Next Article in Journal
A Fast-Response Ultraviolet Phototransistor with a PVK QDs/ZnO Nanowire Heterostructure and Its Application in Pharmaceutical Solute Detection
Previous Article in Journal
Microstructural Characterization of Al/CNTs Nanocomposites after Cold Rolling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Highly Stable and Enhanced Performance of p–i–n Perovskite Solar Cells via Cuprous Oxide Hole-Transport Layers

1
Institute of Materials Science and Engineering, National Taiwan University, Taipei 106, Taiwan
2
Department of Mechanical Engineering, National Taiwan University of Science and Technology, Taipei 106, Taiwan
3
Department of Materials Engineering and Center for Plasma and Thin Film Technologies, Ming Chi University of Technology, New Taipei City 243, Taiwan
4
Mechanical Engineering Department, Faculty of Engineering and Quantity Surveying, INTI International University, Nilai 71800, Negeri Sembilan, Malaysia
5
College of Engineering and Center for Green Technology, Chang Gung University, Taoyuan 333, Taiwan
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(8), 1363; https://doi.org/10.3390/nano13081363
Submission received: 6 March 2023 / Revised: 6 April 2023 / Accepted: 10 April 2023 / Published: 14 April 2023

Abstract

:
Solar light is a renewable source of energy that can be used and transformed into electricity using clean energy technology. In this study, we used direct current magnetron sputtering (DCMS) to sputter p-type cuprous oxide (Cu2O) films with different oxygen flow rates (fO2) as hole-transport layers (HTLs) for perovskite solar cells (PSCs). The PSC device with the structure of ITO/Cu2O/perovskite/[6,6]-phenyl-C61-butyric acid methyl ester (PC61BM)/bathocuproine (BCP)/Ag showed a power conversion efficiency (PCE) of 7.91%. Subsequently, a high-power impulse magnetron sputtering (HiPIMS) Cu2O film was embedded and promoted the device performance to 10.29%. As HiPIMS has a high ionization rate, it can create higher density films with low surface roughness, which passivates surface/interface defects and reduces the leakage current of PSCs. We further applied the superimposed high-power impulse magnetron sputtering (superimposed HiPIMS) derived Cu2O as the HTL, and we observed PCEs of 15.20% under one sun (AM1.5G, 1000 Wm−2) and 25.09% under indoor illumination (TL-84, 1000 lux). In addition, this PSC device outperformed by demonstrating remarkable long-term stability via retaining 97.6% (dark, Ar) of its performance for over 2000 h.

Graphical Abstract

1. Introduction

Solar energy is the most abundant and promising alternative to fossil energy. It can be collected and converted easily in the form of electricity for industrial and household usage via various technologies such as silicon-based solar cells (SSCs) [1], dye-sensitized solar cells (DSSCs) [2], organic photovoltaics (OPVs) [3,4], and perovskite solar cells (PSCs) [5,6]. Among these, PSCs have gained the most interest due to their impressive power conversion efficiency (PCE) [7,8,9,10]. The first PSC, introduced by Miyasaka in 2009, had a PCE of 3.8% [11], which now exceeds 25.7% [12] within quite a few years of development. Because of their easy fabrication in low temperatures, inexpensive solution process, and excellent performance attributed to outstanding optoelectronic properties [13,14,15,16,17,18,19,20,21,22,23,24,25], PSCs have attracted more attention. Perovskites typically have the chemical formula ABX3, where A is a monovalent cation (such as methyl ammonium [CH3NH3+ (MA)], NH = CHNH3+ (FA), or Cs+), B is either a Pb2+ or Sn2+ divalent metallic cation, and X is a halide anion (I, Br, or Cl) [26,27].
The classifications of PSC device architecture based on arrangement of heterojunctions are n−i−p (conventional) and p−i−n (inverted). The selection of materials for hole-transport layers (HTLs) shows the impact on the performance and stability of PSCs. Commonly used hole-transport materials (HTMs) are 2,2′,7,7′-tetrakis-(N, N-di-4-methoxyphenylamino)-9,9′-spirobifluorene (Spiro-OMeTAD) [28], poly(3,4-ethylene dioxythiophene) (PEDOT:PSS) [29], poly(3-hexylthiophene) (P3HT) [30], and poly[bis(4-phenyl) (2,4,6-trimethylphenyl)-amine] (PTAA) [31]. These organic materials become highly unstable when exposed to air (moisture) for a long time. Thereby, the replacement of organic HTMs with inorganic HTMs is being considered for comparatively higher stability. For instance, NiOx has been used as an efficient HTL possessing an outstanding performance with excellent air stability [19,32,33]. Cuprous oxide (Cu2O) offers a band gap in the range of 2.2–2.8 eV with an enormous hole mobility of (100 cm2V−1s−1) [34,35,36] and the features of nontoxicity, abundance on the earth’s crust, and low manufacturing cost, which make it a candidate for HTLs of PSCs [37]. Relentless research on Cu2O-based, transparent thin-film PSCs has achieved significant theoretical energy conversion efficiency [38,39]. There are various chemical/solution and physical deposition methods to prepare Cu2O thin films, such as sol-gel [40], thermal oxidation [34], chemical vapor deposition [41], RF/DC sputtering [42,43], electrodeposition [44], pulsed laser deposition [45], and high-power impulse magnetron sputtering (HiPIMS) [46]. Even though Cu2O thin films have outstanding properties, it is still challenging to deposit uniform, dense, smooth, and pinhole-free layers for PSC application.
Various sputtering methods have been considered, such as direct current magnetron sputtering (DCMS) (i.e, a plasma process used for film deposition). The HiPIMS technique offers a high ionization rate of sputtered target atoms and high plasma density, which is advantageous for metal-oxide thin-film deposition [47]. Unfortunately, HiPIMS has the drawback of a low deposition rate, which limits its application. Here, middle-frequency (MF) pulses during the off-time of HiPIMS, also known as superimposed HiPIMS (HiPIMS+MF), step in to overcome the deposition rate of the HiPIMS process without sacrificing the target ionization rate [48]. In this work, we demonstrated Cu2O as the HTLs of PSCs from different sputtering modes (DCMS, HiPIMS, and superimposed HiPIMS). We analyzed the structural, morphological, and optoelectronic properties of Cu2O derived PSCs. The DC, HiPIMS, and superimposed HiPIMS derived devices showed performances of 7.12 ± 0.64, 9.42 ± 0.92%, and 13.04 ± 1.61%, respectively. To the best of our knowledge, the PCE of the superimposed HiPIMS derived PSCs is among the best reported for CH3NH3PbI3 (MAPbI3) based devices. Furthermore, the respective device showed remarkable stability by retaining 97.6% of its performance (i.e., storage in the dark [Ar] for over 2000 h). Hence, this study explains the effect of depositing TCO films using different sputtering modes on the PCEs and stability of PSC devices.

