Next Article in Journal
High-Entropy Oxides: Advanced Research on Electrical Properties
Next Article in Special Issue
Partial Replacement of Dimethylformamide with Less Toxic Solvents in the Fabrication Process of Mixed-Halide Perovskite Films
Previous Article in Journal
Properties of Cold Spray Coatings for Restoration of Worn-Out Contact Wires
Previous Article in Special Issue
Cesium-Containing Triple Cation Perovskite Solar Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Combustion Processed Nickel Oxide and Zinc Doped Nickel Oxide Thin Films as a Hole Transport Layer for Perovskite Solar Cells

by
Ponmudi Selvan Thiruchelvan
1,
Chien-Chih Lai
1,2 and
Chih-Hung Tsai
2,*
1
Department of Physics, National Dong Hwa University, Hualien 97401, Taiwan
2
Department of Opto-Electronic Engineering, National Dong Hwa University, Hualien 97401, Taiwan
*
Author to whom correspondence should be addressed.
Coatings 2021, 11(6), 627; https://doi.org/10.3390/coatings11060627
Submission received: 1 April 2021 / Revised: 18 May 2021 / Accepted: 22 May 2021 / Published: 24 May 2021
(This article belongs to the Special Issue Organic and Hybrid Thin Films for Solar Cells)

Abstract

:
Combustion processed nickel oxide (NiOx) thin film is considered as an alternative to the sol-gel processed hole transport layer for perovskite solar cells (PSCs). In this paper, NiOx thin film was prepared by the solution–combustion process at 250 °C, a temperature lower than the actual reaction temperature. Furthermore, the properties of the NiOx hole transport layer (HTL) in PSCs were enhanced by the incorporation of zinc (Zn) in NiOx thin films. X-ray diffraction and X-ray photoelectron spectroscopy results revealed that the formation of NiOx was achieved at lower annealing temperature, which confirms the process of the combustion reaction. The electrical conductivity was greatly improved with Zn doping into the NiOx crystal lattice. Better photoluminescence (PL) quenching, and low PL lifetime decay were responsible for better charge separation in 5% Zn doped NiOx, which results in improved device performance of PSCs. The maximum power conversion efficiency of inverted PSCs made with pristine NiOx and 5% Zn-NiOx as the HTL was 13.62% and 14.87%, respectively. Both the devices exhibited better stability than the PEDOT:PSS (control) device in an ambient condition.

1. Introduction

Researchers have promised that the perovskite solar cells (PSCs) could soon be released as a commercial product in the next few years [1,2]. PSCs have become a hot topic among researchers because of their superior optoelectronic properties [3,4,5,6,7]. The ease of device fabrication kept this device in the limelight among other next-generation solar cells [8]. In inverted PSCs, the most commonly used HTL is poly(3,4-ethylene dioxythiophene) polystyrene sulfonate (PEDOT:PSS); it can exhibit better device performance at lower processing temperature, but the stability of the device is worse because of its hygroscopic and acidic nature [9]. Replacing the organic PEDOT:PSS as hole transport layer (HTL) in perovskite solar cells with inorganic p-type metal oxide materials is an effective way to improve the device’s performance and stability [10]. The basic requirement of the HTL is to be highly transparent throughout the solar spectra, good electrical conductivity, and a high work function for effective charge transport [11]. Among the various metal oxides, nickel oxide (NiOx) is an ideal candidate for this special purpose [11,12,13,14,15]. In physical vapor deposition techniques, these metal oxides require special processing conditions, including a vacuum environment and heat treatment, etc. [16]. In contrast, solution-processed NiOx has a variety of options that does not require any special processing conditions [17]. The preparation of the NiOx HTL by the sol-gel method requires it to be annealed at a high temperature above 300 °C to achieve high performance from the PSCs [18,19]. This high-temperature process increases the cost of device fabrication at an industrial scale and restricts the substrate options [20]. Low-temperature processing or crystallization of metal oxide thin films has become popular research in the last few years [21].
During the solution process, the temperature of NiOx must be raised to a minimum of 275 °C to commence converting the Ni precursor into NiOx [22,23]. Another means of implementing the sol-gel process is to employ combustion chemistry to create solution-processed transition metal oxide (TMO) films with high crystallinity without using high temperatures [24,25,26,27]. The unique properties of self-generated energy and exothermic reactions permit the use of lower processing temperatures, replacing sol-gel endothermic reactions that require high levels of thermal energy to enable the formation of TMO lattices and the removal of any organic impurity. Combustion chemistry has been employed for the successful preparation of a variety of TMO thin films and at lower temperatures than sol-gel technique. In PSCs, the combustion processed NiOx was first reported by Bai et al. It was annealed at 175 °C, with glycine as a fuel [28]. Later, Jen et al. replaced glycine with acetylacetone and developed NiOx HTL at 150 °C [24]. Chang et al. used a similar method but annealed at 250 °C and achieved a PCE of 20.2% for mixed cation PSCs [29]. Recently, Cai et al. studied the effect of various concentrations of fuel in the combustion processed NiOx HTL for PSCs [30]. Zhang et al. improved the efficiency with a two-step coating of mixed perovskites [31]; however, the intrinsic electrical conductivity of NiOx is very low, which restricts the device’s performance. Metal ion doping into NiO can improve its electrical conductivity even at lower annealing temperatures. This has been successfully used in sol-gel processed NiOx doped with various metal ions such as Ag [32], Cs [23], Y [33], K [34], Co [35], etc. Even in the combustion reaction, Cu doped NiOx thin film was successfully developed as the HTL in inverted PSCs [24]. Influenced by this idea, we adopted this technique to prepare zinc doped NiOx thin film as the HTL in PSCs.
Here, we report the successful development of NiOx and Zn doped NiOx at a relatively low annealing temperature as the HTL for inverted PSCs. It was achieved by utilizing combustion chemistry, which significantly reduces its actual reaction temperature compared with traditional sol-gel methods. The effect of Zn doping in NiOx was studied by investigating the structural, morphological, and optoelectronic properties and the carrier dynamics of the films. The photovoltaic performance was analyzed and the average PCE of NiOx and 5% Zn doped NiOx was 13.52% and 14.67%, respectively. Overall, the device performance was significantly improved with Zn doping in the NiOx HTL.

