Facile Approach to Preparing a Vanadium Oxide ... - ACS Publications

May 4, 2017 - Key Laboratory of Advanced Carbon Materials and Wearable Energy Technologies of Jiangsu Province, Soochow Institute for Energy and Mater...
0 downloads 0 Views 1MB Size
Subscriber access provided by HKU Libraries

Article

A Facile Approach to Prepare a Vanadium Oxide Hydrate Layer as a Hole Transport Layer for High-Performance Polymer Solar Cell Hailin Cong, Dongwei Han, Bingbing Sun, Dongying Zhou, Chen Wang, Ping Liu, and Lai Feng ACS Appl. Mater. Interfaces, Just Accepted Manuscript • Publication Date (Web): 04 May 2017 Downloaded from http://pubs.acs.org on May 5, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Applied Materials & Interfaces is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

A Facile Approach to Prepare a Vanadium Oxide Hydrate Layer as a Hole Transport Layer for High-Performance Polymer Solar Cell Hailin Cong,a,b,‡ Dongwei Han,a,b,‡ Bingbing Sun,b Dongying Zhou,b Chen Wang,b Ping Liu,b Lai Fengb,* a

Laboratory for New Fiber Materials and Modern Textile, Growing Base for State

Key Laboratory, College of Chemical Engineering, Qingdao University, Qingdao, 266071, China b

Key Laboratory of Advanced Carbon Materials and Wearable Energy Technologies

of Jiangsu Province, Soochow Institute for Energy and Materials Innovations, College of Physics, Optoelectronics and Energy, Soochow University, Suzhou 215006, China ‡

These authors contribute equally to this work.

*Corresponding

author

(Tel:

+86

512

65223650;

E-mail

address:

[email protected] (Lai Feng)).

1

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 28

ABSTRACT We demonstrate a facile and green approach to prepare a vanadium oxide hydrate (VOx·nH2O) layer to serve as the hole transport layer (HTL) in high-performance polymer solar cells (PSCs). The VOx·nH2O layer was in-situ prepared by a combined H2O2 and UVO processing on a VOx layer. The as-prepared VOx·nH2O layer featured a work function of 5.0±0.1 eV, high transmittance and better interface properties compared to the generally prepared VOx (UVO or thermal annealing) layers. PSCs based on PTB7-th:PC71BM using the VOx·nH2O layer as HTL yielded high power conversion efficiencies (PCEs) up to 8.11%, outperforming the devices with the VOx layers (PCE of 6.79% for UVO processed VOx layer and 6.10% for thermal annealed VOx layer) and conventional PEDOT-PSS layer (PCE of 7.67%). The improved PCE was attributed to the enhanced JSC and/or FF, which mainly correlates to the improved interfacial contact between the photoactive layer and ITO/HTL or cathode when using the VOx·nH2O layer as HTL. Similar improvement in PCE was also observed for the PSCs based on P3HT:PC61BM. In addition, PSCs with a VOx·nH2O layer as HTL showed higher stability than those with a PEDOT:PSS layer. Hence, it would be possible to use this simply and in-situ prepared VOx·nH2O layer as an inexpensive HTL for high-performance PSCs.

KEYWORDS: polymer solar cells, hole transport layer, vanadium oxide hydrate, H2O2-UVO processing, interfacial contact

2

ACS Paragon Plus Environment

Page 3 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

1. INTRODUCTION Polymer solar cells (PSCs) have received recent attentions owing to their advantages

of

lightweight,

flexibility,

low-cost

and

environment-friendly

characteristics.1-4 Through long-term efforts, the efficiency of single-junction PSCs has been steadily increased up to >10%,5-8 which makes them competitive alternatives or supplementary to the silicon or inorganics-based photovoltaics. Nevertheless, in order to realize the practical use of PSCs in near future, their stability issue and applicability for large-scale production (for example, a roll-to-roll process) still need to be significantly improved.9-12 The interface layers (i.e., hole transport layer and electron transport layer) between the electrode and the photoactive layer are important components of PSC, which are of remarkable benefit to the charge separation and charge collection.13,14 Particularly, polyethylenedioxythiophene:polystyrenesulfonate (PEDOT:PSS) has been

widely

used

for

preparing

hole

transport

layer

(HTL).

However,

PEDOT:PSS-based HTL is still far from being ideal choice due to its high-cost as well as unfavored acidic and hygroscopic natures,15 which not only increase the fabrication cost but also lead to reduced stability of PSCs. Alternatively, a variety of semiconducting transition metal oxides such as NiOx,16-18 MoOx,19-22 VOx,23-27 CuOx28,29 and WO3,30,31 ReOx,32 have been developed as HTL materials for fabricating high-stability and cost-effective PSCs. Among them, VOx is particularly promising because of its low-cost, high ambient stability, nice optical and electrical properties. In early works, V2O5-based HTL was prepared via vacuum thermal evaporation.33-35 3

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 28

However, such a technology requires expensive facilities, which shows low compatibility with the easy processing of PSCs. Recent works focus on low-temperature solution-processed V2O5 or VOx. In a typical study, Hou J. H. et al. reported using hydrogen vanadium bronzes (HxV2O5), prepared by the reaction of vanadium metal powder and H2O2 in presence of ethanol,23 to fabricate HTL for achieving high-performance PSCs. Though device tests revealed that the performance of HxV2O5-based PSCs was still inferior to that of the standard device with a PEDOT:PSS layer, the effectiveness of vanadium oxides as HTL was clearly evidenced. In a more advanced study, Yang R. Q. et al. demonstrated another V2O5-based HTL processed by O2 plasma, which even outperformed the conventional PEDOT:PSS layer due to the excellent surface property.25 Nevertheless, such a processing is not suitable for large-scale fabrication of PSCs owing to the fact that the generation of O2 plasma usually requires an expensive equipment. Moreover, in recent studies, vanadium pentaoxide hydrate (V2O5·nH2O) sol made by aging metavanadic acid solution or through a melting-quenching sol−gel method was applied as HTL in PSCs.26,27 Though the V2O5·nH2O-based PSCs exhibited better performance than that of the standard device with a PEDOT:PSS layer, their cost effectiveness is relatively lower due to the long time aging of metavanadic acid (ca. 20 days) or high-temperature melting of V2O5 (ca. 800

o

C). A facile approach for high

cost-effectiveness thus remains to be achieved. According to previous works, vanadia gels can be alternatively prepared by a simple and green reaction of V2O5 and H2O2.36-40 Taking this strategy into account, in 4