2. Materials and Methods

Cu2O films (10 nm) were deposited by DCMS, HiPIMS [49], and superimposed HiPIMS [48] techniques, subsequently, on an ITO substrate. Before spin-coating the perovskite precursor solution, we used plasma to modify the Cu2O surface for 30 s and coat the precursor solution smoothly. Details of the device fabrication and characterization of the materials are shown in the Supplementary Materials.

3. Results

We fabricated the PSCs with a p–i–n structure of Glass/ITO/Cu2O/MAPbI3/PC61BM/BCP/Ag. Figure 1 displays the X-ray diffraction patterns of Cu2O from DCMS and superimposed HiPIMS at different oxygen flow rates (fO2 = O2/(O2 + Ar) × 100%, Ar: Argon O2: Oxygen). For the DCMS (Figure 1a), the increase in oxygen flow rate from fO2 = 10% started to form the Cu2O phase and became pure Cu2O phase at fO2 = 17.5%. Beyond this, it started to form a mixed Cu2O phase. For the normal HiPIMS sample, we observed the Cu2O phase at fO2 = 20%, which is consistent with our previously published study [49]. For the XRD of the superimposed HiPIMS process (Figure 1b), the increase in oxygen flow rate from fO2 = 17.5% started to form the Cu2O phase and became pure Cu2O phase at fO2 = 30%, and later, with a further increase in fO2, it became a mixed-Cu2O phase. Accordingly, we obtained the optimal parameters for the Cu2O phase and used them as the HTLs of PSCs for further study.
We determined the UV-Vis transmittance of these Cu2O-deposited ITO samples using air as a background (Figure 2a). The transmittance of the HiPIMS sample was slightly more than 80%, while in contrast, that of the DCMS and superimposed HiPIMS samples were each a bit less than 80% (<3%), in a wavelength range (500-900 nm). Due to the high ionization rate and high-energy plasma of the HiPIMS process, it offers dense and highly crystalline Cu2O films. Therefore, we observed higher transmittance of respective Cu2O in comparison with the DCMS sample (i.e., the structure was less dense). Although the superimposed HiPIMS process has the same advantages as the HiPIMS process, the MF power applied during the off-time bombarded more copper that became deposited as film, possibly contributing to the decrease in transmittance.
The values of Ra (the arithmetic average of the roughness profile) of the samples were determined by tapping-mode atomic force microscopy (AFM) to understand the surface morphology of ITO/Cu2O affecting the growth of perovskite; Figure 2b represents surface roughness through AFM images of the samples. We obtained values of Ra for the DCMS (3.06 nm), HiPIMS (2.94 nm), and superimposed HiPIMS (2.93 nm) samples. We speculated that HiPIMS would result in higher density and higher energy plasma than DCMS during deposition, which would result in a smoother surface. The superimposed HiPIMS sample had the same characteristics as the HiPIMS sample, leading to similar Ra values.
Figure 3 shows the contact angles of the ITO/Cu2O surfaces when we dropped the perovskite precursor solution on top. As indicated in Figure 3a, the perovskite precursor formed a sphere on the HiPIMS Cu2O samples (i.e., a high angle of 85.13°), spreading poorly on the Cu2O surface and causing failure growth of the perovskite layer. The DCMS and superimposed HiPIMS Cu2O samples also had high angles, of 89.95° and 69.31°, respectively, as shown in Figure S1, which also led to poor spread of the perovskite precursor. To resolve this issue, we applied a plasma treatment to induce an ultraviolet effect that modified the surface of the Cu2O films. As indicated in Figure 3b, the Cu2O film became hydrophilic with a low angle of 9.99°, leading to the preferred growth of the perovskite layer.
Table 1 and Figure 4 summarize the device performances for various Cu2O films. Table 1 displays the standard deviations of the PSCs calculated from a minimum of three samples deposited using the same technique. The perovskite deposition was from a solution, which may cause slight variations during each deposition. Due to the variability in optoelectronic and surface properties caused by the deposition process of HTLs, the PSCs produced using various conditions of HTLs exhibited different device performances. DCMS-derived devices provided a PCE of 7.12 ± 0.64%, with a short-circuit current density (Jsc) of 13.74 ± 0.97 mAcm−2, a value of open-circuit voltage (Voc) of 0.89 ± 0.02 V, and a fill factor (FF) of 58.1 ± 3.93%. The HiPIMS-based devices showed slightly higher efficiency, with a PCE of 9.42 ± 0.92%, a value of Jsc of 17.36 ± 0.94 mAcm−2, a value of Voc of 0.90 ± 0.00 V, and an FF of 61.1 ± 6.95%. The superimposed HiPIMS devices showed the highest performance, with a PCE of 13.04 ± 1.61%, a value of Jsc of 19.27 ± 1.20 mAcm−2, a value of Voc of 0.98 ± 0.00 V, and an FF of 69.5 ± 4.98%. The improvement in the PCE of the overlaid HiPIMS PSCs was brought on by substantial increases in the Jsc, Voc, and FFs. From the device’s J-V curves (Figure 4), we computed the series (Rs) and shunt (Rsh) resistances. The resistances of the electrodes, the interfacial resistance, the hole/electron transporting layers, and the perovskite all contributed to the value of Rs. The highest value of Rs was for the DCMS device (9.47 Ω cm2), and the lowest value of Rs was for the superimposed HiPIMS device (1.42 Ω cm2). Further, we performed a four-point probe experiment to understand the electrical properties of Cu2O. This analysis showed the highest value of electrical conductivity (σ) for the superimposed HiPIMS Cu2O thin film (3.33 S cm−1) and the lowest for the DCMS Cu2O thin film (0.11 S cm−1). In addition, through XRD analysis (Figure S2), the crystallinity of the Cu2O improved from the DCMS to superimposed HiPIMS deposition techniques. According to previous studies, an increase in film’s crystallinity helps to increase the