimized PSC dropped from 21.17 to ∼18% after aging for 250 h under constant illumination in a nitrogen atmosphere. Apparently, the adoption of multiple-cation perovskite materials is a good choice for commercialized production in the near future.
PSCs have been extensively developed in two different device configurations, that is, regular and inverted types. The regular PSC has a device structure of anode/electron-transport layer (ETL)/perovskite/hole-transport layer (HTL)/cathode since it is derived from the first perovskite-related literature.1 However, the n–i–p configuration usually encounters a serious drawback of hysteresis, which causes efficiency drop and instability of devices.13−15 Another problem is device instability which arises from the morphological deformation of (2,2′,7,7′-tetrakis(N,N-di-p-methoxyphenylamine)-9,9′-spirobifluorene) (spiro-MeOTAD) as the HTL.16,17 To avoid the above problems, inverted PSCs with the p–i–n configuration have been developed because of their simple device structure, free of high-temperature processing, and little hysteresis effect.18−20 Besides, inverted PSCs can be combined with traditional solar cells such as silicon or copper indium gallium selenide solar cells to construct tandem devices with high efficiency.21,22 In the inverted PSC structure, poly(3,4-ethylenedioxythiophene):poly(styrene sulfonate) (PEDOT:PSS) has been widely used as the HTL owing to the simple fabrication process and hole extraction ability from the perovskite layer.23,24 However, organic PEDOT:PSS may corrode and deteriorate transparent conductive oxide electrodes, such as fluorine-doped tin oxide (FTO) and indium tin oxide, due to its acidic nature.25
Apart from organic materials such as PEDOT:PSS, inorganic materials can also serve as the HTL in inverted p–i–n devices, which possess better thermal stability, higher hole mobility, and lower expense compared with traditional organic hole-transport materials.26 It is believed that the development of inorganic hole-transport materials plays an important role in the commercialization of PSCs with low cost, high efficiency, and long-term stability. Among miscellaneous metal oxide materials, nickel oxide (NiOx) is particularly attractive as it belongs to p-type semiconductors with high hole mobility, excellent chemical stability, and good transmittance in the visible range.26,27 The matched energy level alignment between NiOx and the perovskite facilitates hole extraction and electron blocking.28,29 Many approaches have been proposed to prepare NiOx HTLs for inverted PSCs, including sol–gel deposition,30 nanoparticle dispersion,31 spray pyrolysis,32 and atomic layer deposition.33 Moreover, an additional passivation layer can be introduced between NiOx and the perovskite layer to reduce pinhole defects on the NiOx surface and increase the crystallinity of the perovskite. Wang et al. utilized polystyrene as a passivation medium to modify the surface of the NiOx film and increase the perovskite quality.34 Highly efficient devices with a larger perovskite grain size, fewer interfacial defects, and suppressed charge recombination were achieved. A very high PCE of 19.99% and a VOC of 1.149 V were obtained, while no hysteresis was observed. Chen et al. adopted a conjugated poly(bithiophene imide) (PBTI) for the passivation of grain boundaries in inverted planar PSCs.35 The incorporation of PBTI between NiOx and the perovskite resulted in lower defect density, reduced charge recombination, and high efficiency of devices. An optimized PCE of 20.67% was obtained with the PBTI treatment.
In this research, we investigated the formation mechanism and characterization of nanoporous NiOx for the fabrication of inverted PSCs. Nanoporous NiOx has been reported to assist the deposition of a pinhole-free perovskite film with larger grain size.36 Different precursors including nickel sulfate heptahydrate (NiSO4·7H2O), nickel acetate tetrahydrate (Ni(CH3COO)2·4H2O), and nickel chloride (NiCl2) have been adopted to examine the formation condition of nanoporous NiOx layers. Transmission spectroscopy and ultraviolet photoelectron spectroscopy (UPS) were carried out to explore the transmittance in the visible range and energy levels of nanoporous NiOx, respectively. From the literature survey, we notice that UPS has not been applied to investigate the band structure of nanoporous NiOx so far. As for the fabrication of inverted PSCs, [6,6]-phenyl-C61-butyric acid methyl ester (PC61BM) doped with tetrabutylammonium tetrafluoroborate (TBABF4) and titanium (diisopropoxide) bis(2,4-pentanedionate) (TIPD) were chosen as ETLs. The device with the configuration of FTO/NiOx/perovskite/PC61BM/TIPD/Ag was fabricated and evaluated. Both NiOx thin film and nanoporous NiOx layer were used as the HTL for comparison.
2. Experimental Section
2.1. Materials
FTO-coated glass substrates were purchased from Ruilong Optoelectronics Technology Co., Ltd. Potassium persulfate (K2S2O8, purity 98%) was purchased from Showa. NiSO4·7H2O (purity 98%) and aqueous ammonia (NH3(aq), 25–28 wt %) were bought from Acros and Sigma-Aldrich, respectively. High-purity perovskite precursors including lead iodide (PbI2, purity 99.9985%), lead bromide (PbBr2, purity 99.99%), and methylammonium bromide (MABr, purity 99.5%) were purchased from Luminescence Technology Corp., Taiwan. Formamidinium iodide (FAI, purity 98%) was bought from STAREK Scientific Co., Ltd. Cesium iodide (CsI, purity 99.9%) was bought from Alfa Aesar. PC61BM (purity 99%) was purchased from Solenne B.V., the Netherlands. Other chemicals and solvents were purchased from Alfa Aesar or Acros and used without further purification.
2.2. Preparation of the NiOx Thin Film
The NiOx thin film was spin-cast from its precursor solution on the FTO-coated glass substrate, which was cleaned stepwise in detergent, deionized water, acetone, and IPA under ultrasonication for 30 min each, followed by UV–ozone exposure for 25 min. To prepare the 0.1 M NiOx precursor solution, 2.62 mg of NiSO4·7H2O was dissolved in 10 mL of methanol at 80 °C with stirring in a sealed glass vial overnight. The NiOx precursor solution was then spin-coated on the FTO substrate at 1500 rpm for 30 s, followed by drying at 80 °C for 10 min. After drying, the substrate was moved into a high-temperature oven for calcination at 450 °C for 1 h to obtain the NiOx thin film.
