Next Article in Journal
Retraction: Kim, J.R. and Kim, J.J. Epoxy Resins Toughened with Surface Modified Epoxidized Natural Rubber Fibers by One-Step Electrospinning. Materials 2017, 10, 464, doi:10.3390/ma10050464
Next Article in Special Issue
Plasma-Exposure-Induced Mobility Enhancement of LiTFSI-Doped Spiro-OMeTAD Hole Transport Layer in Perovskite Solar Cells and Its Impact on Device Performance
Previous Article in Journal
Permeability versus Design in TPMS Scaffolds
Previous Article in Special Issue
Natural Dyes and Their Derivatives Integrated into Organic Solar Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Enhanced Open-Circuit Voltage in Perovskite Solar Cells with Open-Cage [60]Fullerene Derivatives as Electron-Transporting Materials

1
Department of Chemistry, University of Texas at El Paso El Paso, TX 79968, USA
2
Institut de Química Computacional i Catàlisi (IQCC), Department de Química, Universitat de Girona, 17003 Girona, Catalonia Spain
3
Department of Chemistry, Columbia University, New York, NY 10027, USA
*
Authors to whom correspondence should be addressed.
These authors contributed equally.
Materials 2019, 12(8), 1314; https://doi.org/10.3390/ma12081314
Submission received: 25 March 2019 / Revised: 15 April 2019 / Accepted: 18 April 2019 / Published: 23 April 2019
(This article belongs to the Special Issue Advances in Organic and Perovskite Solar Cells)

Abstract

:
The synthesis, characterization, and incorporation of open-cage [60]fullerene derivatives as electron-transporting materials (ETMs) in perovskite solar cells (PSCs) with an inverted planar (p-i-n) structure is reported. Following optical and electrochemical characterization of the open-cage fullerenes 2ac, p-i-n PSCs with a indium tin oxide (ITO)/poly(3,4-ethylenedioxythiophene)-polystyrene sulfonate (PEDOT:PSS)/perovskite/fullerene/Ag structure were prepared. The devices obtained from 2ab exhibit competitive power conversion efficiencies (PCEs) and improved open-circuit voltage (Voc) values (>1.0 V) in comparison to a reference cell based on phenyl-C61-butyric-acid methyl-ester (PC61BM). These results are rationalized in terms of a) the higher passivation ability of the open-cage fullerenes with respect to the other fullerenes, and b) a good overlap between the highest occupied molecular orbital/lowest unoccupied molecular orbital (HOMO/LUMO) levels of 2ab and the conduction band of the perovskite.

1. Introduction

Perovskite solar cells (PSCs) are an emerging class of photovoltaic devices, which promise to rival the performance of state-of-the-art cells, with current record power conversion efficiencies (PCEs) recently reaching 24.2% [1] A major advantage of PSCs is their facile manufacturing process, which is mostly based on solution processing. However, a number of challenges need to be addressed before a marketable technology is available, including (a) cell performance, (b) cell stability, and (c) upscaling beyond laboratory scale toward the industrial production of commercially viable photovoltaic devices [2,3,4,5,6,7]. Another desirable yet unmet objective is the replacement of Pb by less toxic metals in the perovskite structure [8].
A PSC consists of a sandwiched structure containing a transparent conductive oxide, a hole transport layer (HTL), a perovskite photo-absorber layer, an electron-transport layer (ETL), and a back-contact electrode [9,10,11,12,13,14,15]. Among the various cell configurations available for PSCs, the inverted planar structure (p-i-n) (Figure 1) is the most attractive in terms of manufacturing, because the ETL, which is typically a fullerene derivative, is solution-processed [16], as opposed to the metal oxides employed in the regular planar and mesoscopic configurations, which require high-temperature annealing steps [17,18,19]. Importantly, p-i-n PSCs can be easily integrated in flexible devices [20]. A downside of the p-i-n configuration is that high open circuit voltage (Voc) values are difficult to achieve. Successful strategies to overcome this limitation rely on the incorporation of dopant materials [21,22,23], interfacial engineering [24,25,26], morphology control [27,28], or the replacement of the HTL [29,30].
On the other hand, the replacement of phenyl-C61-butyric-acid methyl-ester (PC61BM) by other fullerenes as the ETL has found limited success, with only a few reports of Voc values beyond the 1.0 V threshold [31,32]. For the specific case of poly(3,4-ethylenedioxythiophene)-polystyrene sulfonate (PEDOT:PSS)-based devices, examples are even scarcer [11,33,34]. The development of PSCs with PC61BM surrogates relies mostly on the incorporation of highly crystalline ETLs. The resulting cells benefit from reduced energy disorder and improved charge trap passivation [35,36]. While this approach has been implemented with a variety of fullerenes, the vast majority of them consist of [3+2], [4+2], or cyclopropane adducts [15]. Thus, the introduction of structurally novel fullerene scaffolds that can expand the ETL repertoire is highly desirable.
Overall, the highest certified PCE value reported for PSCs with p-i-n configurations is 20.9% [37]. The development of efficient ETMs that can increase the efficiency of PSCs without the need of additives or complex manufacturing techniques is a major challenge in the field of inverted planar PSCs.
Open-cage fullerenes are a family of synthetic derivatives in which the three-dimensional backbone of the fullerene cage is distorted by the scission of one or more C–C bonds. Open-cage fullerenes have been successfully used as electron-accepting materials or as additives in bulk-heterojunction solar cells [38,39,40]. However, their use as ETLs in PSCs remains unexplored. Some of us recently reported a straightforward methodology for the synthesis of open-cage fullerene derivatives (Scheme 1) [41,42]. The promising electrochemical properties and remarkable stability of these compounds, together with the fact that fullerene derivatives are so far the materials of choice as the ETL for p-i-n PSCs, prompted us to study the use of these compounds for PSCs.
Herein, we report for the first time the incorporation of open-cage fullerenes in p-i-n PSCs. The cells have been thoroughly characterized, and their photovoltaic performance has been studied. We demonstrate that open-cage fullerenes exhibit improved performances with respect to PC61BM-based devices.