charge-carrier mobility (μ) [50]. In addition, varying the exposure times of the sputtering and plasma treatments impacted the optoelectronic properties of the thin films. In an attempt to reduce the exposure time during sputtering, we observed that the quality of the Cu2O film deteriorated significantly, potentially due to the film becoming discontinuous. Conversely, increasing the exposure time led to a well-formed Cu2O film. However, it resulted in unsatisfactory formation of the perovskite layer, and we are currently investigating the underlying reasons for this. As for the plasma treatment, we have not yet examined the impact of different exposure times on the optoelectronic properties of Cu2O film, as our primary goal was to improve its wettability. However, we speculate that excessively long exposure times during plasma treatment may damage Cu2O film and decrease its photoelectric properties. Lower values of Rs appear to be associated with higher σ and μ of Cu2O thin films and increased Jsc values of the device. We speculate that the increase in Jsc could have been due to a reduction in the carrier recombination rate at the p-i interface. This explanation is supported by other studies reporting that increasing hole conductivity can lead to higher Jsc [51]. We acknowledge that we do not have the necessary facilities to perform spectral-response characterization experiments, but our findings align with the existing literature. Overall, the use of superimposed HiPIMS Cu2O as HTLs resulted in higher FF, Jsc, and PCE values compared with the devices based on DCMS and HiPIMS. As indicated in Figure 4, 15.20% was the best performance of the superimposed HiPIMS devices. Table 2 compares Cu2O films as the HTLs of PSCs prepared by different processes; Voc, Jsc, FF, and PCE are the best performance values of the device. The Cu2O films were prepared by various methods such as spin coating [52,53], successive ionic layer adsorption and reaction (SILAR) [54], sputtering [55,56], the thermal oxidation method [57,58], electrodeposition [59,60], chemical vapor deposition (CVD) [61], etc. Although it is known from simulation results that, theoretically, the PCE of Cu2O-based PSCs can be as high as 25.2% [62], it was found that the actual device PCE was mostly in the range of 6 to 13%. Based on the above method, we observed that the improvement in PCE was due to the low series and high shunt resistance of the respective devices. The high conductivity of the Cu2O thin film resulted in low recombination loss and smaller bulk resistance of the device. These factors contributed to a reduction in the energy loss within the device and increased the efficiency of charge extraction and transfer. This phenomenon is consistent with findings in previous reports [54,60].
In this study, we achieved a competitive result of 15.21% via embedding Cu2O thin films using superimposed HiPIMS. The MF power-supply sputtering in the superimposed HiPIMS system allowed us to retain a high ionization rate as the advantage of the HiPIMS. Compared with DCMS processes, HiPIMS and superimposed HiPIMS can bombard sputtering ions with higher energy, which helps to form the preferred stoichiometry of Cu2O films, resulting in a denser coating with lower surface roughness. It helped to reduce the defects in the film and led to a smoother interface, resulting in a decrease in the leakage current and an improvement in the PCE. Our EPMA analysis (Table S1) showed that the Cu2O chemical composition ratio as Cu/O was 1.9, which is close to the form of perfect stoichiometric Cu2O.
In addition to evaluating the surface morphologies of the Cu2O films, we also evaluated the energy level alignment at the Cu2O/p-i interface using an incident light energy of 21.2 eV (He(I) emission) via ultraviolet photoelectron spectroscopy (UPS). Figure 5 presents the ECut-off regions of the DCMS, HiPIMS, and superimposed HiPIMS Cu2O samples (WF = 21.2 − ECut-off) [63,64]. The DCMS Cu2O film had a work function (WF) of –6.40 eV, while the WF of the HiPIMS Cu2O film was –6.38 eV, which was 0.02 eV lesser than that of the DCMS Cu2O film. The WF of the superimposed HiPIMS sample was –6.07 eV (the negative work function means how much energy is required to be added to the bound electron by the photon it absorbs). Changes in the WF at the p-i interface can significantly impact the energy alignment of Cu2O with perovskite, hole transport, and effectiveness of charge extraction as well as transfer. Therefore, a shift in the energy levels of the Cu2O may have contributed to the improved performance of the Cu2O devices. We believe that the significantly different composition ratio (Table S1) and phase structure (Figure S2) of the superimposed HiPIMS Cu2O film may have affected the WF, resulting in the improved device performance.
We examined the shelf lives of the unencapsulated devices under an argon glove box in the dark to assess the stability of the superimposed HiPIMS Cu2O sample for PSC applications (we measured the PCE results at least three times to calculate the error bars). The devices initially offered the best PCE of 15.20%, as shown in Figure 6. After testing for 168 h, the devices retained 99.4% of their initial performance; after testing over 1000 h, the devices retained 98.3% of their initial performance; and after testing over 2000 h, the devices retained 97.6% of their initial performance. Thus, our PSC device suggested high stability. Finally, we determined the performance of the superimposed HiPIMS device under dim light (a TL-84 fluorescent lamp with an illumination of 1000 lux). Table 3 shows the performance of the respective device. The device showed a PCE of 25.09%. Based on these results, we demonstrated that Cu2O film has great potential as HTLs of PSCs under one sun and indoor illumination.