2.3. Preparation of the Nanoporous NiOx Layer
The nanoporous NiOx layer was prepared according to the previous literature with some modified parameters.36 The pre-cleaned FTO substrate was taped in a Petri dish with FTO face upward. To prepare the NiOx growth solution, 8 mL of 0.1 M NiSO4·7H2O was added to a solution of 6 mL of 0.025 M K2S2O8 in DI water and 2 mL of aqueous ammonia (25–28 wt %). The mixed growth solution was shaken and immediately poured into the Petri dish. The reaction was carried out for 5 min and the substrate was taken out, cleaned with deionized water to remove loose black particles, and further dried at 100 °C for 1 h. After drying, the substrate was moved into a high-temperature oven for calcination at 450 °C for 1 h to obtain the nanoporous NiOx layer.
2.4. Device Fabrication
After the preparation of the NiOx thin film or nanoporous NiOx layer, the absorbing perovskite Cs0.05FA0.81MA0.14Pb(Br0.15I0.85)3 layer was deposited by spin coating. For the perovskite solution used in this research, a mixture of CsI (17.5 mg), FAI (190.2 mg), MABr (21.8 mg), PbI2 (548.6 mg), and PbBr2 (77.1 mg) was dissolved in a mixed solvent (1 mL) consisting of N,N-dimethylformamide and dimethyl sulfoxide with a 4:1 volume ratio at 70 °C for 1 h with stirring. The perovskite film was deposited on top of the NiOx thin film or nanoporous NiOx layer with a two-step spin-coating process in a nitrogen-filled glovebox. The first step was 1000 rpm for 10 s with a ramp-up rate of 200 rpm/s, and the second step was 5000 rpm for 20 s with a ramp-up rate of 1000 rpm/s. Chlorobenzene (300 μL) was dropped onto the substrate 10 s before the end of the second spin-coating step. The substrate was then annealed at 100 °C for 1 h. The PC61BM solution (20 mg/mL) containing 2 wt % of TBABF4 in chlorobenzene was spin-coated on the formed perovskite layer at 3000 rpm for 30 s, followed by heating at 100 °C for 10 min. Afterward, 0.1 wt % of TIPD solution in isopropanol was spin-coated on the PC61BM layer at 5000 rpm for 30 s. Finally, 100 nm of the Ag electrode was deposited by thermal evaporation at a base pressure of 7 × 10–6 Torr. The active area of each device is 4 mm2.
2.5. Characterization Methods
The cross-sectional and top-view scanning electron microscopy (SEM) micrographs of samples were investigated with an ultrahigh-resolution ZEISS AURIGA Crossbeam scanning electron microscope. The morphology and roughness of samples were examined with a Bruker Innova atomic force microscopy (AFM) with the tapping mode. The UPS measurements for the NiOx thin film and nanoporous NiOx layers were performed on a Thermo VG-Scientific/Sigma Probe spectrometer. A He I (hν = 21.22 eV) discharge lamp was used as the excitation source. X-ray photoelectron spectroscopy (XPS) measurements were conducted by a Thermo K-Alpha X-ray photoelectron spectrometer for elemental composition analysis of samples. The steady-state photoluminescence (PL) spectra of the perovskites on different substrates were measured using a Princeton Instruments Acton 2150 spectrophotometer. A KIMMON KOHA He–Cd laser with double excitation wavelengths at 325/442 nm was utilized as the light source. The transmission spectra were recorded with the same spectrophotometer using a xenon lamp (ABET Technologies LS 150) as the light source. To perform time-resolved PL (TR-PL) measurements, a 473 nm pulsed laser (Omicron) was utilized as an excitation light source. The TR-PL signals were recorded by a time-correlated single-photon counting module (PicoQuant MultiHarp 150 4N) combined with a photomultiplier tube through an Andor Kymera 328i spectrometer. The apparatus was assembled by LiveStrong Optoelectronics Co., Ltd. from Taiwan. X-ray diffraction (XRD) patterns and crystallinity of samples were measured by a Rigaku D/MAX2500 X-ray diffractometer. The current density–voltage (J–V) characteristics of the PSCs were measured using a Keithley 2400 SourceMeter under AM 1.5G simulated sunlight exposure (Yamashita Denso YSS-150A equipped with a xenon short arc lamp, 1000 W) at 100 mW/cm2 under an ambient environment. The scan rate for J–V measurements was 25 mV/s. The external quantum efficiency (EQE) measurements were performed on an assembled apparatus in the laboratory, comprising a solar simulator (ABET Technologies LS 150, USHID UXL-150MO), a monochromator (Prince Instruments Acton 2150), and a Keithley 2400 SourceMeter.
3. Results and Discussion
3.1. Characterization of the NiOx Thin Film and Nanoporous Layers
The NiOx thin film was prepared via the sol–gel process and its formation has been discussed previously.37,38 Meanwhile, the nanoporous NiOx layer was prepared by chemical bath deposition.39−41 The involved chemical reactions are listed as follows.