2. Materials and Methods

2.1. Synthesis of NiOx and Zn Doped NiOx Thin Films

Preparation of the NiO precursor solution took place as previously described [25] with certain adaptations. Briefly, 0.145 g of nickel nitrate hexahydrate underwent dissolution in 5 mL of 2-methoxyethanol; this solution underwent stirring for 30 min until complete dissolution of the precursor was achieved. This solution was then supplemented with 50 µL of acetylacetonate and stirring was maintained overnight. Following this, a polytetrafluoroethylene (PTFE) syringe filter was used to filter the solution. To deposit the thin film, the precursor solution underwent spin coating onto a clean fluorine-doped tin oxide (FTO) substrate for 45 s at 2500 rpm. A hotplate was then used to anneal the spin-coated films in the air for one hour at 250 °C. These prepared films were then given UV-ozone treatment for five minutes prior to being placed in the glovebox. To make the doping successful, zinc nitrate hexahydrate was used to replace the variety of mole percentages of nickel precursors.

2.2. PSC Device Fabrication

An ultrasonic bath was employed to clean the laser patterned FTO substrates in sequence with detergent, acetone, and IPA for 10 min. Subsequently, drying of the substrate occurred employing a nitrogen gun and then exposed to ultraviolet (UV)-ozone treatment (15 min). A spin coater was used to deposit the NiOx and Zn-NiOx thin films with annealing occurring at 250 °C as previously described. Following cooling, these substrates were then placed in the glovebox. Following this, production of a perovskite active layer occurred by dissolving 1.4 M methylammonium iodide (MAI) and 1.4 M lead iodide (PbI2) in γ-butyrolactone (GBL) and dimethyl sulfoxide (DMSO) (7:3). This solution was then subjected to overnight stirring inside the glove box at 60 °C. A 0.45-µm PTFE syringe filter was used to filter the solution before deposition. The solution underwent spin coating onto FTO/NiOx substrates (1000 rpm/15 s and 5000 rpm/25 s). Toluene was employed as an anti-solvent and was added precisely five seconds prior to completing the second stage. Following this, the substrates were annealed for 10 min at 100 °C. Preparation of the electron transport layer (ETL) was achieved in the following manner: 20 mg/mL of phenyl-C61-butyric acid methyl ester (PCBM) in 1,4-dichlorobenzene (DCB) underwent spin-coating onto the substrates at 2000 rpm for 30 s, and a saturated solution of bathocuproine (BCP) in methanol was dynamically spin-coated at 6000 rpm for 30 s. Lastly, thermal evaporation of an Ag electrode was achieved employing a metal mask.

2.3. Characterization

Characterization of the NiOx and Zn-NiOx thin films was undertaken implying X-ray diffraction (XRD) techniques using a Rigaku D/Max-2500V system (Rigaku Corp, Tokyo, Japan). The surface morphology and topography underwent inspection using a field emission scanning electron microscope (FESEM) (morphology) (JEOL JSM-7000F, JEOL Inc., Tokyo, Japan) and an atomic force microscope (AFM) (AutoProbe CP, Thermomicroscopes, Sunnyvale, CA, USA) (topography). Elemental composition analysis was undertaken using X-ray photoelectron spectroscopy (XPS) (Thermo K-Alpha, Thermo Fisher Scientific, Inc., Waltham, MA, USA). Measurement of UV-visible absorption and transmission spectra was undertaken employing a UV-Vis spectrophotometer (Hitachi U 3900, Tokyo, Japan). Measurements of steady-state photoluminescence (PL) and time-resolved photoluminescence (TRPL) were undertaken employing a solid-state laser (λ = 405 nm, PDL 800-D, PICOQUANT, Berlin, Germany) equipped with a laser scanning confocal microscope and a spectrometer equipped with a thermoelectrically cooled charge-coupled detector (for PL) and a time-correlated single-photon counting detector (TCSPC) (for TRPL). Analysis of the PSC device performance was undertaken employing current density–voltage (JV) curves, external quantum efficiency (EQE), and electrochemical impedance spectroscopy (EIS). Measurement of PSC JV characteristics was undertaken using a 1 sun illumination (AM 1.5 G spectrum) employing a 550 W Xenon lamp solar simulator (Sun 3000 Class AAA, Abet Technologies, Milford, MA, USA).