4. Conclusions

In this study, we prepared Cu2O films via DCMS, HiPIMS, and superimposed HiPIMS and used them as HTLs for PSC applications. In the comparison of different process technologies for Cu2O samples, superimposed HiPIMS enhanced the WF (measured by UPS), surface roughness (measured by AFM), and optoelectronic properties of the Cu2O films and, in turn, increased carrier extraction and transport at the perovskite–Cu2O interface. Because the optimized Cu2O film was obtained from superimposed HiPIMS technology, we observed an increase in the PCE of the PSCs from 7.91 (DCMS) to 15.20%. The extended shelf life of the unencapsulated devices in the glove box attested to their excellent stability and support Cu2O as a contender for p-type transporting layers in PSC applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano13081363/s1, Experimental Details, Figure S1: Cu2O film prepared by (a) DCMS, (b) superimposed HiPIMS of contact angle.; Figure S2: X-ray diffraction spectra of different process technologies.; Table S1: Chemical composition of different process technologies. References [19,65,66] are cited in the supplementary materials.

Author Contributions

Conceptualization, T.-H.C. and S.-C.C.; data curation, Y.-H.C.; investigation, Y.-H.C.; writing—original draft preparation, S.S. and Y.-H.C.; writing—review and editing, S.S. and C.-P.C.; supervision, T.-H.C., W.-C.C., C.K.C., C.-P.C. and S.-C.C.; resources, T.-H.C. and S.-C.C. All authors have read and agreed to the published version of the manuscript.

Funding

We sincerely appreciate the financial assistance provided by the Ministry of Science and Technology of Taiwan (MOST 111-2221-E-131-027).