2. Results

The open-cage derivatives used in this study, 2ac, were prepared by the photochemical oxidation of bis(fulleroids) 1ac, which can be obtained from [60]fullerene in one step using our Rh(I)-catalyzed cycloaddition protocol (Scheme 1) [41]. Importantly, unlike precursors 1a-c, open-cage derivatives 2ac do not suffer further degradation upon exposure to light and air. In addition to the parent open-cage derivative 2a, we selected compounds possessing desirable features for PSC manufacturing, such as the improved solubility of 2b or the light-harvesting ability of 2c. With 2ac in hand, we first assessed their optical and electrochemical properties. In solution, 2ac display remarkable absorption maxima in the visible region (λmax = 705–710 nm, Figure S12). The electrochemical properties of 2ac (Figure S13) and PC61BM (3) were determined by cyclic voltammetry (CV) in ortho-dichlorobenzene (o-DCB) (see the Supporting Information).
Compounds 2ac exhibit three fully reversible cathodic electrochemical behaviors between −0.8 and −2.3 V at a scan rate of 100 mV s−1. The highest occupied molecular orbital/lowest unoccupied molecular orbital (HOMO/LUMO) values were estimated from the ultraviolet (UV) and CV measurements [43].
The optical properties of compounds 2ac are summarized in Table 1. Overall, the photophysical and electrochemical properties of the open-cage fullerenes 2a–c are very similar to those of PC61BM, even though the C60 cage skeleton is significantly altered with respect to the latter. These results encouraged us to incorporate 2ac as the electron-transporting materials (ETMs) in PSCs.
Figure 2a shows the energy level diagram estimated from the onset potential of the first reductions and the maximum onset absorption from UV-vis spectra for all the compounds [44]. The electrical conductivities for PC61BM and compounds 2ac films were compared by recording current–voltage (J-V) curves for electron-only devices with a structure of ITO/Al/ETM/Al. All the ETLs showed similar electron conductivities (4.8, 3.5, 2.8 and 3.7 × 10−4 cm2 V−1·s−1 for 2a, 2b, 2c, and PC61BM, respectively).
To probe the passivation ability of compounds 2ac, we studied the photoluminescence (PL) and time-resolved (TR) PL of the photoactive layer (perovskite) with and without 2ac, using PC61BM as the control (Figure 2b and Figure S14). A significant PL quenching effect was observed for the perovskite layer coated with the open-cage fullerenes 2a, 2b, and PC61BM (Figure 2b). Meanwhile, the PL intensity of the perovskite increases when using 2c, which is an effect that can be attributed to the lower solubility of 2c in chlorobenzene (CB) (Figure S15). Compound 2a exhibits a higher passivation ability than the other fullerenes, resulting in a more pronounced inhibition of the electron–hole recombination processes [45].
Figure S14 shows the TR-PL decay measurements, monitoring the emission peak of PC61BM and 2ac coated perovskite layers as a function of time. The pristine perovskite layer exhibits a PL lifetime of about 25.6 ns, whereas perovskite/2a, perovskite/2b, perovskite/2c, and perovskite/PC61BM exhibit PL lifetimes of 3.8 ns, 4.7 ns, 9.2 ns, and 14.1 ns, respectively. Faster decays are measured for the samples coated with 2a and 2b, indicating that the charge transfer processes are faster than the charge recombination in the perovskite layer [46].
Compounds 2ac were incorporated in PSCs with an ITO/PEDOT:PSS/perovskite/fullerene/Ag structure (Figure 1, see Supporting Information for details). The 2a (1.01 V) and 2b (0.97 V)-based devices showed a significant enhancement of Voc values compared to the PC61BM (0.92 V)-based devices.
On the other hand, the lower solubility of 2c led to low-quality films, resulting in lower photovoltaic performances for the 2c-based devices (Figure 3). Table 2 summarizes the performances of the PSC devices incorporating PC61BM and compounds 2ac. The work functions of the charge transport materials affect the Voc of PSCs significantly, so the higher Voc values obtained from 2a and 2b-based devices can be attributed to their higher LUMO values, when compared with PC61BM [25,26,46,47]. Commonly, the Voc values are improved by inserting a work-function interlayer between the perovskite and the ETL [26].
PSCs based on all the fullerene derivatives showed negligible hysteretic behavior (Figure S16). Device performance reproducibilities were calculated from the PCE distributions measured for 25 independent cells (Figure 3b). Figure S17 shows the external quantum efficiency (EQE) of the PSCs based on PC61BM and 2ac; the devices based on 2a show higher photoresponse around 600 nm and 750 nm. The integrated photocurrent densities based on EQE measurements (Figure S17) are consistent with those from J-V measurements (Table 2). PC61BM devices exhibited a PCE value of 16.22% with a Voc value of 0.92 V, a short circuit current (Jsc) value of 21.77 mA·cm−2, and a fill factor (FF) value of 0.80. In contrast, 2a devices exhibited a PCE value of 16.92% with a Voc value of 1.01 V, a Jsc value of 21.21 mA·cm−2, and a FF value of 0.79. The improved device performance was attributed to the better passivation ability of compounds 2a and 2b, because of their higher work function, which matches well with the conduction band of the perovskite [25,26,46,47].
The stabilities of PSCs fabricated with PC61BM, 2a, and 2b were monitored under ca. 25% humidity in air at room temperature without encapsulation for 10 days. The normalized PCEs against time are shown in Figure S18. PC61BM-based devices lost 67% of their initial PCE; this was similar to the PC61BM-based devices 2a and 2b-based devices, which lost 65% and 63% of their initial PCE, respectively. Meanwhile, the devices based on 2c lost 71% of their initial PCE. Thus, the open-cage compounds are comparable to PC61BM in terms of cell stability.