| 1 |
| 2 |
The first step (1) describes the reaction between the starting material NiSO4 and ammonia in water to form nickel hydroxide Ni(OH)2. Then, the second step (2) reveals the formation of the species NiO(OH) by reacting Ni(OH)2 with S2O82– from K2S2O8, as indicated in the part of Section 2.3. Obviously, in this step, the persulfate salt induces Ni2+ oxidation to generate Ni3+ species NiO(OH). Afterward, the as-deposited precursor film containing Ni(OH)2 and NiOOH was thermally converted to NiO by calcination. The advantages of this approach include simple processing, usage of DI water as solvent, and short reaction time within 10 min. Figure 1a, c shows the top-view and cross-sectional SEM images of the NiOx thin film, respectively. It can be seen that a thin and dense NiOx layer is deposited on the FTO surface that possesses many grain boundaries. The thickness of the NiOx thin film was estimated to be ca. 30 nm. Figure 1b,d reveals the top-view and cross-sectional SEM images of the nanoporous NiOx layer, respectively. Sponge-like nanostructures and interconnecting networks were clearly observed, exhibiting higher surface area for perovskite filling and crystallization. The formed nanoporous NiOx layer has a high degree of porosity with a pore size of 100–300 nm and a wall thickness of 20–30 nm. Furthermore, nanoporous NiOx flakes were vertically aligned on the FTO substrate with a thickness ranging from 100 to 120 nm. The cross-sectional SEM image of the nanoporous NiOx layer can be seen in Figure S1 in the Supporting Information, revealing a flake-like morphology. The surface morphology of our nanoporous NiOx layer is similar to those in the previous literature.36,39 Apart from the SEM observation, AFM experiments were also carried out to investigate the morphology and average roughness (Ra) of the prepared samples. Figure 1e,f shows the topographic AFM images of the NiOx thin film and nanoporous layer, respectively. Many NiOx nanospheres were formed on the FTO substrate, with a low Ra value of 11.1 nm. In Figure 1f, highly porous flakes were aligned on the FTO surface, which are consistent with the SEM observation. The Ra value of the nanoporous NiOx layer is estimated to be 13.3 nm, which is slightly larger than that of the NiOx thin film due to its sponge-like interconnecting networks.
Figure 1.
Top-view and cross-sectional SEM images of the (a,c) NiOx thin films and (b,d) nanoporous NiOx layer deposited on FTO substrates; AFM topographic images of the (e) NiOx thin film and (f) nanoporous NiOx layer.
Figure 2a shows the transmission spectra of the NiOx thin film and nanoporous layer from 300 to 750 nm. It is seen that the nanoporous NiOx layer has a higher transmittance of 60–90% in the range of 300–450 nm and an even higher transmittance of 90–95% in the range of 450–750 nm, as compared with the NiOx thin film. This is beneficial for incident photons to enter devices and to be absorbed by the absorbing layer. Figure 2b reveals the absorption spectra of the NiOx thin film and nanoporous layer in the visible range. The optical band gaps (Eg) of the NiOx thin film and nanoporous layer were estimated from their absorption edges around 350 nm to give 3.48 and 3.5 eV, respectively, which is similar to the previous literature.42 From transmission and absorption measurements, we conclude that light is easier to pass through the nanoporous NiOx layer owing to its porous structure, as compared with the NiOx thin film.
Figure 2.

(a) Transmission and (b) absorption spectra of the NiOx thin film and nanoporous NiOx layer.
The UPS spectra of the NiOx thin film and nanoporous layer were measured to examine the change of the energy levels, as shown in Figure 3. The work function (φw) is derived by subtracting the binding energy cutoff in the high binding energy region (around 14.1 eV) from the He I photon energy (21.22 eV).23 Since φw is defined as the energy difference between the EF and the vacuum level, the EF values of the NiOx thin film and nanoporous layer were determined to be −7.05 and −7.1 eV, respectively, from Figure 3a. Furthermore, the binding energy cutoffs in the low binding energy region around −1.9 eV indicate the energy difference between the EF and the valence band (VB) level.43 Therefore, the VB levels of the NiOx thin film and nanoporous layer were calculated to be −5.16 and −5.23 eV, respectively, from Figure 3b. By combining VB levels from UPS experiments and Eg values from optical measurements, the conduction band (CB) levels of the NiOx thin film and nanoporous layer were calculated to be −1.68 and −1.73 eV, respectively. The above results demonstrate that the energy levels of NiOx can be slightly altered by different nanostructures. The downshifted VB level of the nanoporous NiOx layer is matched better with the perovskite absorbing layer than the NiOx thin film, which can improve the hole extraction from perovskite to NiOx HTL.
Figure 3.

UPS spectra of the NiOx film and nanoporous NiOx layer at (a) high and (b) low binding energy regions.
The XRD patterns of the NiOx thin film and nanoporous layer on the FTO substrates are shown in Figure 4. The diffraction signals of the NiOx thin film are found at 2θ = 37.0, 43.2, and 62.8°, corresponding to the (111), (200), and (220) planes, respectively.44 According to the XRD patterns, the prepared NiOx is well consistent with the cubic phase.45 Furthermore, the three diffraction signals of the nanoporous NiOx layer are also observed at similar 2θ positions, confirming that the crystalline structure of NiOx is not affected by different nanostructures.
Figure 4.
XRD patterns of the NiOx thin film and nanoporous NiOx layer deposited on FTO substrates.
To further probe Ni3+ and Ni2+ components in the NiOx thin film and nanoporous layer, XPS experiments were performed and the corresponding Ni 2p3/2 and O 1s signals are displayed in Figure 5. The multicomponent band can be deconvoluted into four different states at 854.0 (Ni2+), 855.8 (Ni3+), 861.2 (Ni2+ satellite), and 864.1 eV (Ni3+ satellite) that have been reported in the previous literature.29,30 The received Ni3+/Ni2+ ratios were calculated from Figure 5a,b to be 1.23 and 1.25 for the NiOx thin film and nanoporous layer, respectively. The slightly increased Ni3+/Ni2+ ratio means that the nanoporous structure has better hole-transport ability than the thin film. Figure 5c,d shows the O 1s spectra of the NiOx thin film and nanoporous layer, being fitted with two states at around 529.3 eV (O2– from NiO) and 531.3 eV (O2– from Ni2O3).36 The ratios of the obtained O element from NiO and Ni2O3 were calculated to be 0.424 and 0.454 for the NiOx thin film and nanoporous layer. Similarly, it shows that the NiOx nanoporous layer has a higher proportion of Ni3+ to contribute to hole conductivity compared to the thin-film state.
Figure 5.
XPS spectra of Ni 2p3/2 and O 1s elements in the (a,c) nanoporous NiOx layer and (b,d) NiOx thin film.