3. Results and Discussion

As mentioned, the NiOx and Zn doped NiOx thin films were prepared using the combustion spin-coating technique, and the films were annealed at a lower temperature (250 °C) than the sol-gel technique. The inverted PSCs with an architecture of FTO/NiOx/MAPbI3/PCBM/BCP/Ag were fabricated, as shown in Figure 1a. Figure 1b,c show the energy level diagram of inverted PSCs and the photograph of the fabricated device. The current density–voltage (JV) curves of the PSCs made with NiOx and Zn doped NiOx HTL are shown in Figure 2a. The device made with 5% Zn-NiOx exhibited the best PCE of 14.87% versus pristine NiOx (13.62%). The detailed device parameters of the PSCs made with both the NiOx and Zn doped NiOx thin films annealed at 250 °C are summarized in Table 1. The results were reproducible and the average values were calculated from eight best devices from different batches. The devices performed well with NiOx HTL annealed at a relatively lower temperature. The current density (Jsc) of the pristine NiOx device was improved with an increase in Zn doping and may be due to an improvement in the electrical conductivity of HTL. The lower doping concentration of 2% Zn-NiOx exhibited relatively similar efficiency with a minimal increase in Jsc. The device made with 5% Zn-NiOx showed improved Jsc and minimal reduction in FF and Voc. There was no significant change in Voc perhaps because there were no major changes in valence band edge position after Zn doping with lower concentration. The Voc and FF were decreased with further increasing of Zn concentration (10%) into NiOx HTL. It may be due to the higher Zn2+ dopant concentration induce the transformation of p-type to n-type NiOx, which results in poor device performance [36]. The JV characteristics revealed that the current flow across the device was increased due to the Zn doping into NiOx. The external quantum efficiency (EQE) spectra of the devices made using NiOx and Zn doped NiOx HTL are shown in Figure 2b. The devices made with NiOx and Zn doped NiOx films displayed satisfactory results throughout the entire visible region. In comparison, a similar kind of EQE curve was observed in all the devices. The increment in the EQE value for the 5% Zn-NiOx device versus the pristine NiOx device indicates the improvement in device performance. These results agreed with the JV characteristics results. Therefore, the pristine NiOx and 5% Zn-NiOx were chosen for further detailed studies.
Apart from the performance provided by the HTL, it has a strong impact on the stability of PSCs. This study performed an investigation on the stability of unencapsulated devices stored in dark conditions at room temperature with about 40% of relative humidity. Figure 3a,b show the stability graph and the photograph of PSCs based on PEDOT:PSS, NiOx, and 5% Zn-NiOx HTLs, respectively. It can be seen that the device made with PEDOT:PSS degrades fast and lost its total efficiency in 4 days. The fast degradation of perovskite happened at the active area of PEDOT:PSS device was due to the reaction of Ag electrodes with volatile degradation products from the perovskite and ion migration. The acidic and hygroscopic nature of PEDOT:PSS degrades the MAPbI3 into PbI2 and releases its by-products in the form of I2, CH3I, and HI, which can readily react with Ag to form AgI [37]. By contrast, the rate of degradation is usually very slow in NiOx based PSCs. Both NiOx and 5% Zn-NiOx based PSCs have retained 80% of their initial efficiency after 108 h.
The XRD pattern of NiOx and 5% Zn-NiOx thin films on glass substrate and annealed at 250 °C is shown in Figure 4. The cubic structure of NiO has peaked at 37.3°, 43.3°, and 62.9°, corresponding to the hkl value of (111), (002), and (022), respectively. This can be observed in both NiOx and 5% Zn-NiOx thin films. It clearly shows that thin films had poor polycrystalline nature but significantly higher crystallinity than the film annealed at 250 °C in the sol-gel process in previous work [38], and similar kinds of results were reported [29,33]. This result shows that the formation of NiOx occurred below the actual reaction temperatures. It confirms the process of the combustion reaction.
X-ray photoelectron spectroscopy (XPS) was used to analyze the elemental composition of the NiOx and 5% Zn-NiOx thin films annealed at 250 °C. Figure 5a,b illustrate the XPS spectra for the Ni 2p and O 1s core levels of NiOx and 5% Zn-NiOx thin films. The NiOx peaks primarily comprise Ni2+ and Ni3+ states. At 861 eV, there is a satellite peak attributable to the NiOx shake-up process. The deconvoluted peak at 854 eV corresponds with NiO (Ni2+), 856 eV with Ni2O3 (Ni3+), and 857 eV with NiOOH [39]. Similarly, the O 1s spectra has three peaks, the deconvoluted peak at 529 eV corresponding with NiO (Ni2+), 531 eV with Ni2O3 (Ni3+), and 533 eV with NiOOH. Generally, Ni3+ states cause p-type conductivity, indicating that quasi-localized holes have been formed around the lattice position’s Ni2+ vacancies [40]. There is a higher Ni3+ state with 5% Zn-NiOx versus pristine NiOx as seen in the spectra of Ni 2p and O 1s. This was due to the replacement of the Ni atom with Zn in the crystal lattice. Zn incorporation could alter the electronic structure of NiOx. The increase in the Ni3+ state denotes the presence of more Ni vacancies. The higher Ni3+/Ni2+ ratio in 5% Zn-NiOx indicates the reduction in ionization energy of the Ni vacancies and increased hole density [41]. The successful incorporation of Zn in NiOx can be seen in Figure 5c,d. The divalent state of Zn peaks of 2p3/2 and 2p1/2 were observed at 1021 eV and 1045 eV, respectively.