Acknowledgments

We thank H.-C. Lin and C.-Y. Kao of the Instrumentation Center, National Taiwan University for their assistance with the EPMA experiments.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ellis, F., Jr.; Delahoy, A. Optical properties of hydrogenated amorphous silicon based solar cells. Sol. Energy Mater. 1986, 13, 109–132. [Google Scholar] [CrossRef]
  2. Chen, S.; Kumar, R.V.; Gedanken, A.; Zaban, A. Sonochemical synthesis of crystalline nanoporous zinc oxide spheres and their application in dye-sensitized solar cells. Isr. J. Chem. 2001, 41, 51–54. [Google Scholar] [CrossRef]
  3. Kushto, G.P.; Kim, W.; Kafafi, Z.H. Flexible organic photovoltaics using conducting polymer electrodes. Appl. Phys. Lett. 2005, 86, 093502. [Google Scholar] [CrossRef]
  4. Jiang, B.H.; Peng, Y.J.; Su, Y.W.; Chang, J.F.; Chueh, C.C.; Shieh, T.S.; Huang, C.I.; Chen, C.P. A polymer donor with versatility for fabricating high-performance ternary organic photovoltaics. Chem. Eng. J. 2022, 431, 133950. [Google Scholar] [CrossRef]
  5. Chen, W.; Wu, Y.; Yue, Y.; Liu, J.; Zhang, W.; Yang, X.; Chen, H.; Bi, E.; Ashraful, I.; Grätzel, M. Efficient and stable large-area perovskite solar cells with inorganic charge extraction layers. Science 2015, 350, 944–948. [Google Scholar] [CrossRef] [Green Version]
  6. Lin, S.C.; Cheng, T.H.; Chen, C.P.; Chen, Y.C. Structural effect on triphenylamine dibenzofulvene based interfacial hole transporting materials for high-performance inverted perovskite solar cells. Mater. Chem. Phys. 2022, 288, 126385. [Google Scholar] [CrossRef]
  7. Christians, J.A.; Schulz, P.; Tinkham, J.S.; Schloemer, T.H.; Harvey, S.P.; Tremolet de Villers, B.J.; Sellinger, A.; Berry, J.J.; Luther, J.M. Tailored interfaces of unencapsulated perovskite solar cells for >1000 hour operational stability. Nat. Energy 2018, 3, 68–74. [Google Scholar] [CrossRef]
  8. Leijtens, T.; Giovenzana, T.; Habisreutinger, S.N.; Tinkham, J.S.; Noel, N.K.; Kamino, B.A.; Sadoughi, G.; Sellinger, A.; Snaith, H.J. Hydrophobic organic hole transporters for improved moisture resistance in metal halide perovskite solar cells. ACS Appl. Mater. Interfaces 2016, 8, 5981–5989. [Google Scholar] [CrossRef] [PubMed]
  9. Mazzarella, L.; Lin, Y.H.; Kirner, S.; Morales Vilches, A.B.; Korte, L.; Albrecht, S.; Crossland, E.; Stannowski, B.; Case, C.; Snaith, H.J. Infrared light management using a nanocrystalline silicon oxide interlayer in monolithic perovskite/silicon heterojunction tandem solar cells with efficiency above 25%. Adv. Energy Mater. 2019, 9, 1803241. [Google Scholar] [CrossRef]
  10. Jiang, Q.; Tong, J.; Xian, Y.; Kerner, R.A.; Dunfield, S.P.; Xiao, C.; Scheidt, R.A.; Kuciauskas, D.; Wang, X.; Hautzinger, M.P. Surface reaction for efficient and stable inverted perovskite solar cells. Nature 2022, 611, 278–283. [Google Scholar] [CrossRef]
  11. 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]
  12. Kim, M.; Jeong, J.; Lu, H.; Lee, T.K.; Eickemeyer, F.T.; Liu, Y.; Choi, I.W.; Choi, S.J.; Jo, Y.; Kim, H.B. Conformal quantum dot–SnO2 layers as electron transporters for efficient perovskite solar cells. Science 2022, 375, 302–306. [Google Scholar] [CrossRef] [PubMed]
  13. Qin, P.L.; He, Q.; Chen, C.; Zheng, X.L.; Yang, G.; Tao, H.; Xiong, L.B.; Xiong, L.; Li, G.; Fang, G.J. High-Performance Rigid and Flexible Perovskite Solar Cells with Low-Temperature Solution-Processable Binary Metal Oxide Hole-Transporting Materials. Sol. RRL 2017, 1, 1700058. [Google Scholar] [CrossRef]
  14. Qin, P.L.; Yang, G.; Ren, Z.W.; Cheung, S.H.; So, S.K.; Chen, L.; Hao, J.; Hou, J.; Li, G. Stable and efficient organo-metal halide hybrid perovskite solar cells via π-conjugated lewis base polymer induced trap passivation and charge extraction. Adv. Mater. 2018, 30, 1706126. [Google Scholar] [CrossRef] [PubMed]
  15. Zhu, H.L.; Liang, Z.; Huo, Z.; Ng, W.K.; Mao, J.; Wong, K.S.; Yin, W.J.; Choy, W.C. Low-Bandgap Methylammonium-Rubidium Cation Sn-Rich Perovskites for Efficient Ultraviolet–Visible–Near Infrared Photodetectors. Adv. Funct. Mater. 2018, 28, 1706068. [Google Scholar] [CrossRef]
  16. Zhu, Z.; Zhao, D.; Chueh, C.C.; Shi, X.; Li, Z.; Jen, A.K.Y. Highly efficient and stable perovskite solar cells enabled by all-crosslinked charge-transporting layers. Joule 2018, 2, 168–183. [Google Scholar] [CrossRef] [Green Version]
  17. Correa Baena, J.P.; Saliba, M.; Buonassisi, T.; Grätzel, M.; Abate, A.; Tress, W.; Hagfeldt, A. Promises and challenges of perovskite solar cells. Science 2017, 358, 739–744. [Google Scholar] [CrossRef] [Green Version]
  18. Huang, J.; Yuan, Y.; Shao, Y.; Yan, Y. Understanding the physical properties of hybrid perovskites for photovoltaic applications. Nat. Rev. Mater. 2017, 2, 17042. [Google Scholar] [CrossRef]
  19. Hsu, H.L.; Hsiao, H.T.; Juang, T.Y.; Jiang, B.H.; Chen, S.C.; Jeng, R.J.; Chen, C.P. Carbon nanodot additives realize high-performance air-stable p–i–n perovskite solar cells providing efficiencies of up to 20.2%. Adv. Energy Mater. 2018, 8, 1802323. [Google Scholar] [CrossRef]