Figure 6a,b shows the cross-sectional SEM image of the perovskite deposited on the NiOx thin film and nanoporous NiOx layer, respectively. The thickness of the perovskite layer on both HTLs was estimated to be 550 nm. It is evident that the large and intact perovskite crystals with a submicron grain size up to 500 nm were formed when deposited on the nanoporous NiOx layer, as revealed in Figure 6b. In contrast, the perovskite crystals deposited on the NiOx thin film looked smaller with an average grain size of 300 nm, as shown in Figure 6a. The perovskite films with a large grain size are beneficial for reducing charge recombination, allowing effective carrier extraction and transport from the perovskite absorbing layer to the corresponding charge-transport layers.34,46 The top-view SEM images of the perovskite deposited on the NiOx thin film and nanoporous NiOx layer are displayed in Figure 6c,d, respectively. No pinholes could be found for both perovskite films. The grain size of perovskite crystals on the NiOx thin film is estimated to be in the range of 100–200 nm in Figure 6c, while larger perovskite crystals with a grain size of 100–350 nm were observed on the nanoporous NiOx layer in Figure 6d. The highly porous surface of the nanoporous NiOx layer provides nucleation sites for the perovskite growth to achieve a larger grain size compared with the NiOx thin film. Our results prove that the nanoporous NiOx layer can perform as a template to obtain a high-quality perovskite film.
Figure 6.
Cross-sectional SEM images of PSCs based on the (a) NiOx thin film and (b) nanoporous NiOx layer as the HTL; top-view SEM images of the perovskite deposited on the (c) NiOx thin film and (d) nanoporous NiOx layer.
The XRD patterns of the perovskite deposited on the NiOx thin film and nanoporous NiOx layer are shown in Figure 7. The composition of the deposited perovskite layer was Cs0.05FA0.81MA0.14Pb(Br0.15I0.85)3, as mentioned in the Experimental Section. Intense diffraction peaks at 2θ = 14.04, 19.92, 24.52, 28.36, 31.76, 34.96, 40.52, and 43.16° are observed, which correspond to (001), (011), (111), (002), (012), (112), (022), and (003) planes. The location of these diffraction peaks of the perovskite is consistent with the previous reports.47 Moreover, the intensity of diffraction peaks of the perovskite on the nanoporous NiOx is found to be higher than that on the NiOx thin film, implying better crystallization and morphology of the perovskite film. Here again, the nanoporous NiOx layer can serve as a better template for the perovskite growth to obtain higher crystallinity.
Figure 7.
XRD patterns of the perovskite films on the NiOx thin film and nanoporous NiOx layer.
The steady-state PL spectra of the perovskite on the glass, NiOx thin film, and nanoporous NiOx layer are revealed in Figure 8a. The PL emission of the perovskite Cs0.05FA0.81MA0.14Pb(Br0.15I0.85)3 is located at 755 nm that is similar to the previous literature.11 It is seen that the perovskite deposited on the glass has the highest PL intensity, while the one on the nanoporous NiOx layer owns the lowest PL emission. The reduced PL emission can be attributed to the nanoporous structure that helps to extract carriers more efficiently from the perovskite layer, indicative of the decreased charge recombination and improved JSC value for device fabrication.45 The TR-PL measurements of the perovskite on the glass substrate, NiOx thin film, and nanoporous NiOx layer were carried out and the corresponding results are revealed in Figure 8b. It is clearly seen that the perovskite deposited on the nanoporous NiOx layer possesses a faster PL decay curve compared with that on the NiOx thin film, implying that more effective hole–electron separation can be accomplished. The PL decay curves were fitted using a biexponential model and the average lifetime (τavg) was estimated from the fitted curve data using the equation τavg = ∑i=1nAiτi2/∑i=1Aiτi,27,31 where Ai is a constant and τi is the lifetime. The perovskite Cs0.05(MA0.85FA0.15)0.95Pb(Br0.15I0.85)3 on the glass substrate had a τavg of 121.47 ns, while the perovskite deposited on the NiOx thin film and nanoporous NiOx layer showed shorter τavg values of 34.3 and 26.52 ns, respectively. This indicates more effective charge extraction by the nanoporous NiOx HTL from the perovskite active layer as compared with that by the NiOx thin film.
Figure 8.

(a) PL emission spectra and (b) TR-PL decay curves of the perovskite on the glass substrate, NiOx thin film, and nanoporous NiOx layer.
The p–i–n device structure of the inverted PSC based on the nanoporous NiOx HTL is shown in Figure 9a, revealing a sandwiched FTO/nanoporous NiOx/Cs0.05(MA0.85FA0.15)0.95Pb(Br0.15I0.85)3/PCBM/TIPD/Ag architecture. The energy level diagram of the whole device is depicted in Figure 9b. The VB and CB levels of NiOx have been discussed in the previous part, while the energy levels of the remaining components were referred to our previous report.43 In our device architecture, electrons can be smoothly transported from the perovskite absorbing layer to the Ag electrode through PCBM/TIPD, while holes are migrated stepwise from the perovskite layer to nanoporous NiOx and collected at the FTO electrode. Moreover, the deeper VB level of the nanoporous NiOx HTL would result in a higher VOC value. The J–V curves of the devices under AM 1.5 G illumination are shown in Figure 9c, and the measured parameters including JSC, VOC, FF, PCE, series resistance (RS), and shunt resistance (RSh) are summarized in Table 1. The optimized device based on the nanoporous NiOx showed a VOC of 1.02 V, a JSC of 18.9 mA/cm2, a FF of 70%, and a PCE of 13.43% in the reverse scan, which is significantly higher than the one based on the NiOx thin film (VOC = 1 V, JSC = 16 mA/cm2, FF = 66%, and PCE = 10.53%). The statistical distribution of 20 individual devices for all photovoltaic parameters is depicted in Figure S3 in the Supporting Information. It can be seen that our devices possessed good reproducibility and PSCs based on the nanoporous NiOx layer showed relatively higher photovoltaic parameters. The improved device performance is mainly ascribed to the increased JSC value. In order to investigate the reason to the increased JSC, hole-only devices with the configuration of FTO/NiOx thin film or nanoporous NiOx layer/Ag were fabricated and their current–voltage characteristics were measured and are displayed in Figure S2 in the Supporting Information. It reveals that the nanoporous NiOx layer has a higher conductivity than the NiOx thin film, indicative of the enhanced hole-transport ability. The hole mobility (μh) was inferred from the space-charge limited current equation J = (9/8)εε0μh(V2/L3). The μh values of the nanoporous NiOx layer and the NiOx thin film are calculated to be 2.41 × 10–3 and 2.03 × 10–3 cm2/V s, respectively, which are close to the previous report.43 The increased JSC value can be ascribed to the enhanced hole mobility. The reduced charge recombination, as discussed in the PL and TR-PL part, is also responsible for the increased JSC value. We also notice that VOC of the device based on the nanoporous NiOx layer is higher than that on the NiOx thin film, which has been predicted from the downshifted VB levels. Furthermore, the PSC using the nanoporous NiOx HTL exhibited negligible hysteresis compared with that using the NiOx thin film as the HTL. The eliminated hysteresis is realized due to the higher crystallinity and reduced surface defect of the perovskite, which has been reported in the previous study.48Figure 9d shows the EQE spectra and integrated current densities of devices as a function of wavelength using the nanoporous NiOx layer and NiOx thin film as the HTL. The results demonstrate that the device based on the nanoporous NiOx has a higher photon-to-electron conversion capability from 350 to 700 nm compared to that based on the NiOx thin film. The integrated J values for the devices based on the nanoporous NiOx layer and NiOx thin film were calculated to be 18.2 and 15.6 mA/cm2, respectively, which are similar to the JSC values from Table 1. As shown in Figure 9e, both the resulting PSCs using the nanoporous NiOx layer and NiOx thin film as the HTL maintained 80% of their initial efficiency over a period of 50 days stored in the nitrogen glovebox and measured in ambient air at 25 °C, revealing long-term stability for future production.