The surface morphology of NiOx and 5% Zn-NiOx thin films annealed at 250 °C was analyzed by field emission scanning electron microscopy (FESEM). The top view of the FESEM image and energy-dispersive X-ray spectroscopy (EDS) results for NiOx and 5% Zn-NiOx thin films are shown in Figure 6a,b, respectively. Both films exhibit similar morphology and better coverage without any voids on FTO substrate. A fine layer of these thin films is clearly seen on the FTO, and the 5% Zn doped NiOx film looks somewhat smoother than the pristine film. This observation was confirmed by AFM results. EDS results confirmed the presence of Zn in NiOx thin films. The atomic composition ratio was closely related to stoichiometry. Figure 6c shows the cross-sectional FESEM image of the device. The device was fabricated with an architecture of FTO/NiOx/MAPbI3/PCBM/BCP/Ag. All layers can be well distinguished from the cross-sectional FESEM image.
Figure 7a,b show the AFM images of NiOx and 5% Zn-NiOx thin films annealed at 250 °C. The films exhibited similar smooth and compact surface topography on FTO substrates for both NiOx and 5% Zn-NiOx. The root-mean-squared (RMS) roughness values of NiOx and 5% Zn-NiOx are 37.1 nm and 33.9 nm, respectively. The roughness of the film decreased with Zn doping because of the tendency of ZnO to form a crystalline nature at the lower annealing temperature [42]. This result is well-matched with other literature reports [41,43].
Figure 8a illustrates the transmittance spectra for NiOx and 5% Zn-NiOx thin films that underwent annealing at 250 °C. Each film showed a high transmittance level within the visible spectrum range, with average values being above 80%. Figure 8b illustrates the graph of absorption coefficient as a function of excitation energy; calculation of the optical bandgap value was undertaken through an additional plot of the linear portion of (αhν)2 to the x-axis employing a Tauc plot [44]. This bandgap value (Eg) was around 3.95 eV both for NiOx and 5% Zn-NiOx thin films. Very high HTL transmittance is required for minimization of incident light loss for enhancement of the light that reaches the perovskite layer. Any variations in optical properties when Zn is incorporated into NiOx are minimal due to lower film thicknesses (~20 nm).
The impact of Zn doping on NiOx thin-film electrical conductivity was examined by measuring the current–voltage (IV) characteristics of the hole-only device (FTO/NiOx/Ag). Figure 9 shows the IV curves for the hole-only devices based on NiOx and 5% Zn-NiOx thin films that underwent annealing at 250 °C. This demonstrates that there are improvements in electrical conductivity when Zn doping is applied as the dopants have a high level of affinity with electrons.
Charge extraction and non-radiative recombination at the perovskite and HTL interface was evaluated using steady-state photoluminescence (PL) and time-resolved photoluminescence (TRPL). Figure 10a,b show the PL spectrum and TRPL profile for perovskite films over NiOx and 5% Zn-NiOx thin films. The offset in the valence band edge plays a major role in charge recombination. Mismatches in the valence band edge between HTL and perovskite layer leads to more recombination at the interface and results in lowering of the Voc [45]. PL exhibited by the perovskite layer was significantly reduced by the HTLs. Noteworthy levels of PL quenching were shown with both types of film, demonstrating high levels of efficiency of charge extraction for Zn doped NiOx thin films. These findings accord with previous research [43]. The TRPL profile was measured employing time-correlated single-photon counting. Calculation of the PL lifetime decay was undertaken using tail fit with a biexponential decay function, as shown in Equation (1),
y = y 0 + A 1 e ( x x 0 ) / τ 1 + A 2 e ( x x 0 ) / τ 2
where τ1 represents the photogenerated excitons diffusing into defects, and τ2 representing intrinsic electron-hole recombination. The formula illustrated in Equation (2) was employed for the calculation of average PL lifetime values [46]; a detailed summary of these values can be found in Table 2.
τ a v g = ( Σ i A i τ i 2 ) / ( Σ i A i τ i )
This research employed EIS measurements to undertake further verification of NiOx and 5% Zn-NiOx thin film-based devices’ charge transport properties. Figure 11 illustrates the Nyquist plots for devices fabricated using NiOx and 5% Zn-NiOx thin films annealed at 250 °C. EIS measurement was taken at one solar illumination using a frequency range from 10 Hz to 1 MHz with an AC amplitude of 0.01 V, and providing a bias voltage that matched the open-circuit voltage. The impedance curve was observed at three frequency regions where the high-frequency (HF) regime correlates with charge transport resistance (Rct) at the HTL/perovskite interface, mid-frequency (MF) region is related to charge recombination, and low-frequency (LF) has been associated to iodide ion modulated recombination/injection process [47]. The curves were fitted with the equivalent circuit and the calculated Rct values are ~60.38 Ω and 45.62 Ω for NiOx and 5% Zn-NiOx devices, respectively. Typically, low impedance values are expected for better carrier transport functions [48,49]. The device fabricated using Zn doped NiOx had lower impedance values than those fabricated using pristine NiOx. It is clear that interface charge accumulation is suppressed, thus, improving exciton charge separation. Improved charge separation leads to increases in EQE and Jsc values. These findings clearly indicate that charge transport at the HTL-perovskite interfaces can be improved by doping NiOx with Zn, leading to a general improvement in the performance of NiOx HTL.