  20. Singh, M.; Ng, A.; Ren, Z.; Hu, H.; Lin, H.C.; Chu, C.W.; Li, G. Facile synthesis of composite tin oxide nanostructures for high-performance planar perovskite solar cells. Nano Energy 2019, 60, 275–284. [Google Scholar] [CrossRef]
  21. Zhang, H.; Nazeeruddin, M.K.; Choy, W.C. Perovskite photovoltaics: The significant role of ligands in film formation, passivation, and stability. Adv. Mater. 2019, 31, 1805702. [Google Scholar] [CrossRef] [PubMed]
  22. Ono, L.K.; Juarez Perez, E.J.; Qi, Y. Progress on perovskite materials and solar cells with mixed cations and halide anions. ACS Appl. Mater. Interfaces 2017, 9, 30197–30246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Lee, K.M.; Chen, K.S.; Wu, J.R.; Lin, Y.D.; Yu, S.M.; Chang, S.H. Highly efficient and stable semi-transparent perovskite solar modules with a trilayer anode electrode. Nanoscale 2018, 10, 17699–17704. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Hsu, H.L.; Chen, C.P.; Chang, J.Y.; Yu, Y.Y.; Shen, Y.K. Two-step thermal annealing improves the morphology of spin-coated films for highly efficient perovskite hybrid photovoltaics. Nanoscale 2014, 6, 10281–10288. [Google Scholar] [CrossRef]
  25. Zeng, X.; Zhou, T.; Leng, C.; Zang, Z.; Wang, M.; Hu, W.; Tang, X.; Lu, S.; Fang, L.; Zhou, M. Performance improvement of perovskite solar cells by employing a CdSe quantum dot/PCBM composite as an electron transport layer. J. Mater. Chem. A 2017, 5, 17499–17505. [Google Scholar] [CrossRef]
  26. Cao, X.; Zhi, L.; Jia, Y.; Li, Y.; Zhao, K.; Cui, X.; Ci, L.; Zhuang, D.; Wei, J. A review of the role of solvents in formation of high-quality solution-processed perovskite films. ACS Appl. Mater. Interfaces 2019, 11, 7639–7654. [Google Scholar] [CrossRef]
  27. Li, S.; Ren, H.; Yan, Y. Boosting efficiency of planar heterojunction perovskite solar cells to 21.2% by a facile two-step deposition strategy. Appl. Surf. Sci. 2019, 484, 1191–1197. [Google Scholar] [CrossRef]
  28. Lin, Z.; Li, J.; Li, H.; Mo, Y.; Pan, J.; Wang, C.; Zhang, X.L.; Bu, T.; Zhong, J.; Cheng, Y.B. A novel dopant for spiro-OMeTAD towards efficient and stable perovskite solar cells. Sci. China Mater. 2021, 64, 2915–2925. [Google Scholar] [CrossRef]
  29. Li, P.; Mohamed, M.I.O.; Xu, C.; Wang, X.; Tang, X. Electrical property modified hole transport layer (PEDOT: PSS) enhance the efficiency of perovskite solar cells: Hybrid co-solvent post-treatment. Org. Electron. 2020, 78, 105582. [Google Scholar] [CrossRef]
  30. Zhang, Y.; Liu, W.; Tan, F.; Gu, Y. The essential role of the poly (3-hexylthiophene) hole transport layer in perovskite solar cells. J. Power Sources 2015, 274, 1224–1230. [Google Scholar] [CrossRef]
  31. Wang, Y.; Duan, L.; Zhang, M.; Hameiri, Z.; Liu, X.; Bai, Y.; Hao, X. PTAA as Efficient Hole Transport Materials in Perovskite Solar Cells: A Review. Sol. RRL 2022, 6, 2200234. [Google Scholar] [CrossRef]
  32. 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]
  33. Cao, J.; Yu, H.; Zhou, S.; Qin, M.; Lau, T.K.; Lu, X.; Zhao, N.; Wong, C.P. Low-temperature solution-processed NiOx films for air-stable perovskite solar cells. J. Mater. Chem. A 2017, 5, 11071–11077. [Google Scholar] [CrossRef]
  34. Musa, A.; Akomolafe, T.; Carter, M. Production of cuprous oxide, a solar cell material, by thermal oxidation and a study of its physical and electrical properties. Sol. Energy Mater. Sol. Cells 1998, 51, 305–316. [Google Scholar] [CrossRef]
  35. Abdelfatah, M.; Ledig, J.; El-Shaer, A.; Wagner, A.; Sharafeev, A.; Lemmens, P.; Mosaad, M.M.; Waag, A.; Bakin, A. Fabrication and characterization of flexible solar cell from electrodeposited Cu2O thin film on plastic substrate. Sol. Energy 2015, 122, 1193–1198. [Google Scholar] [CrossRef]
  36. Rühle, S.; Barad, H.; Bouhadana, Y.; Keller, D.; Ginsburg, A.; Shimanovich, K.; Majhi, K.; Lovrincic, R.; Anderson, A.; Zaban, A. Combinatorial solar cell libraries for the investigation of different metal back contacts for TiO2–Cu2O hetero-junction solar cells. Phys. Chem. Chem. Phys. 2014, 16, 352156. [Google Scholar] [CrossRef]
  37. Aseena, S.; Abraham, N.; Sahaya Dennish Babu, G.; Kathiresan, S.; Suresh Babu, V. Solution-Synthesized Cu2O as a Hole Transport Layer for a ZnO-Based Planar Heterojunction Perovskite Solar Cell Fabricated at Room Temperature. J. Electron. Mater. 2022, 51, 1692–1699. [Google Scholar] [CrossRef]
  38. Jeong, S.; Mittiga, A.; Salza, E.; Masci, A.; Passerini, S. Electrodeposited ZnO/Cu2O heterojunction solar cells. Electrochim. Acta 2008, 53, 2226–2231. [Google Scholar] [CrossRef]
  39. Yu, L.; Xiong, L.; Yu, Y. Cu2O homojunction solar cells: F-doped N-type thin film and highly improved efficiency. J. Phys. Chem. C 2015, 119, 22803–22811. [Google Scholar] [CrossRef]
  40. Halin, D.; Talib, I.; Daud, A.; Hamid, M. Characterizations of cuprous oxide thin films prepared by sol-gel spin coating technique with different additives for the photoelectrochemical solar cell. Int. J. Photoenergy 2014, 2014, 352156. [Google Scholar] [CrossRef] [Green Version]