Figure 9.
(a) Device structure based on the nanoporous NiOx, (b) energy level diagram, (c) J–V characteristics, (d) EQE spectra and integrated current density, and (e) normalized PCE evolution of devices based on the NiOx thin film and nanoporous NiOx layer.
Table 1. Device Performance of Inverted PSCs Based on Different NiOx HTLs.
| HTL | scan direction | JSC (mA/cm2) | VOC (V) | FF (%) | best PCE (%) | avg PCEa (%) | RS (Ω·cm2) | RSh (kΩ·cm2) |
|---|---|---|---|---|---|---|---|---|
| NiOx thin film | forward | 16.9 | 0.98 | 62 | 9.87 | 9.36 | 278 | 9834 |
| reverse | 16.0 | 1 | 66 | 10.53 | 9.97 | 252 | 10,573 | |
| nanoporous NiOx | forward | 18.9 | 1.02 | 69 | 13.26 | 12.54 | 213 | 13,160 |
| reverse | 18.9 | 1.02 | 70 | 13.43 | 12.73 | 194 | 15,657 |
Data were obtained from 20 devices.
4. Conclusions
In this research, we compared two different nanostructures of NiOx applying in inverted PSCs. The sponge-like nanostructures and interconnecting networks of NiOx can provide nucleation sites for the perovskite growth to achieve a larger grain size compared with the NiOx thin film. The higher crystallinity and pinhole-free perovskite could generate carriers more efficiently, while effective extraction of carriers by the nanoporous NiOx HTL and PCBM/TIPD ETL was achieved to reduce charge recombination and improve JSC of devices. The optimized PSC based on the nanoporous NiOx layer exhibited a high PCE of 13.43%, negligible hysteresis, and excellent device stability for 50 days of storage. To date, our results provide a simple and effective approach to achieve a high-quality perovskite absorbing layer on the nanoporous NiOx layer for future application in photovoltaics.
Acknowledgments
The authors thank the Ministry of Science and Technology (MoST) of Republic of China (grant numbers MOST 107-2221-E-009-042-MY3 and MOST 109-2112-M-006-016) and the Higher Education Sprout Project of the Ministry of Education (MOE), Taiwan, for financially supporting this work.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.1c01378.
Cross-sectional SEM image of the nanoporous NiOx layer; I–V curves of hole-only devices based on the nanoporous NiOx layer and NiOx thin film; and performance variation represented as a standard box plot in PCE, JSC, FF, and VOC from 20 devices based on the nanoporous NiOx layer and NiOx thin film under forward and reverse scans (PDF)
The authors declare no competing financial interest.
Supplementary Material
References
- Kojima A.; Teshima K.; Shirai Y.; Miyasaka T. Organometal Halide Perovskites as Visible-Light Sensitizers for Photovoltaic Cells. J. Am. Chem. Soc. 2009, 131, 6050–6051. 10.1021/ja809598r. [DOI] [PubMed] [Google Scholar]
- National Renewable Energy Laboratory (NREL) . Best research-cell efficiency chart. [Online]. https://www.nrel.gov/pv/cell-efficiency.html (2021-04-12).