4. Conclusions

In this work, we investigated the effect of extrinsic Zn doping in NiOx HTL in inverted PSCs by the combustion technique. Comparatively, the devices made with both pristine and Zn doped NiOx films performed well with the annealing temperature of 250 °C, showing average PCE values of 13.62% and 14.87%, respectively. Both the devices have exhibited better stability than the PEDOT:PSS (control) device in an ambient condition. The crystallinity of both pristine and Zn doped NiOx films is better than the sol-gel derived thin film annealed at 250 °C. The XPS results revealed that the presence of compound formation below its reaction temperature and intensity of the Ni3+ peak (Ni vacancy) was higher in Zn doped NiOx than the NiOx film. The presence of Zn in NiOx was confirmed in XPS but not observed in XRD. Both films exhibited similar optical transparency, and the effect of Zn incorporation on optical properties is negligible. The electrical conductivity was greatly improved with Zn doping into the NiOx crystal lattice. The better PL quenching and low PL lifetime decay responsible for better charge separation were observed in Zn doped NiOx. A small impedance value from the Zn doped device showed less transport charge resistance than the pristine one. Overall, the NiOx thin films made by the combustion technique were the films annealed at 250 °C, which can work as HTL in inverted PSCs. The performance of combustion processed NiOx HTL was improved by Zn doping, showing better efficiency and overall superior performance than pristine NiOx films.