  41. Chua, D.; Kim, S.B.; Gordon, R. Enhancement of the open circuit voltage of Cu2O/Ga2O3 heterojunction solar cells through the mitigation of interfacial recombination. AIP Adv. 2019, 9, 055203. [Google Scholar] [CrossRef] [Green Version]
  42. Sanjana, T.; Sunil, M.A.; Shaik, H.; Kumar, K.N. Studies on DC sputtered cuprous oxide thin films for solar cell absorber layers. Mater. Chem. Phys. 2022, 281, 125922. [Google Scholar] [CrossRef]
  43. Yang, F.F.; Peng, W.B.; Zhou, Y.J.; Li, R.; Xiang, G.J.; Yue Liu, J.M.Z.; Zhang, J.H.; Zhao, Y.; Wang, H. Thermal optimization of defected Cu2O photon-absorbing layer and the steady p-Cu2O/n-Si photovoltaic application. Vacuum 2022, 198, 110876. [Google Scholar] [CrossRef]
  44. Georgieva, V.; Ristov, M. Electrodeposited cuprous oxide on indium tin oxide for solar applications. Sol. Energy Mater. Sol. Cells 2002, 73, 67–73. [Google Scholar] [CrossRef]
  45. Kartha, C.V.; Rehspringer, J.L.; Muller, D.; Roques, S.; Bartringer, J.; Ferblantier, G.; Slaoui, A.; Fix, T. Insights into Cu2O thin film absorber via pulsed laser deposition. Ceram. Int. 2022, 48, 15274–15281. [Google Scholar] [CrossRef]
  46. Sun, H.; Chen, S.C.; Wen, C.K.; Chuang, T.H.; Yazdi, M.A.P.; Sanchette, F.; Billard, A. p-type cuprous oxide thin films with high conductivity deposited by high power impulse magnetron sputtering. Ceram. Int. 2017, 43, 6214–6220. [Google Scholar] [CrossRef]
  47. Sun, H.; Kuo, T.Y.; Chen, S.C.; Chen, Y.H.; Lin, H.C.; Yazdi, M.A.P.; Billard, A. Contribution of enhanced ionization to the optoelectronic properties of p-type NiO films deposited by high power impulse magnetron sputtering. J. Eur. Ceram. Soc. 2019, 39, 5285–5291. [Google Scholar] [CrossRef]
  48. Chuang, T.H.; Wen, C.K.; Chen, S.C.; Liao, M.H.; Liu, F.; Sun, H. p-type semi-transparent conductive NiO films with high deposition rate produced by superimposed high power impulse magnetron sputtering. Ceram. Int. 2020, 46, 27695–27701. [Google Scholar] [CrossRef]
  49. Sun, H.; Wen, C.K.; Chen, S.C.; Chuang, T.H.; Arab Pour Yazdi, M.; Sanchette, F.; Billard, A. Microstructures and optoelectronic properties of CuxO films deposited by high-power impulse magnetron sputtering. J. Alloys Compd. 2016, 688, 672–678. [Google Scholar] [CrossRef]
  50. Sakalley, S.; Saravanan, A.; Cheng, W.C.; Chen, S.C.; Sun, H.; Liao, M.H.; Huang, B.R. Cu3N thin film synthesized by selective in situ substrate heating during high power impulse magnetron sputtering for augmenting UV photodetection. Sens. Actuator A Phys. 2023, 350, 114137. [Google Scholar] [CrossRef]
  51. Yao, Q.; Zhao, L.; Sun, X.; Zhu, L.; Zhao, Y.; Qiang, Y.; Song, J. Na2S decorated NiOx as effective hole transport layer for inverted planar perovskite solar cells. Mater. Sci. Semicond. Process. 2023, 153, 107107. [Google Scholar] [CrossRef]
  52. Zuo, C.; Ding, L. Solution-Processed Cu2O and CuO as Hole Transport Materials for Efficient Perovskite Solar Cells. Small 2015, 11, 5528–5532. [Google Scholar] [CrossRef] [PubMed]
  53. Makenali, M.; Kazeminezhad, I.; Ahmadi, V.; Roghabadi, F.A. Charge transfer balancing of planar perovskite solar cell based on a low cost and facile solution-processed CuOx as an efficient hole transporting layer. J. Mater. Sci. Mater. Electron. 2021, 32, 2312–2325. [Google Scholar] [CrossRef]
  54. Chatterjee, S.; Pal, A.J. Introducing Cu2O Thin Films as a Hole-Transport Layer in Efficient Planar Perovskite Solar Cell Structures. J. Phys. Chem. C 2016, 120, 1428–1437. [Google Scholar] [CrossRef]
  55. Nejand, B.A.; Ahmadi, V.; Gharibzadeh, S.; Shahverdi, H.R. Cuprous Oxide as a Potential Low-Cost Hole-Transport Material for Stable Perovskite Solar Cells. ChemSusChem 2016, 9, 302–313. [Google Scholar] [CrossRef]
  56. Chen, Y.J.; Li, M.H.; Huang, J.C.A.; Chen, P. The Cu/Cu2O nanocomposite as a p-type transparent-conductive-oxide for efficient bifacial-illuminated perovskite solar cells. J. Mater. Chem. C 2018, 6, 6280–6286. [Google Scholar] [CrossRef]
  57. Yu, W.; Li, F.; Wang, H.; Alarousu, E.; Chen, Y.; Lin, B.; Wang, L.; Hedhili, M.N.; Li, Y.; Wu, K.; et al. Ultrathin Cu2O as an efficient inorganic hole transporting material for perovskite solar cells. Nanoscale 2016, 8, 6173–6179. [Google Scholar] [CrossRef] [Green Version]
  58. Chen, L.C.; Chen, C.C.; Liang, K.C.; Chang, S.H.; Tseng, Z.L.; Yeh, S.C.; Chen, C.T.; Wu, W.T.; Wu, C.G. Nano-structured CuO-Cu2O Complex Thin Film for Application in CH3NH3PbI3 Perovskite Solar Cells. Nanoscale Res. Lett. 2016, 11, 402. [Google Scholar] [CrossRef] [Green Version]
  59. Liu, L.; Xi, Q.; Gao, G.; Yang, W.; Zhou, H.; Zhao, Y.; Wu, C.; Wang, L.; Xu, J. Cu2O particles mediated growth of perovskite for high efficient hole-transporting-layer free solar cells in ambient conditions. Sol. Energy Mater. Sol. Cells 2016, 157, 937–942. [Google Scholar] [CrossRef]
  60. Miao, X.; Wang, S.; Sun, W.; Zhu, Y.; Du, C.; Ma, R.; Wang, C. Effect of Cu2O Content in Electrodeposited CuOx Film on Perovskite Solar Cells. Nano 2019, 14, 1950126. [Google Scholar] [CrossRef] [Green Version]
  61. Zhang, Z.; Chen, S.; Li, P.; Li, H.; Wu, J.; Hu, P.; Wang, J. Aerosol-assisted chemical vapor deposition of ultra-thin CuOx films as hole transport material for planar perovskite solar cells. Funct. Mater. Lett. 2018, 11, 1850035. [Google Scholar] [CrossRef]
