• Journal of Semiconductors
  • Vol. 44, Issue 2, 020202 (2023)
Dongmei He1, Shirong Lu2, Juan Hou3、*, Cong Chen4、**, Jiangzhao Chen1、***, and Liming Ding5、****
Author Affiliations
  • 1Key Laboratory of Optoelectronic Technology & Systems (MoE), College of Optoelectronic Engineering, Chongqing University, Chongqing 400044, China
  • 2Department of Material Science and Technology, Taizhou University, Taizhou 318000, China
  • 3Department of Physics, Shihezi University, Shihezi 832003, China
  • 4State Key Laboratory of Reliability and Intelligence of Electrical Equipment, School of Materials Science and Engineering, Hebei University of Technology, Tianjin 300401, China
  • 5Center for Excellence in Nanoscience (CAS), Key Laboratory of Nanosystem and Hierarchical Fabrication (CAS), National Center for Nanoscience and Technology, Beijing 100190, China
  • show less
    DOI: 10.1088/1674-4926/44/2/020202 Cite this Article
    Dongmei He, Shirong Lu, Juan Hou, Cong Chen, Jiangzhao Chen, Liming Ding. Doping organic hole-transport materials for high-performance perovskite solar cells[J]. Journal of Semiconductors, 2023, 44(2): 020202 Copy Citation Text show less

    Abstract

    I can easily diffuse into HTL and react with positively charged radicals, which would deteriorate device performance. Besides, Li+ ions can also diffuse into perovskite layer. To overcome these issues, Yanget al. reported a Li-free doping strategy by coupling positive polymer radicals with molecular anion 1,1,2,2,3,3-hexafluoropropane-1,3-disulfonimide (HFDF)via an ion exchange process[16]. The molecular structures for PTAA, F4TCNQ and LiHFDF are exhibited inFig. 2(a). The doping process can be completed in a few minutes. And the doping mechanism can be depicted as follows:

    (Color online) (a) Comparison between the conventional and ion-modulated (IM) radical doping strategies. (b)J–V characteristics for SnO2-based PSCs (under different doping). (c)J–V curves for TiO2-based PSCs (conventional dopingvs IM radical doping). (d) Moisture stability for unencapsulated PSCs under 70 ± 5% humidity (conventional dopingvs IM radical doping). (e) Thermal stability for the unsealed devices at 70 ± 3 °C. Reproduced with permission[2], Copyright 2022, American Association for the Advancement of Science.

    Figure 1.(Color online) (a) Comparison between the conventional and ion-modulated (IM) radical doping strategies. (b)J–V characteristics for SnO2-based PSCs (under different doping). (c)J–V curves for TiO2-based PSCs (conventional dopingvs IM radical doping). (d) Moisture stability for unencapsulated PSCs under 70 ± 5% humidity (conventional dopingvs IM radical doping). (e) Thermal stability for the unsealed devices at 70 ± 3 °C. Reproduced with permission[2], Copyright 2022, American Association for the Advancement of Science.

    PTAA+F4TCNQ[PTAA+F4TCNQ],

    Some p-type dopants were used to fully supersede tBP and LiTFSI[33,34]. Jiaet al. doped PTAA HTL with a fluorine-containing hydrophobic Lewis acid and achieved a higher PCE of 19.01% than 17.77% for control devices doped by bi-dopants LiTFSI/tBP[33]. Moreover, smaller hysteresis and improved stability were demonstrated. Recently, Nazeeruddinet al. proposed a novel doping approach by employing DIC-PBA dopant with a diphenyl iodide cation and pentafluorophenyl boric acid anion[34]. The devices gave a small-area PCE of 22.86% and a module PCE of 19.13% (33.2 cm2) along with increased ambient stability. DIC-PBA can simultaneously p-dope PTAA and perovskite, which was attributed to ionic interaction-derived dipole arrangement. These results indicate that the total substitution strategy is promising but the PCE needs to be further enhanced. To further ameliorate doping, more doping molecules should be designed. The correlation between molecular structures and hole conductivity, energy levels, interfacial carrier dynamics as well as device performance should be established.

    [PTAA+F4TCNQ]+LiHFDF[PTAA+HFDF]+Li++F4TCNQ.

    Gaoet al. developed an ion-modulated (IM) radical doping strategy where pre-synthesized organic radicals spiro-OMeTAD2·+(TFSI)2 and TBMP+TFSI salt were used to dope spiro-OMeTAD HTL (Fig. 1(a))[2]. Spiro-OMeTAD can be immediately oxidized by spiro-OMeTAD2·+(TFSI)2 into spiro-OMeTAD·+TFSI monoradicalvia comproportionation. The radicals can instantly augment the conductivity and workfunction (WF) through providing hole polarons. In the meantime, ionic salts can further tune WFvia affecting the energetics of hole polarons. A very long oxidation time (~24 h) was required to reach optimal PCE for the devices with conventional doping recipe (Fig. 1(b)). In comparison, the PCE for the device with IM radical doping instantly reaches maximum. Through IM radical doping, FAPbI3 cells offered a PCE of 25.15% (Fig. 1(c)). Under high relative humidity of ~70 ± 5%, the device with IM radical doping showed aT80 of ~1240 h whileT80 was only ~96 h for the device with conventional doping (Fig. 1(d)). The markedly improved thermal stability was also realized as confirmed by much largerT80 value for the target device with radical doping (~796 h) as compared to the control device (~264 h) (Fig. 1(e)). The generality for IM radical doping was revealed by comparing the photovoltaic performance of the devices based on different perovskite compositions and various organic salts with different cations and anions. Developing alternative dopant recipe could improve device performance.

    In order to mitigate hygroscopic problem of LiTFSI dopant, some p-type dopants with better hydrophobic property (e.g., metal organic complex[20-22]) were added into HTL solution. In addition, fluorine substitution in lithium salt was attempted to strengthen the moisture stability of HTLs and device[23]. However, the above methods can not solve thoroughly the instability issue induced by hygroscopic LiTFSI. Given this, a variety of alternative p-dopants to LiTFSI, such as protic ionic liquids[24,25], metal salts[26-28], and ex situ synthesized spiro-OMeTAD2·+(TFSI)2 radicals[2,29,30], have been explored. Nevertheless, to guarantee doping effect, tBP is imperative during the doping process. It should be pointed out that whether tBP doping is conducive to ameliorating device stability is still debatable[31,32]. By-products can be formed through the coordination between tBP and LiTFSI[31], and the generated radicals can be consumedvia its interaction with tBP[32]. The role of tBP in doping process needs further investigation to address the instability issue of classical doping recipe.

    [PTAA·+F4TCNQ·−] was firstly generated by the reaction between PTAA and F4TCNQ. Subsequently, the ion exchange between F4TCNQ·− in [PTAA·+F4TCNQ·−] and HFDF led to the formation of a complex [PTAA·+HFDF·−] (named HFDF-HTL). The doping efficiency was prominently enhanced for the HFDF-HTL compared with traditional LiTFSI/tBP-doped HTL (Li-HTL) as evidenced by 80 times greater conductivity of the former. HFDF-HTL can maintain high hole conductivity and excellent energy band alignment upon extreme iodide invasion. The ion exchange doping strategy enabled the fabrication of PSCs with a PCE of 24.0% (certified 23.9%), much higher than 20.5% for the device with Li-HTL (Figs. 2(b) and2(c)). The unsealed devices were evaluated under AM1.5G radiation and ambient humidity of ~50% (Fig. 2(d)). The device with HFDF-HTL degraded by about 10% after aging for 576 h while the device with Li-HTL degraded by 60% after only 100 h. As shown inFig. 2(e), 92% of the original PCE was maintained for the device with HFDF-HTL after 1000 h aging while the device with Li-HTL exhibited a 49% drop in PCE. The HTL can operate stably under iodide intrusion through developing effective doping strategy.

    Most efforts now focus on improving the hydrophobicity of HTL by developing hydrophobic dopants to replace LiTFSI. The moisture instability is easily resolved by encapsulation technology. The instability caused by migration and diffusion of Li+ (from HTL to perovskite layer) and I (from perovskite layer to HTL) is more challenging. More effective doping strategy should be developed.

    (Color online) (a) Molecular structures for PTAA, F4TCNQ and LiHFDF. (b) Cross-sectional SEM image for PSCs with HFDF-HTL. (c)J–V curves for PSCs (Li-HTLvs HFDF-HTL). (d) Moisture stability for unsealed PSCs under AM1.5G radiation and ~50% RH (Li-HTLvs HFDF-HTL). (e) Thermal stability for the encapsulated devices with different HTLs under AM1.5G illumination at 85 °C. Reproduced with permission[16], Copyright 2022, American Association for the Advancement of Science.

    Figure 2.(Color online) (a) Molecular structures for PTAA, F4TCNQ and LiHFDF. (b) Cross-sectional SEM image for PSCs with HFDF-HTL. (c)J–V curves for PSCs (Li-HTLvs HFDF-HTL). (d) Moisture stability for unsealed PSCs under AM1.5G radiation and ~50% RH (Li-HTLvs HFDF-HTL). (e) Thermal stability for the encapsulated devices with different HTLs under AM1.5G illumination at 85 °C. Reproduced with permission[16], Copyright 2022, American Association for the Advancement of Science.

    Single-junction and tandem perovskite solar cells (PSCs) have achieved impressive power conversion efficiencies (PCEs) of 25.7% and 31.3%, respectively, which makes it to be one of next-generation photovoltaic technologies[1-9]. Interface engineering[3,5,10-12], composition engineering[13] and additive engineering[7,14,15] have made remarkable contributions to efficiency enhancement. Compared with efficiency, the long-term operational stability of PSCs jogs along, which is far from the requirements of commercialization. Currently, almost all regular n–i–p PSCs were accomplished with classical organic hole-transport materials (HTMs), i.e., PTAA[16] and spiro-OMeTAD[2,4,6]. However, the highly efficient PSCs with the above organic hole-transport layers (HTL) usually suffer from instability. To facilitate hole transport and extraction, LiTFSI and tBP are frequently employed to dope organic HTLs but this would sacrifice device stability. The use of these hygroscopic p-dopants endows the devices with poor moisture stability. It is worth noting that small-sized lithium ion (Li+) can easily diffuse into perovskite layer and metal electrode, which deteriorates device performance[17]. Consequently, a critical challenge limiting commercial applications of PSCs is the trade-off between high efficiency and high stability. The traditional doping strategy with LiTFSI and tBP requires a long time (usually several days) to reach optimal doping and device performance, which is not good for mass production. In addition, owing to the intrinsic soft nature of perovskites, iodide ions can easily migrate and diffuse into the HTL[18,19], and then interact with positive radicals in HTLs, diminishing hole transport[16].

    References

    [1] National Renewable Energy Laboratory. Best Research Cell Efficiencies. 2022

    [2] T Zhang, F Wang, H B Kim et al. Ion-modulated radical doping of spiro-OMeTAD for more efficient and stable perovskite solar cells. Science, 377, 495(2022).

    [3] Z Li, B Li, X Wu et al. Organometallic-functionalized interfaces for highly efficient inverted perovskite solar cells. Science, 376, 416(2022).

    [4] M Kim, J Jeong, H Lu et al. Conformal quantum dot-SnO2 layers as electron transporters for efficient perovskite solar cells. Science, 375, 302(2022).

    [5] Q Jiang, J Tong, Y Xian et al. Surface reaction for efficient and stable inverted perovskite solar cells. Nature, 611, 278(2022).

    [6] Y Zhao, F Ma, Z Qu et al. Inactive (PbI2)2RbCl stabilizes perovskite films for efficient solar cells. Science, 377, 531(2022).

    [7] J Jeong, M Kim, J Seo et al. Pseudo-halide anion engineering for α-FAPbI3 perovskite solar cells. Nature, 592, 381(2021).

    [8] A Al-Ashouri, E Köhnen, B Li et al. Monolithic perovskite/silicon tandem solar cell with >29% efficiency by enhanced hole extraction. Science, 370, 1300(2020).

    [9] L Zhang, X Pan, L Liu et al. Star perovskite materials. J Semicond, 43, 030203(2022).

    [10] Z Fang, X Meng, C Zuo et al. Interface engineering gifts CsPbI2.25Br0.75 solar cells high performance. Sci Bull, 64, 1743(2019).

    [11] Z Zhang, J Li, Z Fang et al. Adjusting energy level alignment between HTL and CsPbI2Br to improve solar cell efficiency. J Semicond, 42, 030501(2021).

    [12] M Cheng, C Zuo, Y Wu et al. Charge-transport layer engineering in perovskite solar cells. Sci Bull, 65, 1237(2020).

    [13] G Kim, H Min, K S Lee et al. Impact of strain relaxation on performance of α-formamidinium lead iodide perovskite solar cells. Science, 370, 108(2020).

    [14] L Zhu, X Zhang, M Li et al. Trap state passivation by rational ligand molecule engineering toward efficient and stable perovskite solar cells exceeding 23% efficiency. Adv Energy Mater, 11, 2100529(2021).

    [15] M Li, H Li, Q Zhuang et al. Stabilizing perovskite precursor by synergy of functional groups for niox-based inverted solar cells with 23.5 % efficiency. Angew Chem Int Ed, 61, e202206914(2022).

    [16] T Wang, Y Zhang, W Kong et al. Transporting holes stably under iodide invasion in efficient perovskite solar cells. Science, 377, 1227(2022).

    [17] Z Li, C Xiao, Y Yang et al. Extrinsic ion migration in perovskite solar cells. Energ Environ Sci, 10, 1234(2017).

    [18] Y Wang, T Wu, J Barbaud et al. Stabilizing heterostructures of soft perovskite semiconductors. Science, 365, 687(2019).

    [19] E Bi, H Chen, F Xie et al. Diffusion engineering of ions and charge carriers for stable efficient perovskite solar cells. Nat Commun, 8, 15330(2017).

    [20] L Badia, E Mas-Marzá, R S Sánchez et al. New iridium complex as additive to the spiro-OMeTAD in perovskite solar cells with enhanced stability. APL Mater, 2, 081507(2014).

    [21] T M Koh, S Dharani, H Li et al. Cobalt dopant with deep redox potential for organometal halide hybrid solar cells. ChemSusChem, 7, 1909(2014).

    [22] J Zhang, Q Daniel, T Zhang et al. Chemical dopant engineering in hole transport layers for efficient perovskite solar cells: insight into the interfacial recombination. ACS Nano, 12, 10452(2018).

    [23] T Wu, R Zhuang, R Zhao et al. Understanding the effects of fluorine substitution in lithium salt on photovoltaic properties and stability of perovskite solar cells. ACS Energy Lett, 6, 2218(2021).

    [24] J Zhang, T Zhang, L Jiang et al. 4-tert-butylpyridine free hole transport materials for efficient perovskite solar cells: A new strategy to enhance the environmental and thermal stability. ACS Energy Lett, 3, 1677(2018).

    [25] L Calio, M Salado, S Kazim et al. A generic route of hydrophobic doping in hole transporting material to increase longevity of perovskite solar cells. Joule, 2, 1800(2018).

    [26] B Xu, J Huang, H Ågren et al. AgTFSI as p-type dopant for efficient and stable solid-state dye-sensitized and perovskite solar cells. ChemSusChem, 7, 3252(2014).

    [27] J Y Seo, H S Kim, S Akin et al. Novel p-dopant toward highly efficient and stable perovskite solar cells. Energ Environ Sci, 11, 2985(2018).

    [28] Y Saygili, H-S Kim, B Yang et al. Revealing the mechanism of doping of spiro-MeOTAD via Zn complexation in the absence of oxygen and light. ACS Energy Lett, 5, 1271(2020).

    [29] W H Nguyen, C D Bailie, E L Unger et al. Enhancing the hole-conductivity of spiro-ometad without oxygen or lithium salts by using spiro(TFSI)2 in perovskite and dye-sensitized solar Cells. J Am Chem Soc, 136, 10996(2014).

    [30] B Tan, S R Raga, A S R Chesman et al. LiTFSI-free spiro-OMeTAD-based perovskite solar cells with power conversion efficiencies exceeding 19%. Adv Energy Mater, 9, 1901519(2019).

    [31] S Wang, Z Huang, X Wang et al. Unveiling the role of tBP–LiTFSI complexes in perovskite solar cells. J Am Chem Soc, 140, 16720(2018).

    [32] F Lamberti, T Gatti, E Cescon et al. Evidence of spiro-ometad de-doping by tert-butylpyridine additive in hole-transporting layers for perovskite solar cells. Chem, 5, 1806(2019).

    [33] J Luo, J Xia, H Yang et al. Toward high-efficiency, hysteresis-less, stable perovskite solar cells: unusual doping of a hole-transporting material using a fluorine-containing hydrophobic lewis acid. Energ Environ Sci, 11, 2035(2018).

    [34] J Xia, Y Zhang, C Xiao et al. Tailoring electric dipole of hole-transporting material p-dopants for perovskite solar cells. Joule, 6, 1689(2022).

    Dongmei He, Shirong Lu, Juan Hou, Cong Chen, Jiangzhao Chen, Liming Ding. Doping organic hole-transport materials for high-performance perovskite solar cells[J]. Journal of Semiconductors, 2023, 44(2): 020202
    Download Citation