Author Contributions

Conceptualization, C.-H.T. and P.S.T.; methodology, C.-H.T., P.S.T. and C.-C.L.; validation, C.-H.T.; formal analysis, C.-H.T., P.S.T. and C.-C.L.; investigation, C.-H.T., P.S.T. and C.-C.L.; resources, C.-H.T.; data curation, C.-H.T., P.S.T. and C.-C.L.; writing—original draft preparation, C.-H.T. and P.S.T.; writing—review and editing, C.-H.T. and P.S.T.; supervision, C.-H.T.; project administration, C.-H.T.; funding acquisition, C.-H.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Ministry of Science and Technology (MOST) of Taiwan (Grant No. MOST 108-2221-E-259-011).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data in this work are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Perovskite Solar Out-Benches Rivals in 2021-IEEE Spectrum. Available online: https://spectrum.ieee.org/tech-talk/energy/renewables/oxford-pv-sets-new-record-for-perovskite-solar-cells (accessed on 18 March 2021).
  2. Perovskite Solar | Perovskite-Info. Available online: https://www.perovskite-info.com/perovskite-solar (accessed on 18 March 2021).
  3. 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. [Google Scholar] [CrossRef] [PubMed]
  4. Kulkarni, S.A.; Baikie, T.; Boix, P.P.; Yantara, N.; Mathews, N.; Mhaisalkar, S. Band-gap tuning of lead halide perovskites using a sequential deposition process. J. Mater. Chem. A 2014, 2, 9221–9225. [Google Scholar] [CrossRef] [Green Version]
  5. Noh, J.H.; Im, S.H.; Heo, J.H.; Mandal, T.N.; Seok, S. Il Chemical management for colorful, efficient, and stable inorganic-organic hybrid nanostructured solar cells. Nano Lett. 2013, 13, 1764–1769. [Google Scholar] [CrossRef] [PubMed]
  6. Stranks, S.D.; Eperon, G.E.; Grancini, G.; Menelaou, C.; Alcocer, M.J.P.; Leijtens, T.; Herz, L.M.; Petrozza, A.; Snaith, H.J. Electron-hole diffusion lengths exceeding 1 micrometer in an organometal trihalide perovskite absorber. Science 2013, 342, 341–344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Dong, Q.; Fang, Y.; Shao, Y.; Mulligan, P.; Qiu, J.; Cao, L.; Huang, J. Electron-hole diffusion lengths > 175 μm in solution-grown CH3NH3PbI3 single crystals. Science 2015, 347, 967–970. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Saliba, M.; Correa-Baena, J.-P.; Wolff, C.M.; Stolterfoht, M.; Phung, N.; Albrecht, S.; Neher, D.; Abate, A. How to make over 20% efficient perovskite solar cells in regular (n-i-p) and inverted (p-i-n) architectures. Chem. Mater. 2018, 30, 4193–4201. [Google Scholar] [CrossRef]
  9. Ionescu-Zanetti, C.; Mechler, A.; Carter, S.A.; Lal, R. Semiconductive Polymer Blends: Correlating Structure with Transport Properties at the Nanoscale. Adv. Mater. 2004, 16, 385–389. [Google Scholar] [CrossRef]
  10. Shin, S.S.; Lee, S.J.; Seok, S. Il Metal Oxide Charge Transport Layers for Efficient and Stable Perovskite Solar Cells. Adv. Funct. Mater. 2019, 29, 1900455. [Google Scholar] [CrossRef]
  11. Liu, T.; Chen, K.; Hu, Q.; Zhu, R.; Gong, Q. Inverted Perovskite Solar Cells: Progresses and Perspectives. Adv. Energy Mater. 2016, 6, 1600457. [Google Scholar] [CrossRef]
  12. Singh, R.; Singh, P.K.; Bhattacharya, B.; Rhee, H.-W. Review of current progress in inorganic hole-transport materials for perovskite solar cells. Appl. Mater. Today 2019, 14, 175–200. [Google Scholar] [CrossRef]
  13. Cetin, C.; Chen, P.; Hao, M.; He, D.; Bai, Y.; Lyu, M.; Yun, J.-H.; Wang, L. Inorganic p-Type Semiconductors as Hole Conductor Building Blocks for Robust Perovskite Solar Cells. Adv. Sustain. Syst. 2018, 2, 1800032. [Google Scholar] [CrossRef]
  14. Yin, X.; Guo, Y.; Xie, H.; Que, W.; Kong, L.B. Nickel Oxide as Efficient Hole Transport Materials for Perovskite Solar Cells. Sol. RRL 2019, 3, 1900001. [Google Scholar] [CrossRef]
  15. Ma, F.; Zhao, Y.; Li, J.; Zhang, X.; Gu, H.; You, J. Nickel oxide for inverted structure perovskite solar cells. J. Energy Chem. 2020, 52, 393–411. [Google Scholar] [CrossRef]
  16. Chen, W.; Wu, Y.; Tu, B.; Liu, F.; Djurišić, A.B.; He, Z. Inverted planar organic-inorganic hybrid perovskite solar cells with NiOx hole-transport layers as light-in window. Appl. Surf. Sci. 2018, 451, 325–332. [Google Scholar] [CrossRef]
  17. Seo, Y.H.; Cho, I.H.; Na, S.I. Investigation of sol-gel and nanoparticle-based NiOx hole transporting layer for high-performance planar perovskite solar cells. J. Alloys Compd. 2019, 797, 1018–1024. [Google Scholar] [CrossRef]
  18. Jin, Z.; Guo, Y.; Yuan, S.; Zhao, J.S.; Liang, X.M.; Qin, Y.; Zhang, J.P.; Ai, X.C. Modification of NiOx hole transport layer for acceleration of charge extraction in inverted perovskite solar cells. RSC Adv. 2020, 10, 12289–12296. [Google Scholar] [CrossRef] [Green Version]
  19. Lian, X.; Chen, J.; Shan, S.; Wu, G.; Chen, H. Polymer Modification on the NiOx Hole Transport Layer Boosts Open-Circuit Voltage to 1.19 V for Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2020, 12, 46340–46347. [Google Scholar] [CrossRef] [PubMed]
  20. Najafi, M.; Di Giacomo, F.; Zhang, D.; Shanmugam, S.; Senes, A.; Verhees, W.; Hadipour, A.; Galagan, Y.; Aernouts, T.; Veenstra, S.; et al. Highly Efficient and Stable Flexible Perovskite Solar Cells with Metal Oxides Nanoparticle Charge Extraction Layers. Small 2018, 14, 1–10. [Google Scholar] [CrossRef] [PubMed]
  21. Bretos, I.; Jiménez, R.; Ricote, J.; Calzada, M.L. Low-temperature crystallization of solution-derived metal oxide thin films assisted by chemical processes. Chem. Soc. Rev. 2018, 47, 291–308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Manders, J.R.; Tsang, S.-W.; Hartel, M.J.; Lai, T.-H.; Chen, S.; Amb, C.M.; Reynolds, J.R.; So, F. Solution-Processed Nickel Oxide Hole Transport Layers in High Efficiency Polymer Photovoltaic Cells. Adv. Funct. Mater. 2013, 23, 2993–3001. [Google Scholar] [CrossRef]