3. Conclusions

In conclusion, we have successfully prepared a series of p-i-n type PSCs incorporating dicarbonylic open-cage [60]fullerene derivatives 2ac as the ETL. For those compounds with appropriate solubility, the resulting PSCs offer performances rivaling or even superior to those of analogous cells employing the PC61BM reference. These results are commensurate with a good overlap between the HOMO/LUMO levels of fulleroids 2ab and the conduction band of the perovskite. The modularity of our synthetic approach to open-cage fullerene derivatives 2ac offers a promising opportunity to develop superior PSCs beyond this preliminary account.

Supplementary Materials

The following are available online at https://www.mdpi.com/1996-1944/12/8/1314/s1, Figure S-1. 1H NMR (400 MHz, CDCl3) of compound 3a. Figure S-2. 1H NMR (400 MHz, CDCl3) of compound 3b. Figure S-3. 1H NMR (400 MHz, CDCl3) of compound 3c. Figure S-4. 1H NMR (400 MHz, CDCl3) of compound 1a. Figure S-5. 1H NMR (400 MHz, CDCl3) of compound 1b. Figure S-6. 1H NMR (400 MHz, o-DCB-d4/CS2) of compound S-5. Figure S-7. 1H NMR (400 MHz, CDCl3) of compound 1c. Figure S-8. 1H NMR (400 MHz, CDCl3) of compound 2a. Figure S-9. 1H NMR (400 MHz, CDCl3) of compound 2b. Figure S-10. 13C NMR (100 MHz, CDCl3) of compound 2b. Figure S-11. 1H NMR (400 MHz, CDCl3) of compound 2c. Figure S-12. UV-vis spectra of compounds 2a-c and PC61BM. Figure S-13. Cyclic voltammetry of compounds 2a-c. Figure S-14. Time-resolved photoluminescence of perovskite, perovskite/compounds 2a-c and perovskite/PC61BM films. Figure S-15. Fullerene derivatives 2a-c and PC61BM in o-dichlorobezene (20 mg/mL). Figure S-16. J-V curves of the inverted PSCs based on PC61BM (a) and 2a,b (b and c, respectively) with respect to forward and reverse scan directions (the scanning rate was 100 mV/s). Figure S-17. EQE measurements for 2a-c and PC61BM-based devices. Figure S-18. Stability studies of 2a-c and PC61BM-based devices. Figure S-19. Top-view SEM image of the perovskite film.

Author Contributions

E.C., A.A, F.L. and F.P.: synthesis, characterization, and cells fabrication: E.C., A.A., F.L., F.P., A.P.-Q., A.R., A.L., X.-Y.Z. and L.E.: writing–original draft, review, and editing: L.E. overall supervision of the final work.

Funding

The authors thank the US National Science Foundation (NSF) for its generous support of this work under the NSF-PREM program (DMR 1205302, and CHE-1801317 to L. E.). The Robert A. Welch Foundation is also gratefully acknowledged for an endowed chair to L. E. (Grant AH-0033). Financial support by the Spanish Ministry of Economy and Competitivity (MINECO), European Social Fund (ESF) (Projects CTQ2017-85341-P and CTQ2017-83587-P, FPI predoctoral grant to A. A., RyC contract RYC2012-11112 to A. L.), and the Generalitat de Catalunya (Project 2017-SGR-39) is gratefully acknowledged. The EU has also funded this research under FEDER grant UNGI10-4E-801. X.-Y.Z thanks the Office of Naval Research for support under award no. N00014-16-1-2921. This research was supported in part by the Department of Energy (DOE) Office of Energy Efficiency and Renewable Energy (EERE) Postdoctoral Research Award under the EERE Solar Energy Technologies Office administered by the Oak Ridge Institute for Science and Education (ORISE) for the DOE. ORISE is managed by Oak Ridge Associated Universities (ORAU) under DOE contract number DE-SC00014664. All of the opinions expressed in this paper are the authors’ and do not necessarily reflect the policies and views of DOE, ORAU, or ORISE.

Conflicts of Interest

There are no conflicts to declare.

References

  1. NREL Best Research-Cell Efficiencies. Available online: https://www.nrel.gov/pv/assets/pdfs/best-research-cell-efficiencies-190416.pdf (accessed on 1 April 2019).
  2. Rong, Y.; Hu, Y.; Mei, A.; Tan, H.; Saidaminov, M.I.; Seok, S.I.; McGehee, M.D.; Sargent, E.H.; Han, H. Challenges for commercializing perovskite solar cells. Science 2018, 361, eaat8235. [Google Scholar] [CrossRef]
  3. Li, Z.; Klein, T.R.; Kim, D.H.; Yang, M.; Berry, J.J.; van Hest, M.F.A.M.; Zhu, K. Scalable fabrication of perovskite solar cells. Nat. Rev. Mater. 2018, 3, 18017. [Google Scholar] [CrossRef]
  4. Galagan, Y. Perovskite Solar Cells: Toward Industrial-Scale Methods. J. Phys. Chem. Lett. 2018, 9, 4326–4335. [Google Scholar] [CrossRef]
  5. Wang, F.; Cao, Y.; Chen, C.; Chen, Q.; Wu, X.; Li, X.; Qin, T.; Huang, W. Materials toward the upscaling of perovskite solar cells: progress, challenges, and strategies. Adv. Funct. Mater. 2018, 28, 1803753. [Google Scholar] [CrossRef]
  6. Abate, A.; Correa-Baena, J.-P.; Saliba, M.; Su’ait, M.S.; Bella, F. Perovskite Solar Cells: From the laboratory to the assembly line. Chem. Eur. J. 2018, 24, 3083–3100. [Google Scholar] [CrossRef] [PubMed]
  7. Bella, F.; Renzi, P.; Cavallo, C.; Gerbaldi, C. Caesium for perovskite solar cells: An overview. Chem. Eur. J. 2018, 24, 12183–12205. [Google Scholar] [CrossRef] [PubMed]
  8. Giustino, F.; Snaith, H.J. Toward Lead-free perovskite solar cells. ACS Energy Lett. 2016, 1, 1233–1240. [Google Scholar] [CrossRef]
  9. Castro, E.; Fernandez-Delgado, O.; Arslan, F.; Zavala, G.; Yang, T.; Seetharaman, S.; Dsouza, F.; Echegoyen, L. New thiophene-based C60 fullerene derivatives as efficient electron transporting materials for perovskite solar cells. New J. Chem. 2018, 42, 14551–14558. [Google Scholar] [CrossRef] [PubMed]
  10. Castro, E.; Sisto, T.J.; Romero, E.L.; Liu, F.; Peurifoy, S.R.; Wang, J.; Zhu, X.; Nuckolls, C.; Echegoyen, L. Cove-Edge Nanoribbon materials for efficient inverted halide perovskite solar cells. Angew. Chem. Int. Ed. 2017, 129, 14840–14844. [Google Scholar] [CrossRef]
  11. Castro, E.; Zavala, G.; Seetharaman, S.; D’Souza, F.; Echegoyen, L. Impact of fullerene derivative isomeric purity on the performance of inverted planar perovskite solar cells. J. Mater. Chem. A. 2017, 5, 19485–19490. [Google Scholar] [CrossRef]
  12. Peurifoy, S.R.; Castro, E.; Liu, F.; Zhu, X.Y.; Ng, F.; Jockusch, S.; Steigerwald, M.L.; Echegoyen, L.; Nuckolls, C.; Sisto, T.J. Three-dimensional graphene nanostructures. J. Am. Chem. Soc. 2018, 140, 9341–9345. [Google Scholar] [CrossRef] [PubMed]
  13. Tian, C.; Castro, E.; Betancourt-Solis, G.; Nan, Z.-A.; Fernandez-Delgado, O.; Jankuru, S.; Echegoyen, L. Fullerene derivative with a branched alkyl chain exhibits enhanced charge extraction and stability in inverted planar perovskite solar cells. New J. Chem. 2018, 42, 2896–2902. [Google Scholar] [CrossRef]
  14. Tian, C.; Castro, E.; Wang, T.; Betancourt-Solis, G.; Rodriguez, G.; Echegoyen, L. Improved performance and stability of inverted planar perovskite solar cells using fulleropyrrolidine layers. ACS Appl. Mater. Interfaces. 2016, 8, 31426–31432. [Google Scholar] [CrossRef]
  15. Tian, C.; Kochiss, K.; Castro, E.; Betancourt-Solis, G.; Han, H.; Echegoyen, L. A dimeric fullerene derivative for efficient inverted planar perovskite solar cells with improved stability. J. Mater. Chem. A. 2017, 5, 7326–7332. [Google Scholar] [CrossRef]
  16. Khadka, D.B.; Shirai, Y.; Yanagida, M.; Miyano, K. Unraveling the impacts induced by organic and inorganic hole transport layers in inverted halide perovskite solar cells. ACS Appl. Mater. Interfaces. 2019, 11, 7055–7065. [Google Scholar] [CrossRef]
  17. Castro, E.; Murillo, J.; Fernandez-Delgado, O.; Echegoyen, L. Progress in fullerene-based hybrid perovskite solar cells. J. Mater. Chem. C. 2018, 6, 2635–2651. [Google Scholar] [CrossRef]
  18. Fang, Y.; Bi, C.; Wang, D.; Huang, J. The functions of fullerenes in hybrid perovskite solar cells. ACS Energy Lett. 2017, 2, 782–794. [Google Scholar] [CrossRef]
  19. 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]
  20. Wang, Y.-C.; Li, X.; Zhu, L.; Liu, X.; Zhang, W.; Fang, J. Efficient and hysteresis-free perovskite solar cells based on a solution processable polar fullerene electron transport layer. Adv. Energy Mater. 2017, 7, 1701144. [Google Scholar] [CrossRef]
  21. Xia, X.; Jiang, Y.; Wan, Q.; Wang, X.; Wang, L.; Li, F. Lithium and silver co-doped nickel oxide hole-transporting layer boosting the efficiency and stability of inverted planar perovskite solar cells. ACS Appl. Mater. Interfaces 2018, 10, 44501–44510. [Google Scholar] [CrossRef] [PubMed]
  22. Wang, Y.; Wang, S.; Chen, X.; Li, Z.; Wang, J.; Li, T.; Deng, X. Largely enhanced VOC and stability in perovskite solar cells with modified energy match by coupled 2D interlayers. J. Mater. Chem. A. 2018, 6, 4860–4867. [Google Scholar] [CrossRef]
  23. Chen, W.; Zhou, Y.; Wang, L.; Wu, Y.; Tu, B.; Yu, B.; Liu, F.; Tam, H.-W.; Wang, G.; Djurišić, A.B.; Huang, L.; He, Z. Molecule-doped nickel oxide: Verified charge transfer and planar inverted mixed cation perovskite solar cell. Adv. Mater. 2018, 30, 1800515. [Google Scholar] [CrossRef] [PubMed]
  24. Chen, J.; Lian, X.; Zhang, Y.; Yang, W.; Li, J.; Qin, M.; Lu, X.; Wu, G.; Chen, H. Interfacial engineering enables high efficiency with a high open-circuit voltage above 1.23 V in 2D perovskite solar cells. J. Mater. Chem. A. 2018, 6, 18010–18017. [Google Scholar] [CrossRef]
  25. Xue, Q.; Bai, Y.; Liu, M.; Xia, R.; Hu, Z.; Chen, Z.; Jiang, X.-F.; Huang, F.; Yang, S.; Matsuo, Y.; Yip, H.-L.; Cao, Y. Dual interfacial modifications enable high performance semitransparent perovskite solar cells with large open circuit voltage and fill factor. Adv. Energy Mater. 2017, 7, 1602333. [Google Scholar] [CrossRef]
  26. Bai, Y.; Yu, H.; Zhu, Z.; Jiang, K.; Zhang, T.; Zhao, N.; Yang, S.; Yan, H. High performance inverted structure perovskite solar cells based on a PCBM:polystyrene blend electron transport layer. J. Mater. Chem. A. 2015, 3, 9098–9102. [Google Scholar] [CrossRef]
  27. Cheng, J.; Zhang, H.; Zhang, S.; Ouyang, D.; Huang, Z.; Nazeeruddin, M.K.; Hou, J.; Choy, W.C.H. Highly efficient planar perovskite solar cells achieved by simultaneous defect engineering and formation kinetic control. J. Mater. Chem. A. 2018, 6, 23865–23874. [Google Scholar] [CrossRef]
  28. Li, W.; Wu, X.; Qin, H.; Zhao, Z.; Liu, H. Light-Driven and Light-Guided Microswimmers. Adv. Funct. Mater. 2016, 26, 3164–3171. [Google Scholar] [CrossRef]
  29. Zhang, H.; Wang, H.; Zhu, H.; Chueh, C.-C.; Chen, W.; Yang, S.; Jen, A.K.-Y. Low-temperature solution-processed CuCrO2 hole-transporting layer for efficient and photostable perovskite solar cells. Adv. Energy Mater. 2018, 8, 1702762. [Google Scholar] [CrossRef]
  30. Wang, H.; Yu, Z.; Lai, J.; Song, X.; Yang, X.; Hagfeldt, A.; Sun, L. One plus one greater than two: High-performance inverted planar perovskite solar cells based on a composite CuI/CuSCN hole-transporting layer. J. Mater. Chem. A. 2018, 6, 21435–21444. [Google Scholar] [CrossRef]
  31. Yan, K.; Chen, J.; Ju, H.; Ding, F.; Chen, H.; Li, C.-Z. Achieving high-performance thick-film perovskite solar cells with electron transporting Bingel fullerenes. J. Mater. Chem. A. 2018, 6, 15495–15503. [Google Scholar] [CrossRef]
  32. Meng, X.; Bai, Y.; Xiao, S.; Zhang, T.; Hu, C.; Yang, Y.; Zheng, X.; Yang, S. Designing new fullerene derivatives as electron transporting materials for efficient perovskite solar cells with improved moisture resistance. Nano Energy. 2016, 30, 341–346. [Google Scholar] [CrossRef]
  33. Gil-Escrig, L.; Momblona, C.; Sessolo, M.; Bolink, H.J. Fullerene imposed high open-circuit voltage in efficient perovskite based solar cells. J. Mater. Chem. A. 2016, 4, 3667–3672. [Google Scholar] [CrossRef]
  34. Chiang, C.-H.; Tseng, Z.-L.; Wu, C.-G. Planar heterojunction perovskite/PC71BM solar cells with enhanced open-circuit voltage via a (2/1)-step spin-coating process. J. Mater. Chem. A. 2014, 2, 15897–15903. [Google Scholar] [CrossRef]
  35. Xie, F.; Zhang, L.; Su, D.; Jaroniec, M.; Qiao, S.-Z. Na2Ti3O7@N-Doped carbon hollow spheres for sodium-ion batteries with excellent rate performance. Adv. Mater. 2017, 29, 1700989. [Google Scholar] [CrossRef] [PubMed]
  36. Khadka, D.B.; Shirai, Y.; Yanagida, M.; Noda, T.; Miyano, K. Tailoring the open-circuit voltage deficit of wide-band-gap perovskite solar cells using alkyl chain-substituted fullerene derivatives. ACS Appl. Mater. interfaces. 2018, 10, 22074–22082. [Google Scholar] [CrossRef]
  37. Luo, D.; Yang, W.; Wang, Z.; Sadhanala, A.; Hu, Q.; Su, R.; Shivanna, R.; Trindade, G.F.; Watts, J.F.; Xu, Z.; et al. Enhanced photovoltage for inverted planar heterojunction perovskite solar cells. Science 2018, 360, 1442–1446. [Google Scholar] [CrossRef]
  38. Chen, C.-P.; Huang, C.-Y.; Chuang, S.-C. Highly thermal stable and efficient organic photovoltaic cells with crosslinked networks appending open-cage fullerenes as additives. Adv. Funct. Mater. 2015, 25, 207–213. [Google Scholar] [CrossRef]
  39. Murata, M.; Morinaka, Y.; Murata, Y.; Yoshikawa, O.; Sagawa, T.; Yoshikawa, S. Modification of the σ-framework of [60]fullerene for bulk-heterojunction solar cells. Chem. Commun. 2011, 47, 7335–7337. [Google Scholar] [CrossRef]
  40. Chen, C.-P.; Lin, Y.-W.; Horng, J.-C.; Chuang, S.-C. Open-cage fullerenes as n-type materials in organic photovoltaics: Relevance of frontier energy levels, carrier mobility and morphology of different sizable open-cage fullerenes with power conversion efficiency in devices. Adv. Energy Mater. 2011, 1, 776–780. [Google Scholar] [CrossRef]
  41. Artigas, A.; Pla-Quintana, A.; Lledó, A.; Roglans, A.; Solà, M. Expeditious preparation of open-cage fullerenes by Rhodium(I)-catalyzed [2+2+2] cycloaddition of diynes and C60: An experimental and theoretical study. Chem. Eur. J. 2018, 24, 10653–10661. [Google Scholar] [CrossRef]
  42. Artigas, A.; Lledó, A.; Pla-Quintana, A.; Roglans, A.; Solà, M. Cover feature: A computational study of the intermolecular [2+2+2] cycloaddition of acetylene and C60 catalyzed by wilkinson’s catalyst. Chem. Eur. J. 2017, 23, 14977. [Google Scholar] [CrossRef]
  43. Closs, G.L.; Gautam, P.; Zhang, D.; Krusic, P.J.; Hill, S.A.; Wasserman, E. Steady-state and time-resolved direct detection EPR spectra of fullerene triplets in liquid solution and glassy matrixes: evidence for a dynamic Jahn-Teller effect in triplet C60. J. Phys. Chem. 1992, 96, 5228–5231. [Google Scholar] [CrossRef]
  44. Sun, Q.J.; Wang, H.Q.; Yang, C.H.; Li, Y.F. Synthesis and electroluminescence of novel copolymers containing crown ether spacers. J. Mater. Chem. 2003, 13, 800–806. [Google Scholar] [CrossRef]
  45. Liang, P.-W.; Liao, C.-Y.; Chueh, C.-C.; Zuo, F.; Williams, S.T.; Xin, X.-K.; Lin, J.; Jen, A.K.Y. Additive enhanced crystallization of solution-processed perovskite for highly efficient planar-heterojunction solar cells. Adv. Mater. 2014, 26, 3748–3754. [Google Scholar] [CrossRef]
  46. Wu, C.-G.; Chiang, C.-H.; Chang, S.H. A perovskite cell with a record-high-Voc of 1.61 V based on solvent annealed CH3NH3PbBr3/ICBA active layer. Nanoscale 2016, 8, 4077–4085. [Google Scholar] [CrossRef]
  47. Chen, S.; Hou, Y.; Chen, H.; Richter, M.; Guo, F.; Kahmann, S.; Tang, X.; Stubhan, T.; Zhang, H.; Li, N.; et al. Exploring the limiting open-circuit voltage and the voltage loss mechanism in planar ch3nh3pbbr3 perovskite solar cells. Adv. Energy Mater. 2016, 6, 1600132. [Google Scholar] [CrossRef]
Figure 1. Inverted planar perovskite solar cell (PSC) representation with the structure of indium tin oxide (ITO)/poly(3,4-ethylenedioxythiophene)-polystyrene sulfonate (PEDOT:PSS)/perovskite/electron-transport layer (ETL)/silver.
Figure 1. Inverted planar perovskite solar cell (PSC) representation with the structure of indium tin oxide (ITO)/poly(3,4-ethylenedioxythiophene)-polystyrene sulfonate (PEDOT:PSS)/perovskite/electron-transport layer (ETL)/silver.
Materials 12 01314 g001
Scheme 1. Synthesis of open-cage fullerene derivatives 2ac, and (inset) reference compound used in this study ((3, PC61BM), See Supporting Information for the synthesis of 1c).
Scheme 1. Synthesis of open-cage fullerene derivatives 2ac, and (inset) reference compound used in this study ((3, PC61BM), See Supporting Information for the synthesis of 1c).
Materials 12 01314 sch001
Figure 2. (a) Schematic illustration of the estimated HOMO and LUMO energy levels, estimated from cyclic voltammetry (CV) and UV-vis. (b) Steady-state photoluminescence (PL) spectra of the perovskite and perovskite/ETM films.
Figure 2. (a) Schematic illustration of the estimated HOMO and LUMO energy levels, estimated from cyclic voltammetry (CV) and UV-vis. (b) Steady-state photoluminescence (PL) spectra of the perovskite and perovskite/ETM films.
Materials 12 01314 g002
Figure 3. (a) Current–voltage (J-V) curves under 1 sun of illumination (100 mW/cm2) in forward voltage scans. (b) The power conversion efficiency (PCE) histograms measured for 25 independent cells.
Figure 3. (a) Current–voltage (J-V) curves under 1 sun of illumination (100 mW/cm2) in forward voltage scans. (b) The power conversion efficiency (PCE) histograms measured for 25 independent cells.
Materials 12 01314 g003
Table 1. Optical bandgap, onset reduction, and lowest unoccupied molecular orbital/highest occupied molecular orbital (LUMO/HOMO) energy levels of the open-cage 2ac and compound 3.
Table 1. Optical bandgap, onset reduction, and lowest unoccupied molecular orbital/highest occupied molecular orbital (LUMO/HOMO) energy levels of the open-cage 2ac and compound 3.
Compλmax (nm)Eg (ev)Ered (V)LUMO (ev)HOMO (ev)
2a7091.750.94−3.86−5.61
2b7051.760.99−3.81−5.57
2c7051.760.85−3.95−5.71
PC61BM7181.730.87−3.93−5.66
Table 2. Summary of device performance. The calculated short circuit current (Jsc) values were obtained from the external quantum efficiency (EQE) curves. Values in parentheses represent the best values measured, a are the average values, and * are the calculated values. PC61BM: phenyl-C61-butyric-acid methyl-ester.
Table 2. Summary of device performance. The calculated short circuit current (Jsc) values were obtained from the external quantum efficiency (EQE) curves. Values in parentheses represent the best values measured, a are the average values, and * are the calculated values. PC61BM: phenyl-C61-butyric-acid methyl-ester.
CompoundJsc* (mA cm−2)Jsc (mA cm−2)Voc (V)FF (%)PCE (%)
2a21.0521.17 a (21.21)0.99 a (1.01)0.7916.30 ± 0.62 (16.92)
2b21.1120.98 a (21.37) 0.96 a (0.97)0.7915.77 ± 0.60 (16.37)
2c14.8715.01 a (15.20)0.53 a (0.60)0.413.07 ± 0.67 (3.74)
PC61BM21.2221. 56 a (21.77)0.88 a (0.92)0.8015.66 ± 0.56 (16.22)

Share and Cite

MDPI and ACS Style

Castro, E.; Artigas, A.; Pla-Quintana, A.; Roglans, A.; Liu, F.; Perez, F.; Lledó, A.; Zhu, X.-Y.; Echegoyen, L. Enhanced Open-Circuit Voltage in Perovskite Solar Cells with Open-Cage [60]Fullerene Derivatives as Electron-Transporting Materials. Materials 2019, 12, 1314. https://doi.org/10.3390/ma12081314

AMA Style

Castro E, Artigas A, Pla-Quintana A, Roglans A, Liu F, Perez F, Lledó A, Zhu X-Y, Echegoyen L. Enhanced Open-Circuit Voltage in Perovskite Solar Cells with Open-Cage [60]Fullerene Derivatives as Electron-Transporting Materials. Materials. 2019; 12(8):1314. https://doi.org/10.3390/ma12081314

Chicago/Turabian Style

Castro, Edison, Albert Artigas, Anna Pla-Quintana, Anna Roglans, Fang Liu, Frank Perez, Agustí Lledó, X.-Y. Zhu, and Luis Echegoyen. 2019. "Enhanced Open-Circuit Voltage in Perovskite Solar Cells with Open-Cage [60]Fullerene Derivatives as Electron-Transporting Materials" Materials 12, no. 8: 1314. https://doi.org/10.3390/ma12081314

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