ACS Paragon Plus Environment

Page 5 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

this work, we report a facile way to prepare a vanadium oxide hydrate (VOx·nH2O) based HTL for PSCs, in which a combined H2O2 and UVO processing was introduced after short-time thermal annealing of vanadyl acetylacetonate (VO(acac)2) layer. The in-situ prepared VOx·nH2O layer exhibited improved interfacial properties as compared to the UVO-only processed or thermally annealed VOx layer. Hence, when the

VOx·nH2O

layer

was

used

as

HTL

in

the

PSCs

based

on

poly[(ethylhexyl-thiophenyl)-benzodithiophene-(ethylhexyl)-thienothiophene]:[6,6]phenyl-C71-butyric

acid

methyl

ester

(PTB7-th:PC71BM)

or

poly(3-hexylthiophene):[6,6]-phenyl-C61-butyric acid methyl ester (P3HT:PC61BM), improved PCE with enhanced FF and Jsc was observed. In addition, when comparing with the standard devices, in which PEDOT:PSS layer was used as HTL, the devices with VOx·nH2O layer showed comparable or even improved PCEs with higher stability. 2. EXPERIMENTAL 2.1 Chemicals Vanadyl acetylacetonate (VO(acac)2) (97%) was purchased from Alfa-Aesar. H2O2 (30%) and isopropyl alcohol were purchased from Sinopharm Jiangsu Co. Ltd. PTB7-th, P3HT and fullerene acceptors (PC61BM and PC71BM) were provided by Nano-C Inc.. PEDOT:PSS (Al 4083, Clevios) was purchased from H. C. Stark company. 2.2 VOx·nH2O thin film preparation 5 mg VO(acac)2 was dissolved into 5 mL isopropyl alcohol and the solution was 5

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 28

ultra-sonicated for 3-4 h. After that, the VOx·nH2O layer was prepared as follows. Firstly, the VO(acac)2 solution was filtered with a filter (0.45 µm) and spin-casted on an ITO substrate at 2000 rpm for 30 s, followed by a short-time thermal annealing (150 oC for 2-3 minutes in air). Then, 30% H2O2 (30-40 µL) was spin-casted on the as-prepared VOx layer at 2000 rpm for 30 s and the layer was subjected to an ultraviolet-ozone (UVO) treatment (Ultraviolet Ozone Generator, T1OX10/OES, UVOCS Inc., USA) for 10 min. The as-prepared layer is denoted as VOx·nH2O (H2O2-UVO) layer. In control experiments, a VOx (UVO-only) layer or a VOx (annealing-only) layer was prepared by spin-casting VO(acac)2 solution on an ITO substrate, followed by a UVO treatment (for 10 min) or a thermal annealing process (150 oC in air for 15 min). 2.3 Thin film characterizations X-ray photoelectron spectra (XPS) were recorded with an ESCALAB 250 spectrometer with a Al Kα excitation source. The work function of the VOx·nH2O layer was estimated by ultraviolet photoelectron spectroscopy (UPS) on an AXIS Ultra-DLD using HeI irradiation with hv=21.22 eV. Fourier transform infrared (FTIR) spectra were recorded with a Nicolet 510 spectrometer. Raman spectra were obtained on an instrument (Renishaw inVia-reflex) using a 532 nm excitation laser. Transmittance spectra were obtained on a Shimazu UV2600 spectrometer. Atomic force microscopy (AFM) measurements were conducted on a MFP-3D-BIO (Asylum Research) system in tapping mode. Raman spectra was obtained on an instrument (Horiba Jobin Yvon) using a 532 nm laser as excitation source. The contact angle was 6

ACS Paragon Plus Environment

Page 7 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

measured on a system of MY-SPCX1-01 (Yuming Tech., China). 2.4 Device fabrication and characterizations The PSCs were fabricated with a standard configuration (ITO/HTL/active layer/cathode). The ITO glass was cleaned following a previously reported method.41 Next, VOx·nH2O and various VOx layers were prepared on the ITO substrate as mentioned above. Alternatively, PEDOT:PSS layer with a thickness of 40-60 nm was prepared.41 Then, a photoactive layer was prepared following a previously reported method.41 Finally, the cathode (Ca/Al: 20/80 nm) was vacuum deposited (ca. 10-5 Pa) on the photoactive layer with a shadow mask, giving an active area of 4 mm2. The current density–voltage (J–V) characteristics were measured under a standard condition (AM. 1.5 G) with a Keithley 2400 SourceMeter. The incident photon-to-current conversion efficiency (IPCE) spectra were recorded on a QE-R3011 system (Enli Tech., China). The

hole-only

device

with

a

configuration

of

ITO/HTL/

PTB7-th:PC71BM/MoO3/Ag was fabricated for space-charge-limited-current (SCLC) measurement. The fabrication details were provided in Supporting Information (SI). The hole mobility was calculated from the J-V curve obtained in dark (see SI). 3. RESULTS AND DISCUSSION 3.1. Characterizations of the VOx·nH2O layer. XPS studies were carried out to investigate the chemical compositions of the in-situ prepared VOx·nH2O (H2O2-UVO) layer. Fig. 1 presents the O 1s and V 2p high-resolution spectra of the VOx·nH2O (H2O2-UVO) layer, in comparison with 7

