You are on page 1of 17

Journal Pre-proofs

Short communication

Dopant-free Methoxy Substituted Copper(II) Phthalocyanine for Highly Effi-


cient and Stable Perovskite Solar Cells

Xingdong Ding, Cheng Chen, Li Tao, Cheng Wu, Mengmeng Zheng, Hongfei
Lu, Hui Xu, Huaming Li, Ming Cheng

PII: S1385-8947(20)30121-2
DOI: https://doi.org/10.1016/j.cej.2020.124130
Reference: CEJ 124130

To appear in: Chemical Engineering Journal

Received Date: 6 November 2019


Revised Date: 25 December 2019
Accepted Date: 13 January 2020

Please cite this article as: X. Ding, C. Chen, L. Tao, C. Wu, M. Zheng, H. Lu, H. Xu, H. Li, M. Cheng, Dopant-free
Methoxy Substituted Copper(II) Phthalocyanine for Highly Efficient and Stable Perovskite Solar Cells, Chemical
Engineering Journal (2020), doi: https://doi.org/10.1016/j.cej.2020.124130

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will
undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing
this version to give early visibility of the article. Please note that, during the production process, errors may be
discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2020 Elsevier B.V. All rights reserved.


Dopant-free Methoxy Substituted Copper(II) Phthalocyanine for

Highly Efficient and Stable Perovskite Solar Cells

Xingdong Ding,a Cheng Chen,a* Li Tao,b Cheng Wu,a Mengmeng Zheng,a Hongfei

Lu,b Hui Xu,a Huaming Li,a Ming Cheng a*

a Institute for Energy Research, Key Laboratory of Zhenjiang, Jiangsu University,

Zhenjiang 212013, Jiansu, China

b School of Environmental and Chemical Engineering, Jiangsu University of Science

and Technology, Zhenjiang 212003, Jiangsu, China

*Corresponding authors.

E-mail addresses: chencheng@ujs.edu.cn (C. Chen), mingcheng@ujs.edu.cn (M.

Cheng)

Keywords: Phthalocyanine; Dopant-free; Hole transport material; Mesoporous;

Perovskite solar cell

Abstract: Many researches manifest that the presences of dopant and additive in hole

transport layer (HTL) accelerate the degradation of perovskite solar cell (PSC).

Therefore, in terms of enhancing stability, it is promising to develop dopant-free hole

transport materials (HTMs). Herein, a methoxy substituted copper(II) phthalocyanine

1
CuPc-(OMe)8 was synthesized and successfully applied as dopant-free HTM in the

PSCs. The optimized PSC devices based on the dopant-free CuPc-(OMe)8 achieve the

highest efficiency of 18.3%, which is competitive with traditional doped Spiro-

OMeTAD. More importantly, an impressively superior long-term stability is

represented by dopant-free CuPc-(OMe)8 based device. These results highlight that

CuPc-(OMe)8 is a promising dopant-free HTM for application in PSCs.

Main text

During the past decade, perovskite solar cell (PSC) emerging as the third-generation

PV technology has attracted much attention from worldwide and went through

unprecedented development due to its advantages of easy fabrication, cost-effective and

excellent photovoltaic performance. [1-6] To date, the certified record power

conversion efficiency (PCE) of PSC has rocketed to 25.2%. [7] In PSCs devices, the

efficient charge carrier extraction and transport is crucial for achieving high efficiency,

and hole transport material (HTM) plays an important role in facilitating charge

separation and transport. Conventionally, the HTM applied in highly efficient PSCs is

2,2’,7,7’-tetrakis(N, N-di-p-methoxyphenylamine)-9,9’-spirobifluorene (Spiro-

OMeTAD) due to its advantages of applicable energy levels and good hole mobility.