- Ball J. M.; Stranks S. D.; Hörantner M. T.; Hüttner S.; Zhang W.; Crossland E. J. W.; Ramirez I.; Riede M.; Johnston M. B.; Friend R. H.; Snaith H. J. Optical properties and limiting photocurrent of thin-film perovskite solar cells. Energy Environ. Sci. 2015, 8, 602–609. 10.1039/c4ee03224a. [DOI] [Google Scholar]
- Stranks S. D.; Eperon G. E.; Grancini G.; Menelaou C.; Alcocer M. J. P.; Leijtens T.; Herz L. M.; Petrozza A.; Snaith H. J. Electron-Hole Diffusion Lengths Exceeding 1 Micrometer in an Organometal Trihalide Perovskite Absorber. Science 2013, 342, 341–344. 10.1126/science.1243982. [DOI] [PubMed] [Google Scholar]
- Wang Y.; Zhang Y.; Zhang P.; Zhang W. High intrinsic carrier mobility and photon absorption in the perovskite CH3NH3PbI3. Phys. Chem. Chem. Phys. 2015, 17, 11516–11520. 10.1039/c5cp00448a. [DOI] [PubMed] [Google Scholar]
- Miyata A.; Mitioglu A.; Plochocka P.; Portugall O.; Wang J. T.-W.; Stranks S. D.; Snaith H. J.; Nicholas R. J. Direct measurement of the exciton binding energy and effective masses for charge carriers in organic–inorganic tri-halide perovskites. Nat. Phys. 2015, 11, 582–587. 10.1038/nphys3357. [DOI] [Google Scholar]
- Anaya M.; Correa-Baena J. P.; Lozano G.; Saliba M.; Anguita P.; Roose B.; Abate A.; Steiner U.; Grätzel M.; Calvo M. E.; Hagfeldt A.; Míguez H. Optical analysis of CH3NH3SnxPb1-xI3 absorbers: a roadmap for perovskite-on-perovskite tandem solar cells. J. Mater. Chem. A 2016, 4, 11214–11221. 10.1039/c6ta04840d. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Eperon G. E.; Stranks S. D.; Menelaou C.; Johnston M. B.; Herz L. M.; Snaith H. J. Formamidinium lead trihalide: a broadly tunable perovskite for efficient planar heterojunction solar cells. Energy Environ. Sci. 2014, 7, 982–988. 10.1039/c3ee43822h. [DOI] [Google Scholar]
- Li J.; Janardan D.; Oleksandra S.; Marion A. F.; Hans K.; Markus F.; Christof S.; Bert S.; Justus J.; Daniel M. T.; Antonio A.; Rahim M.; Eva U. 20.8% Slot-Die Coated MAPbI3 Perovskite Solar Cells by Optimal DMSO-Content and Age of 2-ME Based Precursor Inks. Adv. Energy Mater. 2021, 11, 2003460. [Google Scholar]
- Alanazi A. Q.; Kubicki D. J.; Prochowicz D.; Alharbi E. A.; Bouduban M. E. F.; Jahanbakhshi F.; Mladenović M.; Milić J. V.; Giordano F.; Ren D.; Alyamani A. Y.; Albrithen H.; Albadri A.; Alotaibi M. H.; Moser J.-E.; Zakeeruddin S. M.; Rothlisberger U.; Emsley L.; Grätzel M. Atomic-Level Microstructure of Efficient Formamidinium-Based Perovskite Solar Cells Stabilized by 5-Ammonium Valeric Acid Iodide Revealed by Multi-Nuclear and Two-Dimensional Solid-State NMR. J. Am. Chem. Soc. 2019, 141, 17659–17669. 10.1021/jacs.9b07381. [DOI] [PubMed] [Google Scholar]
- Saliba M.; Matsui T.; Domanski K.; Seo J.-Y.; Ummadisingu A.; Zakeeruddin S. M.; Correa-Baena J.-P.; Tress W. R.; Abate A.; Hagfeldt A.; Gratzel M. Incorporation of rubidium cations into perovskite solar cells improves photovoltaic performance. Science 2016, 354, 206–209. 10.1126/science.aah5557. [DOI] [PubMed] [Google Scholar]
- Jeon N. J.; Na H.; Jung E. H.; Yang T.-Y.; Lee Y. G.; Kim G.; Shin H.-W.; Seok S. I.; Lee J.; Seo J. A fluorene-terminated hole-transporting material for highly efficient and stable perovskite solar cells. Nat. Energy 2018, 3, 682–689. 10.1038/s41560-018-0200-6. [DOI] [Google Scholar]
- Saliba M.; Matsui T.; Seo J.-Y.; Domanski K.; Correa-Baena J.-P.; Nazeeruddin M. K.; Zakeeruddin S. M.; Tress W.; Abate A.; Hagfeldt A.; Grätzel M. Cesium-containing triple cation perovskite solar cells: improved stability, reproducibility and high efficiency. Energy Environ. Sci. 2016, 9, 1989–1997. 10.1039/c5ee03874j. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mann D. S.; Patil P.; Kim D.-H.; Kwon S.-N.; Na S.-I. Boron nitride-incorporated NiOx as a hole transport material for high-performance p-i-n planar perovskite solar cells. J. Power Sources 2020, 477, 228738. 10.1016/j.jpowsour.2020.228738. [DOI] [Google Scholar]
- Kim H.-S.; Park N.-G. Parameters Affecting I–V Hysteresis of CH3NH3PbI3 Perovskite Solar Cells: Effects of Perovskite Crystal Size and Mesoporous TiO2 Layer. J. Phys. Chem. Lett. 2014, 5, 2927–2934. 10.1021/jz501392m. [DOI] [PubMed] [Google Scholar]
- Yang D.; Zhou X.; Yang R.; Yang Z.; Yu W.; Wang X.; Li C.; Liu S.; Chang R. P. H. Surface optimization to eliminate hysteresis for record efficiency planar perovskite solar cells. Energy Environ. Sci. 2016, 9, 3071–3078. 10.1039/c6ee02139e. [DOI] [Google Scholar]
- Jena A. K.; Numata Y.; Ikegami M.; Miyasaka T. Role of spiro-OMeTAD in performance deterioration of perovskite solar cells at high temperature and reuse of the perovskite films to avoid Pb-waste. J. Mater. Chem. A 2018, 6, 2219–2230. 10.1039/c7ta07674f. [DOI] [Google Scholar]