  62. Sajid, S.; Alzahmi, S.; Salem, I.B.; Obaidat, I.M. Guidelines for Fabricating Highly Efficient Perovskite Solar Cells with Cu2O as the Hole Transport Material. Nanomaterials 2022, 12, 3315. [Google Scholar] [CrossRef] [PubMed]
  63. Ravindra, P.; Mukherjee, R.; Avasthi, S. Hole-selective electron-blocking copper oxide contact for silicon solar cells. IEEE J. Photovolt. 2017, 7, 1278–1283. [Google Scholar] [CrossRef]
  64. Guo, Y.; Lei, H.; Xiong, L.; Li, B.; Chen, Z.; Wen, J.; Yang, G.; Li, G.; Fang, G. Single phase, high hole mobility Cu2O films as an efficient and robust hole transporting layer for organic solar cells. J. Mater. Chem. A 2017, 5, 11055–11062. [Google Scholar] [CrossRef]
  65. Lee, J.W.; Bae, S.H.; Hsieh, Y.T.; De Marco, N.; Wang, M.; Sun, P.; Yang, Y. A bifunctional lewis base additive for microscopic homogeneity in perovskite solar cells. Chem. 2017, 3, 290–302. [Google Scholar] [CrossRef]
  66. Hsieh, C.M.; Liao, Y.S.; Lin, Y.R.; Chen, C.P.; Tsai, C.M.; Diau, E.W.G.; Chuang, S.C. Low-temperature, simple and efficient preparation of perovskite solar cells using Lewis bases urea and thiourea as additives: Stimulating large grain growth and providing a PCE up to 18.8%. RSC Adv. 2018, 8, 19610–19615. [Google Scholar] [CrossRef] [Green Version]
Figure 1. X-ray diffraction spectra of Cu2O films deposited by (a) DCMS and (b) superimposed HiPIMS at different oxygen flow rates.
Figure 1. X-ray diffraction spectra of Cu2O films deposited by (a) DCMS and (b) superimposed HiPIMS at different oxygen flow rates.
Nanomaterials 13 01363 g001
Figure 2. (a) UV-Vis spectra and (b) AFM topography images of Cu2O films deposited using various processes.
Figure 2. (a) UV-Vis spectra and (b) AFM topography images of Cu2O films deposited using various processes.
Nanomaterials 13 01363 g002
Figure 3. Cu2O film prepared by HiPIMS (a) before plasma cleaning and (b) after plasma cleaning of the contact angle.
Figure 3. Cu2O film prepared by HiPIMS (a) before plasma cleaning and (b) after plasma cleaning of the contact angle.
Nanomaterials 13 01363 g003
Figure 4. J–V curves of PSCs and device structure.
Figure 4. J–V curves of PSCs and device structure.
Nanomaterials 13 01363 g004
Figure 5. UPS spectra of ITO/Cu2O films (DCMS, HiPIMS, and superimposed HiPIMS).
Figure 5. UPS spectra of ITO/Cu2O films (DCMS, HiPIMS, and superimposed HiPIMS).
Nanomaterials 13 01363 g005
Figure 6. Stability data of unsealed devices from superimposed HiPIMS under the argon glove box.
Figure 6. Stability data of unsealed devices from superimposed HiPIMS under the argon glove box.
Nanomaterials 13 01363 g006
Table 1. Different process technologies for solar cell devices.
Table 1. Different process technologies for solar cell devices.
SampleJsc (mA/cm2)Voc
(V)
FF
(%)
PCE
(%)
Best PCE
(%)
Rsa
(Ω cm2)
Rsha
(Ω cm2)
σ
(S cm−1)
DCMS13.74 ± 0.970.89 ± 0.0258.1 ± 3.937.12 ± 0.647.919.47111.820.11
HiPIMS17.36 ± 0.940.90 ± 0.0061.1 ± 6.959.42 ± 0.9210.293.091081.180.40
Superimposed HiPIMS19.27 ± 1.200.98 ± 0.0069.5 ± 4.9813.04 ± 1.6115.201.421080.193.33
Table 2. Performances of Cu2O HTLs in PSCs prepared by different process technologies in the literature.
Table 2. Performances of Cu2O HTLs in PSCs prepared by different process technologies in the literature.
Process TechnologyJsc
(mA/cm2)
Voc
(V)
FF
(%)
PCE
(%)
Year/Ref.
Spin Coating16.521.0775.513.352015/[52]
SILAR16.520.8956.08.232016/[54]
Rotating Sputtering15.800.9659.08.932016/[55]
Cu Thermal Oxidation17.50.9566.211.032016/[57]
Cu Thermal Oxidation14.400.9658.68.102016/[58]
Electrodeposition18.030.8861.09.642016/[59]
RF Sputtering14.940.9367.59.372018/[56]
CVD17.500.9848.08.262018/[61]
Electrodeposition19.030.9773.013.482019/[60]
Spin Coating12.870.7169.06.262021/[53]
Hydrothermal Technique12.580.7663.06.022022/[37]
Superimposed HiPIMS20.340.9876.515.20This work
Table 3. Superimposed HiPIMS device with dim light.
Table 3. Superimposed HiPIMS device with dim light.
SampleJsc (μA/cm2)Voc (V)FF (%)PCE (%)
Superimposed HiPIMS189.510.6266.425.09
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chuang, T.-H.; Chen, Y.-H.; Sakalley, S.; Cheng, W.-C.; Chan, C.K.; Chen, C.-P.; Chen, S.-C. Highly Stable and Enhanced Performance of p–i–n Perovskite Solar Cells via Cuprous Oxide Hole-Transport Layers. Nanomaterials 2023, 13, 1363. https://doi.org/10.3390/nano13081363

AMA Style

Chuang T-H, Chen Y-H, Sakalley S, Cheng W-C, Chan CK, Chen C-P, Chen S-C. Highly Stable and Enhanced Performance of p–i–n Perovskite Solar Cells via Cuprous Oxide Hole-Transport Layers. Nanomaterials. 2023; 13(8):1363. https://doi.org/10.3390/nano13081363

Chicago/Turabian Style

Chuang, Tung-Han, Yin-Hung Chen, Shikha Sakalley, Wei-Chun Cheng, Choon Kit Chan, Chih-Ping Chen, and Sheng-Chi Chen. 2023. "Highly Stable and Enhanced Performance of p–i–n Perovskite Solar Cells via Cuprous Oxide Hole-Transport Layers" Nanomaterials 13, no. 8: 1363. https://doi.org/10.3390/nano13081363

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