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 28

those of UVO-only processed VOx layer. Particularly, the O1s spectrum of the VOx (UVO-only) layer can be fitted into two peaks at 530.4 and 531.6 eV, corresponding to the oxygen species in metal oxide (VOx) and undecomposed precursor (C=O), respectively.42-45 This indicates that most vanadium precursors have been converted to vanadium oxides through a UVO process, similar to those obtained from a thermal annealing process.20 As for the VOx·nH2O (H2O2-UVO) layer, a broader O 1s spectrum was observed, in which additional fitting peak at 533.1 eV can be identified, pointing to the O species in H2O or OH group.46,47 This result indicates that coordinated H2O molecules or OH groups have been introduced during the H2O2-UVO processing due to the reaction between H2O2 and thermally generated VOx.23 On the other hand, similar V 2p spectra were obtained for the VOx·nH2O (H2O2-UVO) and VOx (UVO-only) layers: Either V 2p3/2 or V 2p1/2 profile can be deconvoluted into two peaks (517.7 and 516.8 eV for V 2p3/2; 525.0 and 523.6 eV for V 2p1/2), corresponding to V5+ and V4+ oxidation states, respectively.21,41,42 The presence of V4+ might be resulted from the partial reduction of V5+ due to the UVO or H2O2-UVO treatment. Particularly, in the VOx·nH2O (H2O2-UVO) layer, the ratio of V5+ to V4+ is 2.3:1, while the VOx (UVO-only) layer gives a ratio of 3:1. The slightly higher V4+ content in the VOx·nH2O (H2O2-UVO) layer might indicate more oxygen vacancies formed during the H2O2-UVO process and hence better electric conductivity of the layer.

8

ACS Paragon Plus Environment

Page 9 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Fig. 1 (a) XPS spectra of V 2p and O 1s of (a) the VOx·nH2O (H2O2-UVO) layer, in comparison with (b) the UVO-only processed VOx layer.

Fig. 2 FT-IR spectra of the VOx·nH2O (H2O2-UVO) and VOx (UVO-only) layers.

Fig. 2 presents the FTIR spectrum of the as-prepared VOx·nH2O (H2O2-UVO), in comparison with the UVO-only processed VOx. In both spectra, it is clearly seen the characteristic peaks of the stretching vibration of V=O (997 cm-1) and H–O (around 3500 cm-1). Nevertheless, in the spectrum of VOx·nH2O (H2O2-UVO), the 9

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 28

characteristic peak of H-O is much more intensive than that observed for the UVO-only processed VOx, indicating additional OH groups or water molecules were introduced via the H2O2 processing. Moreover, the micro-Raman spectra of the as-prepared VOx·nH2O (H2O2-UVO) and VOx (UVO-only) layers are shown in Fig. S2. Both spectra showed bands at 989, 685, 519, 477, 423, 280 and 138 cm-1, corresponding to V2O5.40,48 Besides, additional bands at 872 and 158 cm-1 were observed for the VOx·nH2O (H2O2-UVO) layer, indicating the presence of coordinated H2O molecules or OH groups. Thus, this is fully consistent with the XPS result, confirming the formation of vanadium oxide hydrate (VOx·nH2O) through the H2O2-UVO processing. UPS was employed to estimate the work function (WF) of the as-prepared VOx·nH2O (H2O2-UVO) layer. Fig. 3 provides the secondary electron cutoff spectra. The WF of the VOx·nH2O (H2O2-UVO) layer spin-coated on an ITO substrate is calculated to be 5.0±0.1 eV, slightly lower than that of the PEDOT:PSS layer (5.1±0.1 eV) measured under the same conditions or reported in the literatures.49 In addition, the as-prepared VOx·nH2O (H2O2-UVO) layer is highly transparent in the range of 500-800 nm (see transmittance spectra in Fig. S3), similar to that of the VOx (UVO-only) layer as well as the PEDOT:PSS layer. Such an optical feature allows the use of this VOx·nH2O (H2O2-UVO) layer as HTL for PSCs.

10

ACS Paragon Plus Environment

Page 11 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Fig. 3 UPS spectra of PEDOT:PSS and VOx·nH2O (H2O2-UVO) layers.

The surface morphologies of the VOx·nH2O (H2O2-UVO) and VOx (UVO-only) layer were characterized by tapping-mode atomic force microscopy (AFM). As shown in Fig. 4a-c, the VOx·nH2O (H2O2-UVO) layer demonstrates a uniform surface with a root-mean-square (rms) roughness of 2.53 nm. It is lower than that of the VOx (UVO-only) layer (5.13 nm) though higher than that of the PEDOT:PSS layer (1.22 nm), suggesting that additional H2O2 treatment yielded the in-situ formation of VOx·nH2O layer with reduced surface roughness. In addition, the topography of PTB7-th:PC71BM photoactive layers on top of different HTL layers are present in Fig. 4d-f. The rms roughness of the blend film deposited on the VOx·nH2O (H2O2-UVO) layer is 0.56 nm, remarkably lower than that of the blend film on VOx (UVO-only) layer (1.14 nm) and PEDOT:PSS layer (1.07 nm). These results suggest that a smoother photoactive film can be obtained on the top of VOx·nH2O (H2O2-UVO) layer. Hence, the interfacial contact between the active layer and the top electrode can be improved, which benefits charge collection. 11

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 28

Fig. 4 AFM images of HTL layers: (a) VOx·nH2O (H2O2-UVO) layer, (b) VOx (UVO-only) layer, (c) PEDOT:PSS layer, (The AFM image of a bare ITO is present in Fig. S4 as reference.) and (d,e,f) AFM topography images of active layers (AL: PTB7-th:PC71BM) on top of different HTL layers. The scan size is 10 µm × 10 µm. To better understand the influence of H2O2-UVO treatment to the VOx·nH2O interface, we carried out contact angle measurements and the results are shown in Fig. 5. Particularly, the water contact angle of the VOx·nH2O (H2O2-UVO) layer is 11.1°, lower than those observed for the VOx (UVO-only) layer (14.4°) and the VOx (annealing-only) layer (61.2°) as well as the PEDOT:PSS layer (19.7o or 22.17o).25 The low water contact angle of the VOx·nH2O (H2O2-UVO) layer might indicate the presence of many dangling bonds on the surface of the layer, which lead to substantial surface reconstruction and benefit interfacial contact.25 Moreover, as seen in Fig. S5, the contact angle of 1,2-dichlorobenzene on the VOx·nH2O (H2O2-UVO) layer (7.8°) 12

ACS Paragon Plus Environment

Page 13 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

is also slightly lower than those for the VOx (UVO-only) layer (10.4°) and the PEDOT:PSS layer (10.8o), suggesting a slightly better wettability for the VOx·nH2O (H2O2-UVO) layer. Thus, the combined feature of smooth surface and good wettability can account for the formation of a more uniform active layer on the VOx·nH2O (H2O2-UVO) layer as discussed above.

Fig. 5 Measured water contact angle for different HTL. VOx·nH2O (H2O2-UVO), VOx (UVO-only), PEDOT:PSS and VOx (annealing-only) layers from top to bottom, respectively.

3.2 Photovoltaic Performance of the device with a VOx·nH2O layer. To investigate the performance of the VOx·nH2O (H2O2-UVO) layer as HTL, it was applied to fabricate the PTB7-th:PC71BM based PSC with a configuration of ITO/HTL/active layer/Ca/Al. For comparison, a series of PSCs with a PEDOT:PSS 13

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 28

layer or differently processed VOx layer were prepared, optimized and tested. Fig. 6a illustrates the J−V curves of various PSCs under optimized conditions (see Table S1 and Fig. S6 in SI). The averaged device performance and the best PCE values are listed in Table 1. Obviously, the best photovoltaic performance was obtained from the device B with a VOx·nH2O (H2O2-UVO) layer. The average PCE reaches to 7.97% with an open-circuit voltage (VOC) of 0.78 V, a fill factor (FF) of 64.62% and a short-circuit current density (JSC) of 15.76 mA/cm2. However, only moderate PCEs were achieved in the devices C (6.58%) and D (6.04%) with differently processed VOx layers (i.e., UVO or annealing-only) due to their lower VOC, FF and JSC. Additionally, it is found that device B is even superior to the standard device A with a PEDOT:PSS layer, which gave a lower average PCE of 7.52% with a VOC of 0.790 V, a JSC of 14.39 mA/cm2, and an FF of 66.22%, all comparable to those reported in recent literatures.50,51 These results suggest that the VOx·nH2O layer prepared through a simple H2O2-UVO processing might possess a better hole-transport or hole-collection property compared to the regularly prepared VOx layers as well as the conventional PEDOT:PSS layer. Furthermore, to ascertain the accuracy of the J-V results, IPCE spectra were measured for the devices A, B and C. As shown in Fig. 6b, similar IPCE spectra were observed for these devices. Particularly, device B exhibited slightly higher IPCE value in the range of 400-750 nm relative to those of the devices A and C, indicating a better charge collection efficiency of the VOx·nH2O (H2O2-UVO) layer in this range. The calculated Jsc is 14.68 mA/cm2 for device A, 15.80 mA/cm2 for device B and 14.60 14

ACS Paragon Plus Environment

Page 15 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

mA/cm2 for device C, all of which are close to those obtained from the J-V tests, respectively, with an error within 5%. To better understand the slightly higher Jsc of device B relative to those of devices A and C, their J-V characteristics were measured in dark. As shown in Fig. 6c, the current density at reversed biases from device B is smaller than those of devices A and C, indicating a reduced leakage current in device B.52,53 This result is in good agreement with its larger shunt resistance Rsh (938 Ω cm2) relative to device A (532 Ω cm2) and smaller series resistance Rs (6.2 Ω cm2) relative to device C (7.7 Ω cm2, see Table 1). Thus, the higher Jsc of device B might be understood by the enhanced Rsh or reduced Rs,54 which mainly correlates to the better interfacial contact between ITO/HTL and active layer when using a VOx·nH2O (H2O2-UVO) layer as the HTL.

Fig. 6 (a) J−V curves of the best performing devices with different HTLs under 1000 W/m2 AM 15

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 28

1.5G illumination. (b) IPCE curves along with integrated current densities. (c) J−V curves in dark. The device structure is ITO/HTL/PTB7-th:PC71BM/Ca/Al. Table 1. Device performance of the PSCs based on PTB7-th:PC71BM with different HTLs.a

device

A

VOC

JSC

FF

(V)

(mA/cm2)

(%)

0.790

14.39

66.22

0.784

15.76

hole transport layer

PEDOT:PSS

Rs(Ω

Rsh(Ω

cm2)c

cm2)d

7.52(7.67)

6.0

532

64.62

7.97(8.11)

6.2

938

PCE (%)b

VOx·nH2O B (H2O2-UVO) C

VOx (UVO-only)

0.762

14.60

59.20

6.58(6.79)

7.7

901

D

VOx (annealing-only)

0.768

14.14

55.70

6.04(6.10)

11.8

507

a

Statistical data achieved from 10 independent devices. bThe maxima PCEs are in the brackets. c,d

Rs and Rsh were defined using the method reported in literatures.55-57

To explore the hole transport characteristics of the VOx·nH2O (H2O2-UVO) layer, the SCLC measurements were performed and the Mott−Gurney equation58-60 was employed for calculating the hole mobility (µh): JSCLC=9ε0ετµh(V-VBI)2/(8L3). The definitions for ε0, ετ, V, VBI, L and the calculation details for µh are described in SI. The dark J-V curves of hole-only devices are present in Fig. 7. The µh of the device with the

(H2O2-UVO) layer was determined to be 5.59×10-4 cm2 V-1 s-1, higher than that

of the devices with the VOx (UVO-only) layer (3.66×10-4 cm2 V-1 s-1). These results indicate that the hole injected from the active layer, could be more easily extracted for the device with a VOx·nH2O (H2O2-UVO) layer, in agreement with the improved FF as suggested by the J-V tests. The enhanced µh of the hole-only device with a VOx·nH2O (H2O2-UVO) layer further illustrates the positive effect of the improved interface properties for hole extraction. 16

ACS Paragon Plus Environment

Page 17 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Fig. 7 J-V characteristics of the hole-only devices with different HTLs. The fitting results are given as solid lines.

To check the viability of the VOx·nH2O (H2O2-UVO)-based HTL, the PSCs based on P3HT:PC61BM was fabricated. As P3HT is a representative wide bandgap polymer donor, it is very different from the above used low bandgap polymer of PTB7-th. Nevertheless, similar results were obtained (see Table S2 and Fig. S7): the device B′ with a VOx·nH2O (H2O2-UVO) layer demonstrated a superior photovoltaic performance to those of the devices C′, D′ with differently processed VOx layers (UVO or annealed layer). The PCE of device B′ is even comparable to that of the standard device A′ with a PEDOT:PSS layer. The long-term stabilities of devices with PEDOT:PSS and VOx·nH2O (H2O2-UVO) layers, respectively, stored under N2 atmosphere are exhibited in Fig. 8. The device with a VOx·nH2O (H2O2-UVO) layer maintained nearly 85% of its initial PCE after 20 days. Meanwhile, the device with a PEDOT:PSS layer experienced a faster decay with the PCE decreasing to ca. 60% probably owing to the presence of a corrosive and hygroscopic PEDOT:PSS layer. These results suggest that the VOx·nH2O-based HTL features higher stability as compared to the PEDOT:PSS layer, 17

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 28

which makes it more viable HTL for PSCs.

Fig. 8 Degradation of encapsulated devices at room temperature under an N2 atmosphere.

4. CONCLUSIONS In summary, we demonstrated a facile in-situ preparation of VOx·nH2O layer through a combined H2O2-UVO processing on VOx layer, which can be used as HTL for high-performance PSCs. The composition of the as-prepared VOx·nH2O (H2O2-UVO) layer was confirmed by using the XPS, FT-IR and Raman spectroscopies. Further characterizations revealed that the VOx·nH2O (H2O2-UVO) layer features a proper work function of ca. 5.0±0.1 eV, a high transmittance, a good wettability and a smooth surface with many dangling bonds. All these features may induce energy alignment for ITO/HTL/active layer and facilitate the formation of a uniform photoactive layer as well as good interfacial contacts in PSCs. As a result, the PSCs based on PTB7-th:PC71BM with a VOx·nH2O (H2O2-UVO) layer yielded an average PCE of 7.97% with high stability. The PCE is improved by ~32% compared 18

ACS Paragon Plus Environment

Page 19 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

to the devices with the VOx layers and superior to the device with a conventional PEDOT:PSS layer. In addition, the VOx·nH2O (H2O2-UVO) layer was also applicable for P3HT:PC61BM based PSC, giving an average PCE of 3.24%. Thus, the VOx·nH2O (H2O2-UVO) layer reported in this work can be considered as a promising alternative to the conventional PEDOT:PSS layer for high-performance and cost-effective PSCs. ASSOCIATED CONTENT Supporting Information

Schematic molecule structures; Raman spectra; optical transmittance spectra; AFM image, contact angle images; device parameters and J-V curves of the PSCs based on PTB7-th:PC71BM and P3HT:PC61BM, respectively; details for SCLC measurements. This material is available free of charge via the Internet at http://pubs.acs.org. ACKNOWLEDGEMENTS We cordially thank Prof. Zexin Zhang, Mr. Qiang Tao (Soochow University) for the kind help in the contact angle measurements. The work is supported in part by the NNSF of China (No. 51372158), the NSF of Jiangsu Province (No. BK20160325) and Suzhou key laboratory for advanced carbon materials and wearable energy technologies. In addition, this work is also in part supported by Collaborative Innovation Center (CIC) of Suzhou Nano Science and Technology.

19

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 28

REFERENCES 1. Günes, S.; Neugebauer, H.; Sariciftci, N. S., Conjugated Polymer-Based Organic Solar Cells. Chem. Rev. 2007, 107, 1324-1338. 2. Thompso, B.C.; Frechet, M.J., Polymer-Fullerene Composite Solar Cells. Angew. Chem. Int. Ed. 2008, 47, 58-77. 3. Dennler, G.; Scharber, M.C.; Brabec, C.J., Polymer-Fullerene Bulk-Heterojunction Solar Cells. Adv. Mater. 2009, 21, 1323-1338. 4. Lu, L.; Zheng, T.; Wu, Q.; Schneider, A.M.; Zhao, D.; Yu, L., Recent Advances in Bulk Heterojunction Polymer Solar Cells. Chem. Rev. 2015, 115, 12666-12731. 5. Chen, J.D.; Cui, C.; Li, Y.Q.; Zhou, L.; Ou, Q.D.; Li, C.; Li, Y.F.; Tang, J.X., Single-Junction Polymer Solar Cells Exceeding 10% Power Conversion Efficiency. Adv. Mater. 2015, 27, 1035-1041. 6. Yu, W.; Huang, L.; Yang, D.; Fu, P.; Zhou, L.; Zhang, J.; Li, C., Efficiency Exceeding 10% for Inverted Polymer Solar Cells with a ZnO/Ionic Liquid Combined Cathode Interfacial Layer. J. Mater. Chem. A 2015, 3 (20), 10660−10665. 7. Zhao, W.; Qian, D.; Zhang, S.; Li, S.; Inganas, O.; Gao, F.; Hou, J., Fullerene-Free Polymer Solar Cells with over 11% Efficiency and Excellent Thermal Stability. Adv. Mater. 2016, 28, 4734−4739. 8. Liu, X.; Li, X.; Li, Y.; Song, C.; Zhu, L.; Zhang, W.; Wang, H. Q.; Fang, J., High-Performance Polymer Solar Cells with PCE of 10.42% via Al-Doped ZnO Cathode Interlayer. Adv. Mater. 2016, 28, 7405-7412. 9. Lee, J.U.; Jung, J.W.; Jo, J.W.; Jo, W.H., Degradation and Stability of Polymer-Based Solar Cells. J. Mater. Chem. 2012, 22, 24265-24283. 20

ACS Paragon Plus Environment

Page 21 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

10. Jørgensen, M.; Norrman, K.; Gevorgyan, S.; Tromholt, T.; Andreasen, B.; Krebs, F., Stability of Polymer Solar Cells. Adv. Mater. 2012, 24, 580-612. 11. Krebs, F. C.; Tromholt, T.; Jorgensen, M., Upscaling of Polymer Solar Cell Fabrication Using Full Roll-to-roll Processing. Nanoscale 2010, 2, 873-886. 12. Søndergaard, R.; Hösel, M.; Angmo, D.; Larsen-Olsen, T. T.; Krebs, F. C., Roll-to-roll Fabrication of Polymer Solar Cells. Mater. Today 2012, 15, 36-49. 13. Chueh, C.-C.; Lia, C.-Z.; Jen, A. K.-Y., Recent Progress and Perspective in Solution Processed Interfacial Materials for Efficient and Stable Polymer and Organometal Perovskite Solar Cells. Energy Environ. Sci. 2015, 8, 1160-1189. 14. Yin, Z.; Wei, J.; Zheng, Q., Interfacial Materials for Organic Solar Cells: Recent Advances and Perspectives. Adv. Sci. 2016, 3, 1500362. 15. Sun, W. H.; Peng, H. T.; Li, Y. L.; Yan, W. B.; Liu, Z. W.; Bian, Z. Q.; Huang, C. H., Solution-Processed Copper Iodide as an Inexpensive and Effective Anode Buffer Layer for Polymer Solar Cells. J. Phys. Chem. C 2014, 118, 16806-16812. 16. Chen, W.; Wu, Y. Z.; Liu, J.; Qin, C. J.; Yang, X. D.; Islam, A.; Cheng, Y. B.; Han, L. Y., Hybrid Interfacial Layer Leads to Solid Performance Improvement of Inverted Perovskite Solar Cells. Energy Environ. Sci. 2015, 8, 629-640. 17. Kim, J. H.; Liang, P. W.; Williams, S. T.; Cho, N.; Chueh, C. C.; Glaz, M. S.; Ginger, D. S.; Jen, A. K. Y., High-Performance and Environmentally Stable Planar Heterojunction Perovskite Solar Cells Based on a Solution-Processed Copper-Doped Nickel Oxide Hole-Transporting Layer. Adv. Mater. 2015, 27, 695-701. 18. Park, J. H.; Seo, J.; Park, S.; Shin, S. S.; Kim, Y. C.; Jeon, N. J.; Shin, H.-W.; Ahn, T. K.; 21

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 28

Noh, J. H.; Yoon, S. C.; Hwang, C. S.; Seok, S. I., Efficient CH3NH3PbI3 Perovskite Solar Cells Employing Nanostructured p-Type NiO Electrode Formed by a Pulsed Laser Deposition. Adv. Mater. 2015, 27, 4013-4019. 19. Tan, Z.; Zhang, W.; Qian, D.; Cui, C.; Xu, Q.; Li, L.; Li, S. and Li, Y., Solution-Processed Nickel Acetate as Hole Collection Layer for Polymer Solar Cells. Phys. Chem. Chem. Phys. 2012, 14, 14217–14223. 20. Sun, Y.-M.; Takacs, C. J.; Cowan, S. R.; Seo, J. H.; Gong, X.; Roy, A.; Heeger, A. J., Efficient, Air-stable Bulk Heterojunction Polymer Solar Cells Using MoOx as the Anode Interfacial Layer. Adv. Mater. 2011, 23, 2226-2230. 21. Tan, Z.; Zhang, W.; Cui, C.; Ding, Y.; Qian, D.; Xu, Qi; Li, L.; Li, S.; Li, Y., Solution-processed Vanadium Oxide as a Hole Collection Layer on an ITO Electrode for High-performance Polymer Solar Cells. Phys. Chem. Chem. Phys. 2012, 14, 14589–14595 22. Tan, Z.; Qian, D.; Zhang, W.; Li, L.; Ding, Y.; Xu, Q.; Wang, F. and Li, Y., Efficient and Stable Polymer Solar Cells with Solution Processed Molybdenum Oxide Interfacial Layer. J. Mater. Chem. A 2013, 1, 657-664. 23. Xie, F.; Choy, W. C.; Wang, C.; Li, X.; Zhang, S.; Hou, J., Low Temperature Solution-Processed Hydrogen Molybdenum and Vanadium Bronzes for an Efficient Hole-transport Layer in Organic Electronics. Adv. Mater. 2013, 25, 2051−2055. 24. Kim, B. S.; Kim, T. M.; Choi, M. S.; Shim, H. S.; Kim, J. J., Fully Vacuum–Processed Perovskite Solar Cells With High Open Circuit Voltage Using MoO3/NPB as Hole Extraction Layers. Org. Electron. 2015, 17, 102-106. 25. Bao, X.; Zhu, Q.; Wang, T.; Guo, J.; Yang, C.; Yu, D.; Wang, N.; W. Chen,; Yang, R., 22

ACS Paragon Plus Environment

Page 23 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Simple O2 Plasma-Processed V2O5 as an Anode Buffer Layer for High Performance Polymer Solar Cells. ACS Appl. Mater. Interfaces 2015, 7, 7613−7618. 26. Teran-Escobar, G.; Pampel, J.; Caicedo, J. M.; Lira-Cantu, M., Low-Temperature, Solution-Processed, Layered V2O5 Hydrate as the Hole-Transport Layer for Stable Organic Solar Cells. Energy Environ. Sci. 2013, 6, 3088−3098. 27. Jiang, Y.; Xiao, S.; Xu, B.; Zhan, C.; Mai, L.; Lu, X.; You, W., Enhancement of Photovoltaic Performance by Utilizing Readily Accessible Hole Transporting Layer of Vanadium(V) Oxide Hydrate in a Polymer−Fullerene Blend Solar Cell. ACS Appl. Mater. Interfaces 2016, 8, 11658−11666. 28. Xu, Q.; Wang, F.-Z.; Tan, Z.-A.; Li, L.-J.; Li, S.-S.; Hou, X.-L.; Sun, G.; Tu, X.-H.; Hou, J.-H.; Li, Y.-F., High-Performance Polymer Solar Cells with Solution-Processed and Environmentally Friendly CuOx Anode Buffer Layer. ACS. Appl. Mater. Interfaces 2013, 5, 10658-10664 29. Lien, H. T.; Wong, D. P.; Tsao, N. H.; Huang, C. I.; Su, C. C.; Chen, K. H.; Chen, L. C., Effect of Copper Oxide Oxidation State on the Polymer Based Solar Cell Buffer Layer. ACS Appl. Mater. Interfaces 2014, 6, 22445-22450. 30. Qiu, M.; Zhu, D.; Bao, X.; Wang, J.; Wang, X.; Yang, R., WO3 with Surface Oxygen Vacancies as an Anode Buffer Layer for High Performance Polymer Solar Cells. J. Mater. Chem. A 2016, 4, 894-900. 31. Tan, Z.; Li, L.; Cui, C.; Ding, Y.; Xu, Q.; Li, S. Qian, D. and Li, Y., Solution-Processed Tungsten Oxide as an Effective Anode Buffer Layer for High-Performance Polymer Solar Cells. J. Phys. Chem. C 2012, 116, 18626−18632. 23

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 28

32. Tan, Z. A.; Li, L.; Wang, F.; Xu, Q.; Li, S.; Sun, G.; Tu, X.; Hou, X.; Hou, J.; Li, Y., Solution-Processed Rhenium Oxide: A Versatile Anode Buffer Layer for High Performance Polymer Solar Cells with Enhanced Light Harvest. Adv. Energy Mater. 2014, 4, 1300884. 33. Shrotriya, V.; Li, G.; Yao, Y.; Chu, C.; Yang, Y., Transition Metal Oxides as the Buffer Layer for Polymer Photovoltaic Cells. Appl. Phys. Lett. 2006, 88, 073508. 34. Takanezawa, K.; Tajima, K.; Hashimoto, K., Efficiency Enhancement of Polymer Photovoltaic Devices Hybridized with ZnO Nanorod Arrays by the Introduction of a Vanadium Oxide Buffer Layer. Appl. Phys. Lett. 2008, 93, 063308. 35. Shen, L.; Xu, Y.; Meng, F.; Li, F.; Ruan, S.; Chen, W., Semi Transparent Polymer Solar Cells using V2O5/Ag/V2O5 as Transparent Anodes. Org. Electron. 2011, 12, 1223−1226. 36. Fontenot, C. J.; Wiench, J. W.; Pruski, M.; Schrader, G. L., Vanadia Gel Synthesis via Peroxovanadate Precursors. 1. In Situ Laser Raman and

51

V NMR Characterization of the

Gelation Process. J. Phys. Chem. B 2000, 104, 11622-11631. 37. Wang, Y.; Cao, G., Synthesis and Enhanced Intercalation Properties of Nanostructrued Vanadium Oxides. Chem. Mater. 2006, 18, 2787-2804. 38. Wang, Y.; Shang, H.; Chou, T.; Cao, G., Effects of Thermal Annealing on the Li+ Intercalation Properties of V2O5·nH2O Xerogel Films. J. Phys. Chem. B 2005, 109, 11361-11366. 39. Mansour, A. N.; Smith, P. H.; Balasubramanian, M.; Mcbreen, J., Dependence of Electrochemical Properties of Vanadium Oxide Films on Their Nano- and Microstructures. J. Phys. Chem. B 2005, 109, 16700-16704. 40. Jr, W. A.; Ribeiro, C.; Leite, E. R.; Mastelaro, V. R., Vanadium Pentoxide Nanostructures: An 24

ACS Paragon Plus Environment

Page 25 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Effective Control of Morphology and Crystal Structure in Hydrothermal Conditions. Cryst. Growth Des. 2009, 9, 3626-3631. 41. Gu, Z.; Zhou, D.; Sun, B.; Tang, M.; Chen, K.; Feng, L.; Zhou, Y., Potassium-Neutralized Perylene Derivative (K4PTC) and rGO-K4PTC Composite as Effective and Inexpensive Electron Transport Layers for Polymer Solar Cells. Org. Electron. 2016, 37, 47-54. 42. Briggs, D.; Seah, M. P., Practical Surface Analysis—Auger and X-Ray Photoelectron Spectroscopy, John Wiley & Sons, New York, 2nd edn, 1990, 53, 215 43. Suchorski, Y.; Rihko-Struckmann, L.; Klose, F.; Ye, Y.; Alandjiyska, M.; Sundmacher, K.; Weiss, H., Evolution of Oxidation States in Vanadium-Based Catalysts under Conventional XPS Conditions. Appl. Surf. Sci. 2005, 249, 231. 44. Beamson, G.; Briggs, D. The Scienta ESCA 300 Database, Wiley, Chichester, 1992. 45. Crane, J.; Warwick, M.; Smith, R.; Furlan, N.; Binions, R., The Application of Electric Fields to Aerosol Assisted Chemical Vapor Deposition Reactions. J. Electrochem. Soc. 2011, 158, D62-D67. 46. van Spronsen, M. A.; Weststrate, K-J.; Dunnen, A. d.; van Reijzen, M. E.; Hahn, C.; Juurlink, L. B. F., Hydrophilic Interaction Between Low-Coordinated Au and Water: H2O/Au(310) Studied with TPD and XPS. J. Phys. Chem. C 2016, 120, 8693−8703. 47. Kim, Y. J.; Park, C. R., Analysis of Problematic Complexing Behavior of Ferric Chloride with N,N-Dimethylformamide Using Combined Techniques of FT-IR, XPS, and TGA/DTG. Inorg. Chem. 2002, 41, 6211-6216. 48. Fontenot, C. J.; Wiench, J. W.; Pruski, M.; Schrader, G. L., Vanadia Gel Synthesis via Peroxovanadate Precursors. 2. Characterization of the Gels. J. Phys. Chem. B 2001, 105, 25

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 28

10496-10504. 49. Huang, L.Q.; Cheng, X.F.; Yang, J.; Zhang, L.F.; Zhou, W.H.; Xiao, S.Q.; Tan, L.C.; Chen, L.; Chen, Y.W., High-Performance Polymer Solar Cells Realized by Regulating the Surface Properties of PEDOT:PSS Interlayer from Ionic Liquids. ACS Appl. Mater. Interfaces 2016, 8, 27018−27025. 50. Lin, Y.Z.; Wang, J.Y.; Zhang, Z-G; Bai, H.T.; Li, Y.F.; Zhu, D.B.; Zhan, X.W., An Electron Acceptor Challenging Fullerenes for Efficient Polymer Solar Cells. Adv. Mater. 2015, 27, 1170–1174. 51. Zhao, L.; Zhao, S. L.; Xu, Z.; Huang, D.; Zhao, J.; Li, Y.; Xu, X. R., The Effects of Improved Photoelectric Properties of PEDOT:PSS by Two-step Treatments on the Performance of Polymer Solar Cells Based on PTB7-Th:PC71BM. ACS Appl. Mater. Interfaces 2016, 8, 547-552. 52. Destouesse, E.; Chambon, S.; Courtel, S.; Hirsch, L.; Wantz, G.

Solution-Processed

Small-Molecule Bulk Heterojunctions: Leakage Currents and the Dewetting Issue for Inverted Solar Cells. ACS Appl. Mater. Interfaces 2015, 7, 24663−24669. 53. Proctor, C. M.; Nguyen, T.-Q. Effect of Leakage Current and Shunt Resistance on the Light Intensity Dependence of Organic Solar Cells. Appl. Phys. Lett. 2015, 106, 083301. 54. Nian, L.; Zhang, W. Q.; Wu, S. P.; Qin, L. Q.; Liu, L. L.; Xie, Z. Q.; Wu, H. B.; Ma, Y. G., Perylene Bisimide as a Promising Zinc Oxide Surface Modifier: Enhanced Interfacial Combination for Highly Efficient Inverted Polymer Solar Cells. ACS Appl. Mater. Interfaces 2015, 7, 25821−25827. 55. Zhang, Z.G.; Qi, B.Y.; Jin, Z.W.; Chi, D.; Qi, Z.; Li, Y.F.; Wang, J.Z., Perylene Diimides: a 26

ACS Paragon Plus Environment

Page 27 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Thickness-Insensitive Cathode Interlayer for High Performance Polymer Solar Cells. Energy Environ. Sci. 2014, 7, 1966-1973. 56. Liu, X.D.; Jiao, W.-X.; Lei, M.; Zhou, Y.; Song, B.; Li, Y.F., Crown-ether Functionalized Fullerene as a Solution Processable Cathode Buffer Layer For High Performance Perovskite and Polymer Solar Cells. J. Mater. Chem. A 2015, 3, 9278-9284. 57. Kakavelakis, G.; Konios, D.; Stratakis, E.; Kymakis, E., Enhancement of the Efficiency and Stability of Organic Photovoltaic Devices via the Addition of a Lithium-Neutralized Graphene Oxide Electron-Transporting Layer. Chem. Mater. 2014, 26, 5988-5993. 58. Goh, C.; Kline, R. J.; McGehee, M. D.; Kadnikova, E. N.; Frechet, J. M. J., Molecular Weight-Dependent Mobilities in Regioregular Poly(3-hexyl-thiophene) diodes. Appl. Phys. Lett. 2005, 86, 122110. 59. Cheng, Y.J.; Hsieh, C.H.; Li, P.J.; Hsu, C.S., Morphological Stabilization by in Situ Polymerization of Fullerene Derivatives Leading to Efficient, Thermally Stable Organic Photovoltaics. Adv. Funct. Mater. 2011, 21, 1723-1732. 60. Liu, Y.; Wan, X.; Wang, F.; Zhou, J.; Long, G.; Tian, J.; Chen, Y., High-performance Solar Cells Using a Solution-Processed Small Molecule Containing Benzodithiophene Unit. Adv. Mater. 2011, 23, 5387-5391.

27

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 28

TOC

28

ACS Paragon Plus Environment