- Jena A. K.; Ikegami M.; Miyasaka T. Severe Morphological Deformation of Spiro-OMeTAD in (CH3NH3)PbI3 Solar Cells at High Temperature. ACS Energy Lett. 2017, 2, 1760–1761. 10.1021/acsenergylett.7b00582. [DOI] [Google Scholar]
- Huang W.-J.; Huang P.-H.; Yang S.-H. PCBM doped with fluorene-based polyelectrolytes as electron transporting layers for improving the performance of planar heterojunction perovskite solar cells. Chem. Commun. 2016, 52, 13572–13575. 10.1039/c6cc07062k. [DOI] [PubMed] [Google Scholar]
- Yang Z.; Chueh C.-C.; Liang P.-W.; Crump M.; Lin F.; Zhu Z.; Jen A. K.-Y. Effects of formamidinium and bromide ion substitution in methylammonium lead triiodide toward high-performance perovskite solar cells. Nano Energy 2016, 22, 328–337. 10.1016/j.nanoen.2016.02.033. [DOI] [Google Scholar]
- Hu L.; Sun K.; Wang M.; Chen W.; Yang B.; Fu J.; Xiong Z.; Li X.; Tang X.; Zang Z.; Zhang S.; Sun L.; Li M. Inverted Planar Perovskite Solar Cells with a High Fill Factor and Negligible Hysteresis by the Dual Effect of NaCl-Doped PEDOT: PSS. ACS Appl. Mater. Interfaces 2017, 9, 43902–43909. 10.1021/acsami.7b14592. [DOI] [PubMed] [Google Scholar]
- Al-Ashouri A.; Köhnen E.; Li B.; Magomedov A.; Hempel H.; Caprioglio P.; Márquez J. A.; Morales Vilches A. B.; Kasparavicius E.; Smith J. A.; Phung N.; Menzel D.; Grischek M.; Kegelmann L.; Skroblin D.; Gollwitzer C.; Malinauskas T.; Jošt M.; Matič G.; Rech B.; Schlatmann R.; Topič M.; Korte L.; Abate A.; Stannowski B.; Neher D.; Stolterfoht M.; Unold T.; Getautis V.; Albrecht S. Monolithic perovskite/silicon tandem solar cell with >29% efficiency by enhanced hole extraction. Science 2020, 370, 1300–1309. 10.1126/science.abd4016. [DOI] [PubMed] [Google Scholar]
- Mantilla-Perez P.; Feurer T.; Correa-Baena J.-P.; Liu Q.; Colodrero S.; Toudert J.; Saliba M.; Buecheler S.; Hagfeldt A.; Tiwari A. N.; Martorell J. Monolithic CIGS–Perovskite Tandem Cell for Optimal Light Harvesting without Current Matching. ACS Photonics 2017, 4, 861–867. 10.1021/acsphotonics.6b00929. [DOI] [Google Scholar]
- Zhang J.; Mao W.; Hou X.; Duan J.; Zhou J.; Huang S.; Ou-Yang W.; Zhang X.; Sun Z.; Chen X. Solution-processed Sr-doped NiOx as hole transport layer for efficient and stable perovskite solar cells. Sol. Energy 2018, 174, 1133–1141. 10.1016/j.solener.2018.10.004. [DOI] [Google Scholar]
- Hu L.; Li M.; Yang K.; Xiong Z.; Yang B.; Wang M.; Tang X.; Zang Z.; Liu X.; Li B.; Xiao Z.; Lu S.; Gong H.; Ouyang J.; Sun K. PEDOT:PSS monolayers to enhance the hole extraction and stability of perovskite solar cells. J. Mater. Chem. A 2018, 6, 16583–16589. 10.1039/c8ta05234d. [DOI] [Google Scholar]
- Ahn S.; Park M. H.; Jeong S. H.; Kim Y. H.; Park J.; Kim S.; Kim H.; Cho H.; Wolf C.; Pei M.; Yang H.; Lee T. W. Fine Control of Perovskite Crystallization and Reducing Luminescence Quenching Using Self-Doped Polyaniline Hole Injection Layer for Efficient Perovskite Light-Emitting Diodes. Adv. Funct. Mater. 2019, 29, 1807535. 10.1002/adfm.201807535. [DOI] [Google Scholar]
- You J.; Meng L.; Song T.-B.; Guo T.-F.; Yang Y.; Chang W.-H.; Hong Z.; Chen H.; Zhou H.; Chen Q.; Liu Y.; De Marco N.; Yang Y. Improved air stability of perovskite solar cells via solution-processed metal oxide transport layers. Nat. Nanotechnol. 2016, 11, 75–81. 10.1038/nnano.2015.230. [DOI] [PubMed] [Google Scholar]
- Ryu J.; Yoon S.; Park J.; Jeong S. M.; Kang D.-W. Fabrication of nickel oxide composites with carbon nanotubes for enhanced charge transport in planar perovskite solar cells. Appl. Surf. Sci. 2020, 516, 146116. 10.1016/j.apsusc.2020.146116. [DOI] [Google Scholar]
- Wang K.-C.; Shen P.-S.; Li M.-H.; Chen S.; Lin M.-W.; Chen P.; Guo T.-F. Low-Temperature Sputtered Nickel Oxide Compact Thin Film as Effective Electron Blocking Layer for Mesoscopic NiO/CH3NH3PbI3 Perovskite Heterojunction Solar Cells. ACS Appl. Mater. Interfaces 2014, 6, 11851–11858. 10.1021/am503610u. [DOI] [PubMed] [Google Scholar]
- Chen W.; Liu F.-Z.; Feng X.-Y.; Djurišić A. B.; Chan W. K.; He Z.-B. Cesium Doped NiOx as an Efficient Hole Extraction Layer for Inverted Planar Perovskite Solar Cells. Adv. Energy Mater. 2017, 7, 1700722. 10.1002/aenm.201700722. [DOI] [Google Scholar]
- Zhu Z.; Bai Y.; Zhang T.; Liu Z.; Long X.; Wei Z.; Wang Z.; Zhang L.; Wang J.; Yan F.; Yang S. High-Performance Hole-Extraction Layer of Sol-Gel-Processed NiO Nanocrystals for Inverted Planar Perovskite Solar Cells. Angew. Chem., Int. Ed. 2014, 53, 12571–12575. 10.1002/anie.201405176. [DOI] [PubMed] [Google Scholar]
- Zhang H.; Cheng J.; Lin F.; He H.; Mao J.; Wong K. S.; Jen A. K.-Y.; Choy W. C. H. Pinhole-Free and Surface-Nanostructured NiOx Film by Room-Temperature Solution Process for High-Performance Flexible Perovskite Solar Cells with Good Stability and Reproducibility. ACS Nano 2016, 10, 1503–1511. 10.1021/acsnano.5b07043. [DOI] [PubMed] [Google Scholar]
- Chen W.; Wu Y.; Liu J.; Qin C.; Yang X.; Islam A.; Cheng Y.-B.; Han L. Hybrid interfacial layer leads to solid performance improvement of inverted perovskite solar cells. Energy Environ. Sci. 2015, 8, 629–640. 10.1039/c4ee02833c. [DOI] [Google Scholar]
- Seo S.; Park I. J.; Kim M.; Lee S.; Bae C.; Jung H. S.; Park N. G.; Kim J. Y.; Shin H. An ultra-thin, un-doped NiO hole transporting layer of highly efficient (16.4%) organic–inorganic hybrid perovskite solar cells. Nanoscale 2016, 8, 11403–11412. 10.1039/c6nr01601d. [DOI] [PubMed] [Google Scholar]
- Wang T.; Cheng Z.; Zhou Y.; Liu H.; Shen W. Highly efficient and stable perovskite solar cells via bilateral passivation layers. J. Mater. Chem. A 2019, 7, 21730–21739. 10.1039/c9ta08084h. [DOI] [Google Scholar]
- Chen W.; Wang Y.; Pang G.; Koh C. W.; Djurišić A. B.; Wu Y.; Tu B.; Liu F. z.; Chen R.; Woo H. Y.; Guo X.; He Z. Conjugated Polymer-Assisted Grain Boundary Passivation for Efficient Inverted Planar Perovskite Solar Cells. Adv. Funct. Mater. 2019, 29, 1808855. 10.1002/adfm.201808855. [DOI] [Google Scholar]
- Mali S. S.; Kim H.; Kim H. H.; Shim S. E.; Hong C. K. Nanoporous p-type NiOx electrode for p-i-n inverted perovskite solar cell toward air stability. Mater. Today 2018, 21, 483–500. 10.1016/j.mattod.2017.12.002. [DOI] [Google Scholar]
- Cerc Korošec R.; Bukovec P.; Pihlar B.; Vuk A. Š.; Orel B.; Dražič G. Preparation and structural investigations of electrochromic nanosized NiOx films made via the sol–gel route. Solid State Ionics 2003, 165, 191–200. [Google Scholar]
- Paulose R.; Mohan R.; Parihar V. Nanostructured nickel oxide and its electrochemical behavior–A brief review. Nano-Struct. Nano-Objects 2017, 11, 102–111. 10.1016/j.nanoso.2017.07.003. [DOI] [Google Scholar]
- Xia X. H.; Tu J. P.; Zhang J.; Wang X. L.; Zhang W. K.; Huang H. Electrochromic properties of porous NiO thin films prepared by a chemical bath deposition. Sol. Energy Mater. Sol. Cells 2008, 92, 628–633. 10.1016/j.solmat.2008.01.009. [DOI] [Google Scholar]
- Pramanik P.; Bhattacharya S. A chemical method for the deposition of nickel oxide thin films. J. Electrochem. Soc. 1990, 137, 3869–3870. 10.1149/1.2086316. [DOI] [Google Scholar]
- Han S.-Y.; Lee D.-H.; Chang Y.-J.; Ryu S.-O.; Lee T.-J.; Chang C.-H. The Growth Mechanism of Nickel Oxide Thin Films by Room-Temperature Chemical Bath Deposition. J. Electrochem. Soc. 2006, 153, C382–C386. 10.1149/1.2186767. [DOI] [Google Scholar]
- Desai J. D.; Min S.-K.; Jung K.-D.; Joo O.-S. Spray pyrolytic synthesis of large area NiOx thin films from aqueous nickel acetate solutions. Appl. Surf. Sci. 2006, 253, 1781–1786. 10.1016/j.apsusc.2006.03.009. [DOI] [Google Scholar]
- Chen P.-C.; Yang S.-H. Potassium-Doped Nickel Oxide as the Hole Transport Layer for Efficient and Stable Inverted Perovskite Solar Cells. ACS Appl. Energy Mater. 2019, 2, 6705–6713. 10.1021/acsaem.9b01200. [DOI] [Google Scholar]
- Zheng J.; Hu L.; Yun J. S.; Zhang M.; Lau C. F. J.; Bing J.; Deng X.; Ma Q.; Cho Y.; Fu W.; Chen C.; Green M. A.; Huang S.; Ho-Baillie A. W. Y. Solution-Processed, Silver-Doped NiOx as Hole Transporting Layer for High-Efficiency Inverted Perovskite Solar Cells. ACS Appl. Energy Mater. 2018, 1, 561–570. 10.1021/acsaem.7b00129. [DOI] [Google Scholar]
- Chandrasekhar P. S.; Seo Y.-H.; Noh Y.-J.; Na S.-I. Room temperature solution-processed Fe doped NiOx as a novel hole transport layer for high efficient perovskite solar cells. Appl. Surf. Sci. 2019, 481, 588–596. 10.1016/j.apsusc.2019.03.164. [DOI] [Google Scholar]
- Prochowicz D.; Tavakoli M. M.; Solanki A.; Goh T. W.; Pandey K.; Sum T. C.; Saliba M.; Yadav P. Understanding the effect of chlorobenzene and isopropanol anti-solvent treatments on the recombination and interfacial charge accumulation in efficient planar perovskite solar cells. J. Mater. Chem. A 2018, 6, 14307–14314. 10.1039/c8ta03782e. [DOI] [Google Scholar]
- Zheng G.; Zhu C.; Ma J.; Zhang X.; Tang G.; Li R.; Chen Y.; Li L.; Hu J.; Hong J.; Chen Q.; Gao X.; Zhou H. Manipulation of facet orientation in hybrid perovskite polycrystalline fFilms by cation cascade. Nat. Commun. 2018, 9, 2793. 10.1038/s41467-018-05076-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu Y.; Xie F.; Chen H.; Yang X.; Su H.; Cai M.; Zhou Z.; Noda T.; Han L. Thermally Stable MAPbI3 Perovskite Solar Cells with Efficiency of 19.19% and Area over 1 cm2 Achieved by Additive Engineering. Adv. Mater. 2017, 29, 1701073. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.