  23. 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. [Google Scholar] [CrossRef]
  24. Jung, J.W.; Chueh, C.C.; Jen, A.K.Y. 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. 2015, 27, 7874–7880. [Google Scholar] [CrossRef]
  25. Merzhanov, A.G. The chemistry of self-propagating high-temperature synthesis. J. Mater. Chem. 2004, 14, 1779–1786. [Google Scholar] [CrossRef]
  26. Deshpande, K.; Mukasyan, A.; Varma, A. Direct synthesis of iron oxide nanopowders by the combustion approach: Reaction mechanism and properties. Chem. Mater. 2004, 16, 4896–4904. [Google Scholar] [CrossRef]
  27. Epifani, M.; Melissano, E.; Pace, G.; Schioppa, M. Precursors for the combustion synthesis of metal oxides from the sol-gel processing of metal complexes. J. Eur. Ceram. Soc. 2007, 27, 115–123. [Google Scholar] [CrossRef]
  28. Bai, S.; Cao, M.; Jin, Y.; Dai, X.; Liang, X.; Ye, Z.; Li, M.; Cheng, J.; Xiao, X.; Wu, Z.; et al. Low-temperature combustion-synthesized nickel oxide thin films as hole-transport interlayers for solution-processed optoelectronic devices. Adv. Energy Mater. 2014, 4, 1301460. [Google Scholar] [CrossRef]
  29. Liu, Z.; Chang, J.; Lin, Z.; Zhou, L.; Yang, Z.; Chen, D.; Zhang, C.; Liu, S. (Frank); Hao, Y. High-Performance Planar Perovskite Solar Cells Using Low Temperature, Solution–Combustion-Based Nickel Oxide Hole Transporting Layer with Efficiency Exceeding 20%. Adv. Energy Mater. 2018, 8. [Google Scholar] [CrossRef]
  30. Liu, Y.; Cai, H.; Su, J.; Ye, X.; Yang, J.; Liang, X.; Guan, J.; Zhou, X.; Yin, J.; Ni, J.; et al. Solution-combustion-based nickel oxide hole transport layers via fuel regulation in inverted planar perovskite solar cells. J. Mater. Sci. Mater. Electron. 2020, 31, 15225–15232. [Google Scholar] [CrossRef]
  31. Pang, S.; Zhang, C.; Dong, H.; Chen, D.; Zhu, W.; Xi, H.; Chang, J.; Lin, Z.; Zhang, J.; Hao, Y. Efficient NiOx Hole Transporting Layer Obtained by the Oxidation of Metal Nickel Film for Perovskite Solar Cells. ACS Appl. Energy Mater. 2019, 2, 4700–4707. [Google Scholar] [CrossRef]
  32. Wei, Y.; Yao, K.; Wang, X.; Jiang, Y.; Liu, X.; Zhou, N.; Li, F. Improving the efficiency and environmental stability of inverted planar perovskite solar cells via silver-doped nickel oxide hole-transporting layer. Appl. Surf. Sci. 2018, 427, 782–790. [Google Scholar] [CrossRef]
  33. Hu, Z.; Chen, D.; Yang, P.; Yang, L.; Qin, L.; Huang, Y.; Zhao, X. Sol-gel-processed yttrium-doped NiO as hole transport layer in inverted perovskite solar cells for enhanced performance. Appl. Surf. Sci. 2018, 441, 258–264. [Google Scholar] [CrossRef]
  34. 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. [Google Scholar] [CrossRef]
  35. Lee, P.H.; Li, B.T.; Lee, C.F.; Huang, Z.H.; Huang, Y.C.; Su, W.F. High-efficiency perovskite solar cell using cobalt doped nickel oxide hole transport layer fabricated by NIR process. Sol. Energy Mater. Sol. Cells 2020, 208, 110352. [Google Scholar] [CrossRef]
  36. Dewan, S.; Tomar, M.; Tandon, R.P.; Gupta, V. Zn doping induced conductivity transformation in NiO films for realization of p-n homo junction diode. J. Appl. Phys. 2017, 121, 215307. [Google Scholar] [CrossRef]
  37. Svanström, S.; Jacobsson, T.J.; Boschloo, G.; Johansson, E.M.J.; Rensmo, H.; Cappel, U.B. Degradation Mechanism of Silver Metal Deposited on Lead Halide Perovskites. ACS Appl. Mater. Interfaces 2020, 12, 7212–7221. [Google Scholar] [CrossRef] [Green Version]
  38. Ponmudi, T.S.; Lee, C.-W.; Lai, C.-C.; Tsai, C.-H. Comparative study on the effect of annealing temperature on sol–gel-derived nickel oxide thin film as hole transport layers for inverted perovskite solar cells. J. Mater. Sci. Mater. Electron. 2021, 1–10. [Google Scholar] [CrossRef]
  39. Lin, Y.R.; Liao, Y.S.; Hsiao, H.T.; Chen, C.P. Two-step annealing of NiOx enhances the NiOx–perovskite interface for high-performance ambient-stable p–i–n perovskite solar cells. Appl. Surf. Sci. 2020, 504, 144478. [Google Scholar] [CrossRef]
  40. Nandy, S.; Saha, B.; Mitra, M.K.; Chattopadhyay, K.K. Effect of oxygen partial pressure on the electrical and optical properties of highly (200) oriented p-type Ni1-xO films by DC sputtering. J. Mater. Sci. 2007, 42, 5766–5772. [Google Scholar] [CrossRef]
  41. Wan, X.; Jiang, Y.; Qiu, Z.; Zhang, H.; Zhu, X.; Sikandar, I.; Liu, X.; Chen, X.; Cao, B. Zinc as a New Dopant for NiOx-Based Planar Perovskite Solar Cells with Stable Efficiency near 20%. ACS Appl. Energy Mater. 2018, 1, 3947–3954. [Google Scholar] [CrossRef]
  42. Mahmud, M.A.; Elumalai, N.K.; Upama, M.B.; Wang, D.; Chan, K.H.; Wright, M.; Xu, C.; Haque, F.; Uddin, A. Low temperature processed ZnO thin film as electron transport layer for efficient perovskite solar cells. Sol. Energy Mater. Sol. Cells 2017, 159, 251–264. [Google Scholar] [CrossRef]
  43. Lee, J.H.; Noh, Y.W.; Jin, I.S.; Jung, J.W. Efficient planar heterojunction perovskite solar cells employing a solution-processed Zn-doped NiOx hole transport layer. Electrochim. Acta 2018, 284, 253–259. [Google Scholar] [CrossRef]
  44. Tauc, J. Optical properties and electronic structure of amorphous Ge and Si. Mater. Res. Bull. 1968, 3, 37–46. [Google Scholar] [CrossRef]
  45. Hou, D.; Zhang, J.; Gan, X.; Yuan, H.; Yu, L.; Lu, C.; Sun, H.; Hu, Z.; Zhu, Y. Pb and Li co-doped NiOx for efficient inverted planar perovskite solar cells. J. Colloid Interface Sci. 2020, 559, 29–38. [Google Scholar] [CrossRef] [PubMed]