[3, 5, 8] However, the complicated synthesis routes and difficult purification process

raises up the production cost of Spiro-OMeTAD. Moreover, because of intrinsic

relatively low conductivity of pristine Spiro-OMeTAD, all the eye-catching PCEs were

achieved by adding dopants and additives, which further increase the total cost of the

2
devices, as well as reduce the stability of devices. [9-11]

On account of this, tremendous efforts have been made to develop dopant-free and

low-cost alternatives with straightforward synthesis routes; and different design

strategies and different types of dopant-free HTMs have been developed. [12-27]

Thereinto, due to the excellent hole transport properties, good thermal and

photochemical stability, organometallic HTMs are considered to be promising

alternatives of Spiro-OMeTAD. Miyasaka and co-workers employed chlorophylls

derivatives Chl-1 and Chl-2 as dopant-free HTMs, and provided PCEs of 11.44% and

8.06%, respectively. [28] Our group firstly reported dopant-free NiPc-(OBu)8 and V2O5

composed organic-inorganic hybrid HTL, and a PCE of 17.6% was achieved. [29] By

using copper phthalocyanine (CuPc) as HTM, Fang and co-workers reported a fully

vacuum-processed PSC. [30] Combined with the cost-effective carbon counter

electrode, the CuPc-based PSC showed an impressive PCE of 16.1% and greatly

enhanced long-term stability. [31] By tuning the peripheral groups of CuPc, Sun group

further improved the PCE to 17.6%. [32] These results showed that the film-forming

properties and distinct charge carrier mobility could be perfected by molecular

engineering.

Herein, a solution-processable copper (II) 1, 4, 8, 11, 15, 18, 22, 25-octamethyl-

29H, 31H-phthalocyanine, termed CuPc-(OMe)8, have been synthesized (see Scheme

S1) and successfully applied as dopant-free HTM in PSCs. The detailed chemical

structure is shown in Figure 1. The pristine CuPc-(OMe)8 provides comparable hole

mobility and conductivity with the doped Spiro-OMeTAD. The PSC employing CuPc-

3
(OMe)8 as dopant-free HTM showed an impressive PCE of 18.3% at 100 mW cm-2

illumination and exhibited excellent long-term stability.

O O

O N N O
N
N Cu N

N N N
O O

O O

Figure 1. Molecular structure of CuPc-(OMe)8

Figure 2. a) UV-vis spectra of CuPc-(OMe)8, b) CV of CuPc-(OMe)8 and Spiro-

OMeTAD. c) Conductivity measurement of CuPc-(OMe)8 and doped Spiro-OMeTAD.

d) J-V characteristics of the CuPc-(OMe)8 and doped Spiro-OMeTAD based hole-only

devices.

4
Table 1. Optical and electrochemical properties of CuPc-(OMe)8 and Spiro-OMeTAD

λmax λmax Hole


E0-0 EHOMO ELUMO Conductiv
HTM solution film mobility/
/eV /eV d /eV ity /S∙cm-1
/nm /nm cm2∙V-1s-1
CuPc-
727 743 1.62 -5.18 -3.56 1.44×10-4 1.10×10-3
(OMe)8
Spiro-
388 - 3.00 -5.16 -2.16 1.58×10-4 2.19×10-3
OMeTAD

The UV-vis absorption spectra of CuPc-(OMe)8 in dichloromethane and in film

state is shown in Figure 2a) and the relevant optical and electrochemical characteristics

are depicted in Table 1. In dichloromethane solution, CuPc-(OMe)8 shows an intense

absorption peak located at 727 nm, which is assigned to Q-band absorption peak

originated from the π-π* transition from the highest occupied molecular orbital (HOMO)

energy level to the lowest unoccupied molecular orbital (LUMO) energy level. In film

state, a 16 nm redshift of maximum absorption peak and a broadening Q-band

absorption peak were obviously detected, indicating the strong intermolecular π-π

stacking, which is favorable for the hole mobility and conductivity. [15, 33] The optical

band gap of CuPc-(OMe)8 was estimated to be 1.62 eV from the absorption edge of

CuPc-(OMe)8 film by using equation E0-0 = 1240/λ. As shown in Figure 2b), the

HOMO energy level of CuPc-(OMe)8 is calculated from the first oxidation potential,

and the value is extracted to be -5.18 eV, which is comparable with Spiro-OMeTAD (-

5.16 eV). The HOMO energy level is positive than the valence band of perovskite,

therefore, hole extraction and transport from the perovskite to CuPc-(OMe)8 is


5
favorable. According to the formula of ELUMO = EHOMO + E0–0, the LUMO energy level

was estimated to be -3.56 eV, which is much higher than the conduction band of the

perovskite, indicating the charge recombination between the perovskite and Au counter

electrode can be efficiently inhibited.

Figure 3. a) Energy level diagram of the components in PSC. b) Cross-section SEM of

PSC, c) top-view SEM of perovskite film and morphology of hole transport layer

prepared from different concentration (d: 10 mg·mL-1, e: 20 mg·mL-1, f: 30 mg·mL-1,

g: 40 mg·mL-1, h: 50 mg·mL-1)

The hole mobility and conductivity were evaluated by using the space-charge-

limited current (SCLC) method and a two-contact electrical conductivity setup,

respectively. Fitting the J-V curves of a hole-only device (see Figure 2d)), the hole

6
mobility of the pristine CuPc-(OMe)8 is 1.10×10-3 cm2∙V-1s-1, which is slightly lower

than the doped Spiro-OMeTAD (2.19×10-3 cm2∙V-1s-1) but in the same magnitude. The

conductivity of the pristine CuPc-(OMe)8 is 1.44 × 10-4 S∙cm-1, which is also

comparable with the doped Spiro-OMeTAD (1.58×10-4 S∙cm-1). The excellent hole

mobility and conductivity properties make CuPc-(OMe)8 promising dopant-free HTM

candidate for application in PSCs.

Figure 4. a) J-V characteristics of PSCs prepared from different CuPc-(OMe)8

concentration, b) PCE changing trend with different CuPc-(OMe)8 concentration.

Table 2. Photovoltaic performance of PSCs containing CuPc-(OMe)8 as HTM

HTM
Jsc (mA
concentration Voc (V) FF (%) PCE (%) Rs / Ω Rsh / Ω
cm–2)
(mg∙ml-1)
10 1.00 21.4 62.0 13.3 81.2 47325.6
20 1.02 21.8 65.4 14.6 74.6 51097.1
30 1.03 22.0 75.6 17.1 65.9 62197.5
40 1.05 22.1 79.0 18.3 62.2 68308.7
50 1.06 21.4 63.1 14.3 79.3 50169.4

Finally, we employed CuPc-(OMe)8 as dopant-free HTM in PSCs with a device

structure of fluorine doped tin oxide (FTO) glass/ SnO2/perovskite/CuPc-(OMe)8/Au,

as shown in Figure 3a-b). The PSC exhibits a good layer-by-layer structure. The

7
optimized thickness of the CuPc-(OMe)8 HTL layer is 100 nm, which is slightly

thinner than that of Spiro-OMeTAD (150 nm). As we all know, the concentration of

HTM plays an important role in the uniformity, thickness and quality of the formed

film. In this regard, the effects of CuPc-(OMe)8 concentration on film morphology and

devices performance were firstly examined. From the SEM images (see Figure 3c-h)

we can find that, with low concentration, the coverage of HTL is poor. When the

concentration increases to 40 mg · mL-1, the perovskite can be nearly fully covered.

Further increasing the concentration induces serious aggregations, making the film look

very rough. The big differences of HTL morphology directly lead to different

photovoltaic performance. As shown in Figure 4a) and Table 2, with low concentration

of 10 mg∙mL-1, the CuPc-(OMe)8-based PSC represented a passable PCE of 13.3%.

Further increasing CuPc-(OMe)8 concentration to 20 mg∙mL-1, 30 mg∙mL-1 and 40

mg∙mL-1 resulted in a continuously improved PCE (see Figure 4a) and 4b)).

Nevertheless, when the CuPc-(OMe)8 concentration went on increasing to 50 mg∙mL-

1, the open-circuit voltage (Voc) was slightly improved due to the efficiently restrict the

charge recombination by thicker HTL layer, but the dramatically increased series

resistance resulted in dropped FF, correspondingly, the PCE was not further improved.

Through optimization, the appropriate HTM concentration is 40 mg∙mL-1, and the

champion efficiency of 18.3% was obtained, with a Voc of 1.05 V, a short-circuit current

density (Jsc) of 22.1 mA·cm-2 and a fill factor (FF) of 79.0%. As discussed above, the

improved efficiency may come from the more uniform and compact CuPc-(OMe)8 film

formed, which restricts the direct contact of perovskite and Au, correspondingly,

8
suppress the internal charge recombination. Therefore, the following mentioned CuPc-

(OMe)8-based PSCs all adopt the concentration of 40 mg∙mL-1.

Figure 5. J-V characteristics of a) CuPc-(OMe)8-based and b) doped Spiro-OMeTAD-

based PSCs (OC: open-circuit, SC: short-circuit), c) IPCE spectra of CuPc-(OMe)8-

based and doped Spiro-OMeTAD-based PSCs, d) Statistic of photovoltaic parameter

PCE of CuPc-(OMe)8-based and doped Spiro-OMeTAD-based PSCs.

Table 3. Photovoltaic performance of CuPc-(OMe)8-based and doped Spiro-

OMeTAD-based PSCs

Scan
HTM Voc (V) Jsc (mA cm–2) FF (%) PCE (%)
directions
pristine forward 1.05 22.1 79.0 18.3
CuPc-(OMe)8 backward 1.05 22.0 76.4 17.6
Doped Spiro- forward 1.10 22.3 76.3 18.7
OMeTAD backward 1.09 22.3 73.7 17.9

9
The current density-voltage (J-V) characteristics for the PSCs based on dopant-

free CuPc-(OMe)8 and doped Spiro-OMeTAD under the AM 1.5 G irradiation at 100

mW cm-2 are depicted in Figure 5a) and 5b), respectively, and the corresponding

parameters are summarized in Table 3. The champion PSC device using CuPc-(OMe)8

as dopant-free HTM yields a promising PCE of 18.3%, comparable with the devices

based on doped Spiro-OMeTAD as reference (Voc = 1.10 V, Jsc = 22.3 mA cm–2, FF =

76.3%, PCE = 18.7%). Almost neglectable hysteresis for device based on the pristine

CuPc-(OMe)8 is observed between the forward and backward scans. The incident

photo-to-electron conversion efficiency (IPCE) spectra of the devices based on the

pristine CuPc-(OMe)8 and the doped Spiro-OMeTAD were shown in Figure 5c).

Compared with the device based on the doped Spiro-OMeTAD as reference, the IPCE

values for the pristine CuPc-(OMe)8 based device shows a slightly decrease in the

region of 400 – 450 nm and 710 – 750 nm due to the competitive absorption of material

CuPc-(OMe)8 with perovskite, resulting in a slightly lower Jsc. Meanwhile, both the

CuPc-(OMe)8 and Spiro-OMeTAD based PSCs showed good reproducibility and the

devices using CuPc-(OMe)8 as dopant-free HTM obtained an average PCE of 17.1%,

completely the same with that of doped Spiro-OMeTAD (see Figure 5d and Figure S2).

As a key photovoltaic performance parameter, the stability of PSC devices based

on the dopant-free CuPc-(OMe)8 has been further evaluated and devices based on the

doped Spiro-OMeTAD was employed as reference, as shown in Figure 6. The devices

were stored in the ambient condition (the humidity of 40-50%) without capsulation.

The reference device based on the doped Spiro-OMeTAD got a slightly enhanced PCE

10
after first 5 days, but sharply dropped to the 12.5% in the next 25 days. The initial

improvement of PCE can be mainly attributed to the passivation effect of TBP in hole

transport layer. [34] Quite differently, the PSC device based on the dopant-free CuPc-

(OMe)8 still maintain 90% initial efficiency after 30 days, exhibiting extremely

excellent stability. The poor stability of doped Spiro-OMeTAD based PSC could be

attributed to the introduction of LiTFSI and TBP. [10, 11] On the other hand, moisture

also play a key role in the degradation of perovskite, so the hydrophobicity properties

of the pristine CuPc-(OMe)8 and doped Spiro-OMeTAD were subsequently

investigated through water contact angel test. From the test results shown in Figure 7,

the water contact angle of the pristine CuPc-(OMe)8 is estimated to be 112o, which is

much bigger than the doped Spiro-OMeTAD (83o), showing a better hydrophobicity.

Thus, the better hydrophobicity of CuPc-(OMe)8, as well as the absence of dopant and

additives, contribute the excellent stability of the PSC based on CuPc-(OMe)8.

11
Figure 6. a) VOC, b) JSC, c) FF and d) PCE variation tendency of pristine CuPc-(OMe)8

and doped Spiro-OMeTAD based PSCs during aging test.

Figure 7. Water contact angle of pristine CuPc-(OMe)8 and doped Spiro-OMeTAD.

In conclusion, a novel organometallic compound CuPc-(OMe)8 has been selected

and successfully applied as dopant-free HTM in the PSCs. This material has advantages

of low synthesis cost and good stability. Most importantly, the pristine CuPc-(OMe)8

exhibits comparable hole mobility and conductivity properties compared with the

doped Spiro-OMeTAD. These all make it be an excellent dopant-free HTM instead of

traditional doped HTM Spiro-OMeTAD. Applied in PSCs, the optimized devices based

on the dopant-free CuPc-(OMe)8 achieve a highest PCE of 18.3% under the AM 1.5 G

irradiation at 100 mW cm-2, while showing a greatly better long-term stability compared

with Spiro-OMeTAD. Thus, this material provides a new choice of promising dopant-

free HTM for application in PSC field.

Acknowledgements

This work was financially supported by the National Natural Science

Foundation of China (Grants 21805114, 21905119), Natural Science

Foundation of Jiangsu province (BK20180867, BK20180869), China

12
Postdoctoral Science Foundation (2019M651741), Six talent peaks project in

Jiangsu province (XNY066), Jiangsu distinguished professor project, the

Jiangsu University Foundation (17JDG032, 17JDG031), High-tech Research

Key laboratory of Zhenjiang (SS2018002).

Reference

[1] A. Kojima, K. Teshima, Y. Shirai, T. Miyasaka, J. Am. Chem. Soc., 131 (2009)

6050-6051.

[2] N.-G. Park, Mater.Today, 18 (2015) 65-72.

[3] N.J. Jeon, J.H. Noh, W.S. Yang, Y.C. Kim, S. Ryu, J. Seo, S.I. Seok, Nature, 517

(2015) 476-480.

[4] N.J. Jeon, H. Na, E.H. Jung, T.-Y. Yang, Y.G. Lee, G. Kim, H.-W. Shin, S. Il Seok,

J. Lee, J. Seo, Nature Energy, 3 (2018) 682-689.

[5] H. Zhou, Q. Chen, G. Li, S. Luo, T.B. Song, H.S. Duan, Z. Hong, J. You, Y. Liu,

Y. Yang, Science, 345 (2014) 542-546.

[6] H.S. Jung, N.G. Park, Small, 11 (2015) 10-25.

[7] https://www.nrel.gov/pv/assets/pdfs/best-research-cell-efficiencies.20190923.pdf.

[8] Q. Jiang, L. Zhang, HaolinWang, X. Yang, J. Meng, H. Liu, Z. Yin, JinliangWu, X.

Zhang, J. You, Nature Energy, 2 (2016) 16177.

[9] L. Calió, S. Kazim, M. Grätzel, S. Ahmad, Angew.Chem. Int. Ed., 55 (2016) 14522-

14545.

[10] E.J. Juarez-Perez, M.R. Leyden, S. Wang, L.K. Ono, Z. Hawash, Y. Qi, Chem.

13
Mater. , 28 (2016) 5702-5709.

[11] S. Wang, Z. Huang, X. Wang, Y. Li, M. Gunther, S. Valenzuela, P. Parikh, A.

Cabreros, W. Xiong, Y.S. Meng, J. Am. Chem. Soc., 140 (2018) 16720-16730.

[12] C. Chen, X. Ding, H. Li, M. Cheng, H. Li, L. Xu, F. Qiao, H. Li, ACS Appl. Mater.

Interfaces, 10 (2018) 36608-36614.

[13] C. Chen, H. Li, X. Ding, M. Cheng, H. Li, L. Xu, F. Qiao, H. Li, L. Sun, ACS

Appl. Mater. Interfaces, 10 (2018) 38970-38977.

[14] M. Cheng, B. Xu, C. Chen, X. Yang, F. Zhang, Q. Tan, Y. Hua, L. Kloo, L. Sun,

Adv. Energy Mater., 5 (2015) 1401720.

[15] H.-C. Liao, T.L.D. Tam, P. Guo, Y. Wu, E.F. Manley, W. Huang, N. Zhou, C.M.M.

Soe, B. Wang, M.R. Wasielewski, L.X. Chen, M.G. Kanatzidis, A. Facchetti, R.P.H.

Chang, T.J. Marks, Adv. Energy Mater., 6 (2016) 1600502.

[16] F. Zhang, C. Yi, P. Wei, X. Bi, J. Luo, G. Jacopin, S. Wang, X. Li, Y. Xiao, S.M.

Zakeeruddin, M. Grätzel, Adv. Energy Mater., 6 (2016) 1600401.

[17] C. Shen, Y. Wu, H. Zhang, E. Li, W. Zhang, X. Xu, W. Wu, H. Tian, W.H. Zhu,

Angew. Chem. Int. Ed., 58 (2019) 3784-3789.

[18] J. Wang, H. Zhang, B. Wu, Z. Wang, Z. Sun, S. Xue, Y. Wu, A. Hagfeldt, M.

Liang, Angew. Chem. Int. Ed., 58 (2019) 15721-15725.

[19] Q. Xiao, J. Tian, Q. Xue, J. Wang, B. Xiong, M. Han, Z. Li, Z. Zhu, H.L. Yip, Z.

Li, Angew. Chem. Int. Ed., (2019).

[20] J. Zhou, X. Yin, Z. Dong, A. Ali, Z. Song, N. Shrestha, S.S. Bista, Q. Bao, R.J.

Ellingson, Y. Yan, W. Tang, Angew. Chem. Int. Ed., 58 (2019) 13717-13721.

14
[21] K. Rakstys, C. Igci, M.K. Nazeeruddin, Chem. Sci., 10 (2019) 6748-6769.

[22] G.-W. Kim, G. Kang, J. Kim, G.-Y. Lee, H.I. Kim, L. Pyeon, J. Lee, T. Park,

Energy Environ. Sci., 9 (2016) 2326-2333.

[23] C. Huang, W. Fu, C.Z. Li, Z. Zhang, W. Qiu, M. Shi, P. Heremans, A.K. Jen, H.

Chen, J. Am. Chem. Soc., 138 (2016) 2528-2531.

[24] Y. Liu, Q. Chen, H.-S. Duan, H. Zhou, Y. Yang, H. Chen, S. Luo, T.-B. Song, L.

Dou, Z. Hong, Y. Yang, J. Mater. Chem. A, 3 (2015) 11940-11947.

[25] H.D. Pham, S.M. Jain, M. Li, S. Manzhos, K. Feron, S. Pitchaimuthu, Z. Liu, N.

Motta, H. Wang, J.R. Durrant, P. Sonar, J. Mater. Chem. A, 7 (2019) 5315-5323.

[26] C. Chen, M. Cheng, P. Liu, J. Gao, L. Kloo, L. Sun, Nano Energy, 23 (2016) 40-

49.

[27] M. Cheng, K. Aitola, C. Chen, F. Zhang, P. Liu, K. Sveinbjörnsson, Y. Hua, L.

Kloo, G. Boschloo, L. Sun, Nano Energy, 30 (2016) 387-397.

[28] M. Li, Y. Li, S.I. Sasaki, J. Song, C. Wang, H. Tamiaki, W. Tian, G. Chen, T.

Miyasaka, X.F. Wang, ChemSusChem, 9 (2016) 2862-2869.

[29] M. Cheng, Y. Li, M. Safdari, C. Chen, P. Liu, L. Kloo, L. Sun, Adv. Energy Mater.,

7 (2017) 1602556.

[30] W. Ke, D. Zhao, C.R. Grice, A.J. Cimaroli, G. Fang, Y. Yan, Journal of Materials

Chemistry A, 3 (2015) 23888-23894.

[31] F. Zhang, X. Yang, M. Cheng, W. Wang, L. Sun, Nano Energy, 20 (2016) 108-

116.

[32] X. Jiang, D. Wang, Z. Yu, W. Ma, H.-B. Li, X. Yang, F. Liu, A. Hagfeldt, L. Sun,

15
Adv. Energy Mater., 9 (2019) 1803287.

[33] Y. Liu, Z. Hong, Q. Chen, H. Chen, W.H. Chang, Y.M. Yang, T.B. Song, Y. Yang,

Adv. Mater., 28 (2016) 440-446.

[34] C. Wu, C. Chen, L. Tao, X. Ding, M. Zheng, H. Li, G. Li, H. Lu, M. Cheng, J.

Energy Chem., 43 (2020) 98-103.

Highlights

1. A novel solution processable copper(II) phthalocyanine was selected and successfully applied

as dopant-free HTM in the PSCs.

2. The dopant-free CuPc-(OMe)8 delivered comparable PCE with Spiro-OMeTAD.

3. The dopant-free CuPc-(OMe)8 based perovskite solar cell shows enhanced stability.

16

You might also like