  46. Wu, Y.; Yang, X.; Chen, W.; Yue, Y.; Cai, M.; Xie, F.; Bi, E.; Islam, A.; Han, L. Perovskite solar cells with 18.21% efficiency and area over 1 cm2 fabricated by heterojunction engineering. Nat. Energy 2016, 1, 1–7. [Google Scholar] [CrossRef]
  47. García-Rodríguez, R.; Ferdani, D.; Pering, S.; Baker, P.J.; Cameron, P.J. Influence of bromide content on iodide migration in inverted MAPb(I1-XBrx)3 perovskite solar cells. J. Mater. Chem. A 2019, 7, 22604–22614. [Google Scholar] [CrossRef] [Green Version]
  48. Liu, D.; Li, Y.; Yuan, J.; Hong, Q.; Shi, G.; Yuan, D.; Wei, J.; Huang, C.; Tang, J.; Fung, M.K. Improved performance of inverted planar perovskite solar cells with F4-TCNQ doped PEDOT:PSS hole transport layers. J. Mater. Chem. A 2017, 5, 5701–5708. [Google Scholar] [CrossRef]
  49. Tsai, C.H.; Lin, C.M.; Kuei, C.H. Improving the performance of perovskite solar cells by adding 1,8-diiodooctane in the CH3NH3PbI3 perovskite layer. Sol. Energy 2018, 176, 178–185. [Google Scholar] [CrossRef]
Figure 1. (a) Device architecture, (b) energy level diagram of inverted PSCs, and (c) photograph of the fabricated device.
Figure 1. (a) Device architecture, (b) energy level diagram of inverted PSCs, and (c) photograph of the fabricated device.
Coatings 11 00627 g001
Figure 2. (a) JV characteristics and (b) EQE spectra of PSCs made with NiOx, Zn-NiOx, and PEDOT: PSS (control) HTLs.
Figure 2. (a) JV characteristics and (b) EQE spectra of PSCs made with NiOx, Zn-NiOx, and PEDOT: PSS (control) HTLs.
Coatings 11 00627 g002
Figure 3. (a) Stability graph and (b) photograph of the PSCs made with NiOx, 5% Zn-NiOx, and PEDOT: PSS (control) HTLs without encapsulation (under ambient air at room temperature for 4 days).
Figure 3. (a) Stability graph and (b) photograph of the PSCs made with NiOx, 5% Zn-NiOx, and PEDOT: PSS (control) HTLs without encapsulation (under ambient air at room temperature for 4 days).
Coatings 11 00627 g003
Figure 4. XRD patterns of NiOx and 5% Zn-NiOx thin films.
Figure 4. XRD patterns of NiOx and 5% Zn-NiOx thin films.
Coatings 11 00627 g004
Figure 5. XPS spectra of (a) Ni 2p, (b) O 1s, (c) Zn 2p, and (d) survey scan for NiOx and 5% Zn-NiOx thin films.
Figure 5. XPS spectra of (a) Ni 2p, (b) O 1s, (c) Zn 2p, and (d) survey scan for NiOx and 5% Zn-NiOx thin films.
Coatings 11 00627 g005
Figure 6. FESEM image and EDS result of (a) NiOx and (b) 5% Zn-NiOx thin films, (c) cross-sectional FESEM image of the device.
Figure 6. FESEM image and EDS result of (a) NiOx and (b) 5% Zn-NiOx thin films, (c) cross-sectional FESEM image of the device.
Coatings 11 00627 g006
Figure 7. AFM images of (a) NiOx and (b) 5% Zn-NiOx thin films.
Figure 7. AFM images of (a) NiOx and (b) 5% Zn-NiOx thin films.
Coatings 11 00627 g007
Figure 8. (a) Transmittance spectra and (b) Tauc plots of NiOx and 5% Zn-NiOx thin films.
Figure 8. (a) Transmittance spectra and (b) Tauc plots of NiOx and 5% Zn-NiOx thin films.
Coatings 11 00627 g008
Figure 9. IV characteristics of hole-only devices made with NiOx and 5% Zn-NiOx thin films.
Figure 9. IV characteristics of hole-only devices made with NiOx and 5% Zn-NiOx thin films.
Coatings 11 00627 g009
Figure 10. (a) PL spectra and (b) PL lifetime decay of NiOx and 5% Zn-NiOx thin films.
Figure 10. (a) PL spectra and (b) PL lifetime decay of NiOx and 5% Zn-NiOx thin films.
Coatings 11 00627 g010
Figure 11. EIS spectra of PSCs with NiOx and 5% Zn-NiOx HTL.
Figure 11. EIS spectra of PSCs with NiOx and 5% Zn-NiOx HTL.
Coatings 11 00627 g011
Table 1. Summary of device performance of the PSCs made with NiOx, Zn-NiOx, and PEDOT:PSS (control) HTLs.
Table 1. Summary of device performance of the PSCs made with NiOx, Zn-NiOx, and PEDOT:PSS (control) HTLs.
HTLVoc (V)Jsc (mA/cm2)FFPCE (%)Rs (Ω/cm2)Rsh (kΩ/cm2)
NiOxAverage1.047 ± 0.0117.47 ± 0.160.74 ± 0.0113.52 ± 0.106.92.59
best1.05117.390.7513.62
2% Zn-NiOxAverage1.045 ± 0.0217.87 ± 0.420.73 ± 0.0113.61 ± 0.156.81.87
best1.04618.020.7313.76
5% Zn-NiOxAverage1.046 ± 0.0119.26 ± 0.620.73 ± 0.0214.67 ± 0.206.91.78
best1.04619.670.7214.87
10% Zn-NiOxAverage0.942 ± 0.0120.17 ± 0.190.57 ± 0.0110.77 ± 0.139.20.19
best0.94920.280.5710.90
Control
PEDOT:PSS
Average0.905 ± 0.0217.91 ± 0.460.68 ± 0.0111.21 ± 0.227.50.45
best0.92518.030.6911.43
Table 2. PL lifetime parameters from the fitting curves of NiOx and 5% Zn-NiOx thin films.
Table 2. PL lifetime parameters from the fitting curves of NiOx and 5% Zn-NiOx thin films.
SampleA1 (%)τ1 (ns)A2 (%)τ2 (ns)τavg (ns)
FTO/PSK69.420.430.646.033.2
FTO/NiOx/PSK84.94.315.117.910.1
FTO/5%Zn-NiOx/PSK82.04.418.015.69.3
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Thiruchelvan, P.S.; Lai, C.-C.; Tsai, C.-H. Combustion Processed Nickel Oxide and Zinc Doped Nickel Oxide Thin Films as a Hole Transport Layer for Perovskite Solar Cells. Coatings 2021, 11, 627. https://doi.org/10.3390/coatings11060627

AMA Style

Thiruchelvan PS, Lai C-C, Tsai C-H. Combustion Processed Nickel Oxide and Zinc Doped Nickel Oxide Thin Films as a Hole Transport Layer for Perovskite Solar Cells. Coatings. 2021; 11(6):627. https://doi.org/10.3390/coatings11060627

Chicago/Turabian Style

Thiruchelvan, Ponmudi Selvan, Chien-Chih Lai, and Chih-Hung Tsai. 2021. "Combustion Processed Nickel Oxide and Zinc Doped Nickel Oxide Thin Films as a Hole Transport Layer for Perovskite Solar Cells" Coatings 11, no. 6: 627. https://doi.org/10.3390/coatings11060627

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop