You are on page 1of 11

Journal of Cluster Science

https://doi.org/10.1007/s10876-022-02368-6

ORIGINAL PAPER

Solution Processed ­WO3 and PEDOT:PSS Composite for Hole Transport


Layer in ITO‑Free Organic Solar Cells
P. Gurudevi1 · P. Venkateswari2 · T. Sivakumar2 · C. Ramesh3 · P. Vanitha4

Received: 27 June 2022 / Accepted: 10 October 2022


© The Author(s), under exclusive licence to Springer Science+Business Media, LLC, part of Springer Nature 2022

Abstract
We developed PEDOT:PSS [poly (styrenesulfonate)-doped poly(3,4-ethylenedioxythiophene] and tungsten oxide ­(WO3)
based hybrid thin films by facile hydrothermal method. The solution processed W ­ O3–PEDOT:PSS hybrid thin films are fully
characterized by X-ray diffraction (XRD), scanning electron microscope (SEM), UV–Vis (UV), Raman, X-ray photoelectron
spectra (XPS) and Brunauer–Emmett–Teller (BET) surface area analysis. XRD and SEM results suggest that ­WO3 has crys-
talline structure with monoclinic phase and spherical shaped nanoparticles (20–25 nm), which is uniformly decorated on the
surface of PEDOT:PSS sheets. The value of E ­ g [calculated by linear extrapolating of (F(R ∞) hν)2 and hν] was evaluated to
be 3.11 and 2.85 eV for pure ­WO3 and ­WO3/PEDOT:PSS, respectively. The PL intensity of ­WO3/PEDOT:PSS is maximum
lower than the pristine ­WO3 and PEDOT:PSS, implying an improved interfacial contact quality at the ­WO3/PEDOT:PSS.
Photovoltaic performance of ­WO3–PEDOT:PSS devices were measured under AM 1.5 G 1 sun light intensity of 100 mW/
cm2. Organic solar cell devices with W ­ O3–PEDOT:PSS layer give rise to a significant increase in J­ SC from 9.35 ± 0.03 to
16.2 ± 0.01 mA/cm2 respectively and in ­VOC from 0.28 V to 0.72 V respectively, resulting in a remarkable PCE increase from
0.592 ± 0.03 to 0.714 ± 0.02%. The ­WO3/PEDOT:PSS composite deliver a PCE of 7.81 ± 0.01%. The superior performance
of ­WO3/PEDOT:PSS composite was attributed to the reduced current leakage, enhanced hole extraction characteristics, and
less trap-assisted interfacial recombination via current density–voltage.

Keywords WO3–PEDOT:PSS · Composite thin films · Spin coating · Solar light · Organic solar cells · Photo conversion
efficiency

Introduction well as the simplicity of the printing-based processes used


to manufacture the cells [1–3]. In academic and industrial
Organic solar cells (OSCs) based on polymer–fullerene com- research, power conversion efficiencies (PCEs) of 3–5%
posite active layers have been considered as next generation have been easily obtained from bulk heterojunction OSCs
photovoltaic devices following Si-based photovoltaics due with poly 3-hexylthiophene (P3HT) and [6, 6]-phenyl ­C61
to their low fabrication cost, light weight, and flexibility, as butyric acid methyl ester (PCBM) active layers due to the
simplicity of both their structure and the fabrication pro-
cess. In particular, a recent important report on OSCs with
* P. Gurudevi a high PCE of 9.2% suggests that cost-efficient and high-
gurudevi.mphil@gmail.com
performance OSCs may be mass-produced in the near future
1
Department of Electronic Science, A.V.C. College [4]. Wide bandgap polymer donor PM6 and narrow band-
(Autonomous), Mannampandal, Mayiladuthurai, gap polymer acceptor PY-IT were selected to construct all-
Tamil Nadu 609305, India polymer solar cells (all-PSCs) with a layer-by-layer (LbL)
2
Department of Electronics, Rathnavel Subramaniam College or bulk heterojunction (BHJ) structure. The solvent additive
of Arts and Science (Autonomous), Sulur, Coimbatore, 1-chloronaphthalene (CN) plays a vital role in improving
Tamil Nadu 641402, India
the performance of all-PSCs. The PCE of LbL all-PSCs is
3
Department of Mechanical Engineering, M. Kumarasamy improved to 15.81% from 13.67% by incorporating 1 vol%
College of Engineering, Karur, Tamil Nadu 639006, India
CN in PY-IT solution, benefiting from the simultaneously
4
Department of Management Studies, M. Kumarasamy enhanced short circuit current density (­ JSC) of 22.61 mA/cm2
College of Engineering, Karur, Tamil Nadu 639006, India

13
Vol.:(0123456789)
P. Gurudevi et al.

and fill factor (FF) of 73.62% [5]. At present, the majority OSCs. Choi et al. reported on the merits of an W ­ O3 HIL,
of research into OSCs has been devoted to the development specifically citing its high optical transmittance over a wide
of new organic active materials or device structures (e.g., solar spectral range and proper energy level alignment for
tandem or inverted OSCs) to improve the performance of hole-extraction from P3HT [29]. Tao et al. also suggested
photovoltaics [6]. ­WO3 with a high work function of 4.80 eV to efficiently
Although the choice of organic active materials and extract holes and suppress electrons from the active layer
device structure is critical to the performance of an OSC, [30]. Although W-doped indium oxide has been suggested as
the importance of a stable transparent anode cannot be an alternative to ITO anode for OSCs, a new electrode con-
ignored because exciton formation and carrier extraction is cept combining a buffer layer and anode for PEDOT:PSS-
significantly affected by the transmittance and resistivity of free OSCs has not yet been reported [31]. In this study, we
the transparent anode. One of the major challenges to be design and fabricate OSCs of ­WO3–PEDOT:PSS hybrid
overcome for the commercialization of OSs is the develop- thin films with high PCE and stability for the first time. The
ment of transparent electrodes that are inexpensive, solution- PEDOT:PSS acts as an conductive binder to connect the
processable, and suitable for OPV devices [7–10]. Indium ­WO3 nanoparticles leading to reduced impedance and bet-
tin oxide (ITO) is widely used as a transparent conductive ter electrochemical properties. The improved photovoltaic
electrode because of its high transparency (~ 90%) and low mechanism of the proposed materials was also discussed.
sheet resistance (~ 10 Ω/sq). However, the application of
ITO in flexible electronics is challenging owing to its poor
mechanical properties and high-temperature manufacturing Experimental Section
processes, as well as the limited supply sources of indium.
Although various materials, such as carbon nanotubes, gra- Materials
phene, metallic nanowires, and conductive polymers, have
been proposed to replace ITO, many of them are expensive The conducting polymer (PEDOT:PSS) was used for the
within the framework of commercialization [11]. Transpar- synthesis of hybrid material and purchased from Agfa-
ent electrodes for flexible electronics require a high electri- Gevart, Group Belgium. W ­ O3 prepared by hydrothermal
cal conductivity, a high transmittance for visible light, and method, the main precursor i.e. sodium tungstate dehydrate
simple processability [12–19]. Conductive polymers, such ­(Na2WO4·2H2O ≥ 99.09%) was used in synthesis process
poly(3,4-ethylenedioxythiophene):poly(styrene sulfonate) that was purchased from Sigma Aldrich. The chemical struc-
(PEDOT:PSS) complexes, have great potential for large- ture and energy level of PEDOT:PSS was displayed in Fig.
area deposition through solution-based processes with an S1a, b.
aqueous dispersion as well as cost reduction via mass pro-
duction [20–22]. PEDOT:PSS has been widely used as the Fabrication Process of ­WO3 and ­WO3/PEDOT:PSS
flexible transparent electrode because of the advantage of Thin Films
high mechanical flexibility, high conductivity, high trans-
parency in the visible range and excellent thermal stability. Solution A was made by dissolving ammonium molybdate
PEDOT:PSS films are typically fabricated by spin coating ­Na2WO4·2H2O in water to make a 0.01 mol/L solution. Solu-
onto substrates. The film coats on the entire substrate. When tion B was a 2 mol/L hydrochloric acid (HCl) water solution.
used as the bottom electrode, patterning of the PEDOT:PSS The pH of the solution was measured by using digital pH
is needed to avoid or reduce the shunt current to improve the meter. The pH of the mixed solution was changed between 1
device performance. PEDOT:PSS has exhibited good stabil- and 1.5 by dropping solution B into solution A. After 30 min
ity under ambient conditions and a direct-current (DC) con- in an ultrasonic bath, glass substrates were cleaned with
ductivity over a wide range from 0.01 S/cm to the order of distilled (DI) water, acetone, and alcohol. The final solvent
­103 S/cm [23]. However, it is known that the acidic nature of solution was then added to the treated and dried glass sub-
the PEDOT:PSS layer leads to severe etching of the ITO and strate using spin deposition techniques at 3000 rpm for 60 s
the diffusion of indium into the organic active layer, which and dried for 30 min on a hot plate at 100 °C. In the prepa-
in turn causes degradation of the OSC [24]. To solve the ration of W­ O3/PEDOT:PSS composite films, PEDOT:PSS
problems associated with acidic PEDOT:PSS, metal oxides aqueous solution was mixed with ­WO3 aqueous solution
such as NiO, ­V2O5, ­MoO3, and ­WO3 have been introduced as with the ratio of 1:1. The final aqueous solution was applied
a hole injection layer (HIL) between the organic active layer to the washed and dried glass substrate using a spin coat-
and ITO anode in order to improve the reliability of OSCs ing process at 3000 rpm for 60 s and dried for 30 min on
[25–28]. These oxide HILs could assist in hole extraction at a hot plate at 100 °C. A pure PEDOT:PSS film was also
the ITO anode and improve OSC performance. In particular, made without the addition of ­WO3 aqueous solution as a
­WO3 has been reported as an effective and stable HIL for comparative. In the preparation process of pristine ­WO3,

13
Solution Processed ­WO3 and PEDOT:PSS Composite for Hole Transport Layer in ITO‑Free Organic…

ammonium molybdate (1.2 g) was diluted with 50 mL of DI Results and Discussion


water. After that HCl solution was added drop by drop under
strong magnetic stirring until the pH value reaches to 8. The Structural Analysis
final solvent solution was then added to the treated and dried
glass substrate using spin coating techniques at 3000 rpm for To confirm the formation of W ­ O3/PEDOT:PSS compos-
60 s and dried for 30 min on a hot plate at 100 °C. ites, XRD patterns was conducted. Figure 1 shows the
XRD pattern of PEDOT:PSS, W ­ O3 and W­ O3/PEDOT:PSS
composite samples. In XRD spectra of bare PEDOT:PSS,
Characterization Techniques there was no diffraction peak observed for commercially
available PEDOT:PSS conducting polymer due to amor-
To identify the crystal structure, X-ray diffraction (XRD) phous nature [32]. The XRD pattern of pristine ­WO3 thin
studies were carried out using PANalytical X-ray diffrac- film sample, there are series of significant peaks were
tometer with nickel-filtered Cu Ka (30 kV, 30 mA). The observed at 2θ = 23.45°, 23.88°, 24.64°, 26.85°, 29.28°,
surface morphology of the samples was studied using scan- 33.60°, 33.89°, 34.46°, 42.23°, 47.52°, 55.11°, 60.19°
ning electron microscopy (SEM; S-4100, Hitachi). The and 75.16° ICDD-20-1324 [33, 34]. Since there were no
functional group present in the nanocomposite was con- peaks of conducting polymer PEDOT: PSS in the XRD
firmed by Raman spectroscopy RA400, Mettler Toledo. The pattern of ­WO3/PEDOT:PSS composite, which signify the
X-ray photoelectron spectroscopy spectra (XPS) of ­WO3/ amorphous nature of conducting polymer. After the incor-
PEDOT:PSS nanocomposite were recorded via Thermo poration of ­WO3 in to PEDOT:PSS, the peak intensity is
ESCALAB 250 at room temperature. Photoadsorption spec- increased indicating that increase in crystalline nature of
tra of catalysts were tested by UV–Vis diffuse reflectance composite materials. The average grain size was calculated
spectroscopy (DRS Hitachi UV-3600). Photoluminescence from Debye–Scherrer’s formula.
spectra (PL) of CdS-DETA-based samples were tested with
FLS920 fluorescence lifetime and steady state spectrom- K𝜆
d= ,
eter. Specific surface area values ­(SBET) were tested with 𝛽 cos 𝜃
ASAP2020 instrument. where d is the mean crystallite size, K is the shape factor
taken as 0.89, λ is the wavelength of the incident beam, β
is the full width at half maximum and θ is the Bragg angle.
Construction of Organic Solar Cell Setup The average crystalline size of was found as 21.8  nm,
34.1 nm, and 41.6 nm for pure ­WO3, PEDOT:PSS, and ­WO3/
A thin ­WO3–PEDOT:PSS composite film was deposited by PEDOT:PSS composite films, respectively.
using spin coated technique (3000 rpm for 60 s). The pre-
pared glass was sintered at 100 °C for 30 min while sheltered
with a glass Petri dish. Spin-coating P3HT [poly(3-hex-
ylthiophene)] and PCBM (1-(3-methoxycarbonyl)-propyl-
1-phenyl-[6,6]C61) (Sigma Aldrich) in o-dichlorobenzene
(ODCB) at 600 rpm for 60 min resulted in the active layer.
Thermal evaporation was used to deposit a LiF (0.3 nm)
and aluminium (150 nm) cathode through a shadow mask
at a base pressure of 2 × ­10−7 Torr created a device with a 6
­mm2 active area. The gadget was covered with a glass Petri
dish and thermally annealed at 150 °C for 15 min. Five cells
were prepared for each material. Photocurrent–photovoltage
(J–V) curves were recorded under white light irradiation
of 100 mW/cm2 (AM 1.5 G) by a xenon arc lamp (CHF-
XM500, Trusttech Co., Ltd., China) in the air and at room
temperature. All the tests were measured five times in the
same environment and the average data were taken. The
long-term stability of the performance of the DSSCs was
measured with an AM 1.5 G solar simulator (Bunkoh-keiki
Co., Ltd., OTENTO-SUN5) with and without a UV cut-off
filter (< 420 nm). The test was carried out at RT with relative Fig. 1  XRD pattern of the PEDOT:PSS, ­WO3, and ­WO3/PEDOT:PSS
humidity of 71%. composite thin film samples

13
P. Gurudevi et al.

Morphological Studies NPs are adsorbed homogeneously (Fig. 2e). In addition,


more ­WO3 NPs are observed in the AFM height images,
The surface morphology of the PEDOT:PSS, W ­ O3 and implying that the W ­ O3 NPs adsorbed on top layer may be
hybrid film are also investigated by SEM and AFM analy- interdigitated with the ones adsorbed in the underneath lay-
sis. The PEDOT:PSS film reveals a smooth and featureless ers owing to the much smaller thickness of the PEDOT:PSS
morphology as shown in Fig. 2a. On the other hand, the than the size of the ­WO3 NPs. Similar trend is also observed
­WO3 film shows a coarse and rough surface owing to the for the hybrid films (Fig. 2f) because the thickness of a
stacked nanoparticles (Fig. 2b). The stacked nanoparticles PEDOT:PSS layer is also much smaller than the size of W ­ O3
form much aggregation with grain size within 30–40 nm NPs. The room mean square roughness of the PEDOT:PSS,
well-distributed on the surface. In the hybrid film, it can be ­WO3 and hybrid films obtained from the AFM analysis is
seen that the W ­ O3 nanoparticles are well-dispersed in the 0.824, 5.64 and 2.23 nm, respectively. Therefore, both the
PEDOT:PSS matrix. AFM can provide detailed informa- SEM and AFM images of the hybrid film show a smoother
tion about surface morphology, topography and homogene- morphology compared with that of ­WO3 film. In addition,
ity of the films. The AFM micrographs in Fig. 2d–f show the the PEDOT:PSS acts as a conductive binder to connect the
surface features of the PEDOT:PSS, ­WO3-NPs and hybrid ­WO3 nanoparticle which cab reduce the electrical resistance
films, respectively. From Fig. 2d, it is obvious that in the of the resultant film. The chemical compositions were fur-
pristine PEDOT:PSS, the thickness and roughness of the ther characterized by EDX, which is suitable for element
PEDOT:PSS films gradually increase. By contrast, the ­WO3 analysis of thin films. As shown in Fig. 2g, the films consist

Fig. 2  SEM images of a PEDOT:PSS, b ­WO3, c ­WO3/PEDOT:PSS, AFM 2D images of d PEDOT:PSS, e ­WO3, f ­WO3/PEDOT:PSS, and g EDS
image of ­WO3/PEDOT:PSS

13
Solution Processed ­WO3 and PEDOT:PSS Composite for Hole Transport Layer in ITO‑Free Organic…

of W, O and S elements. This suggest that PEDOT:PSS was and peaks centered at frequency 285 ­cm−1 and 333 ­cm−1 rep-
successfully incorporated in the ­WO3 framework. resents the binding modes linkages of oxygen molecules in
­WO3. The Raman impressions of PEDOT:PSS are described
Raman Spectra Analysis in the following references [36, 37]. The different modes
of vibration for PSS are centered at 995 ­cm−1, 1137 ­cm−1,
The better understanding of structural analyses and configu- 1571 ­cm−1. The peaks at frequency 1272 ­cm−1, 1374 ­cm−1,
ration of pure PEDOT:PSS polymer and ­WO3/PEDOT:PSS 1447  ­cm−1 and 1544 ­cm−1 are identified as the inter-ring
hybrid composite films were studied at excitation wave- stretching C–C, stretching of C–C, symmetrical C=C, and
length 532 nm as presented in Fig. 3. The band assigned at antisymmetrical C=C vibration modes, respectively. The
frequency 285 ­cm−1, 333 ­cm−1, 735 ­cm−1 and 821 ­cm−1 are obtained Raman results confirm that prepared hybrid com-
identified as the pure monoclinic phase of crystalline nature posite films properties do not depend on the mixing of indi-
of ­WO3 nanomaterial [35]. The peaks at 735  ­cm−1 and vidual ­WO3 material but also on the process of bond sharing
821 ­cm−1 are attributed to the stretching modes (W–O–W) between ­WO3 nanoparticles and PEDOT:PSS conducting
polymer.

Optical Characteristic Studies

The UV–Vis absorption spectra of pure conducting poly-


mer and hybrid composites films exhibit four absorption
bands  (Fig.  4). The conducting polymer PEDOT: PSS
spectra exhibit absorption bands at 473  nm, while the
­WO3/PEDOT:PSS hybrid composite films the absorption
bands are found to be at position 525 nm. The peaks cen-
tered nearby absorption wavelength 473 nm were due to the
π–π*, π–polaron and polaron–π* transition, respectively. The
localized polaron band between 400 and 600 nm confirmed
that bonds chain of PEDOT:PSS conducting polymer were
compressed. Absorption spectra also indicate the shift in
absorption peaks towards longer wavelength region depict-
ing that the energy band gap of nanocomposites decreased.
Fig. 3  Raman spectra of (a) PEDOT:PSS, (b) W
­ O3, and (c) W
­ O 3/ From these spectra optical band-gap energy (­ Eg) was calcu-
PEDOT:PSS composite thin film samples lated using Kubelka–Munk model, which is described by

Fig. 4  a UV–Vis absorption spectra and b band gap plot of PEDOT:PSS, W


­ O3, and ­WO3/PEDOT:PSS thin films

13
P. Gurudevi et al.

F(R∞) = K/S = (1 − ­R∞)2/2R∞, where ­R∞ = ­Rsample/Rstandard measured for analyzing the charge-transfer processes at the
is the reflectance of an infinitely thick specimen, while K ­WO3/PEDOT:PSS interfaces. As shown in Fig. 5, the inten-
and S are the absorption and scattering coefficients, respec- sity of the PL spectra obtained for the ­WO3/PEDOT:PSS is
tively. F(R∞). hν)2 = A (hν − ­Eg). The variation of F(R∞). maximum lower than the pristine ­WO3 and PEDOT:PSS,
hν)2 versus hν was plotted and the straight line range of these implying an improved interfacial contact quality at the ­WO3/
plots is extended on the x-axis (hν) to obtain the values of PEDOT:PSS. The improved interfacial contact quality is
optical band gap (­ Eg). The value of E­ g (calculated by lin- favourable for an enhanced charge transfer to take place at
ear extrapolating of (F(R ∞) hν)2 and hν) was evaluated to the ­WO3 and PEDOT:PSS interface.
be 2.95 eV, pure conducting polymer. Whereas, the band
gap energy of pure ­WO3 and ­WO3/PEDOT:PSS were 3.11 Surface Area and Elemental Composition Analysis
and 2.85 eV, respectively. The PL behaviors of the pure
PEDOT:PSS, ­WO3 and ­WO3/PEDOT:PSS composites were Brunauer–Emmett–Teller (BET) characterization dem-
onstrated the mesoporous nature of W ­ O 3 and W
­ O 3/
PEDOT:PSS (Fig. 6). The nitrogen adsorption/desorption
isotherms were similar in shape. However, ­WO3 and ­WO3/
PEDOT:PSS showed a more distinct hysteresis loop in the
P/P0 range of 0.5 to 1.0 than that of W ­ O3 (Fig. 6a). This
finding can be attributed to the porosity of PEDOT:PSS and
the presence of ultrafine ­WO3 nanoparticles deposited on
PEDOT:PSS nanosheets. The mesoporous feature can be
further studied using the pore size distribution curves. The
pore size range of ­WO3 and ­WO3/PEDOT:PSS was 4 nm
to 12 nm (Fig. 6b). The surface area of W ­ O3/PEDOT:PSS
was calculated to be 135.8 ­m2/g, which was higher than
­ O3 (60.9 ­m2/g). This mesoporous structure with
that of W
high surface area of composites can play an important role
in providing rapid electrolyte transport, shorter diffusion
paths, and more active sites for photochemical reactions
on the electrode surface. XPS is a surface technique that is
able to provide information on the chemical and electronic
Fig. 5  Photoluminescence spectra of PEDOT:PSS, W
­ O3, and W
­ O 3/ structure of molecules. The survey spectrum of the ­WO3/
PEDOT:PSS composite thin film samples PEDOT–PSS consists of W 4f, O 1s, C 1s, and S 2p, which

­ O3, and ­WO3/PEDOT:PSS and b pore size curve


Fig. 6  a ­N2 adsorption and desorption analysis of W

13
Solution Processed ­WO3 and PEDOT:PSS Composite for Hole Transport Layer in ITO‑Free Organic…

is shown in Fig. 7a. The XPS core level spectra of W 4f, O shown in Table 1, the device with the W ­ O3/PEDOT:PSS
1s, C 1s, and S 2p of the W ­ O3/PEDOT–PSS are shown in composite layer shows the highest PCE of 7.81 ± 0.01%,
Fig. 7b–e. In Fig. 7b, there are two peaks at approximate with an open-circuit voltage (­ VOC) of 0.714 ± 0.02% V, a
34.6 eV and 36.8 eV corresponding to W 4­ f7/2 and W 4­ f5/2, short-circuit current (­ JSC) of 16.2 ± 0.01%, mA/cm2 and a
respectively [38], which is in good agreement with the exist- FF of 58.1 ± 0.03%. while the ­WO3 exhibited the lowest
ence of W­ 6+ in W
­ O3/PEDOT:PSS. The peak at 532.5 eV can ­JSC, FF, and PCE of (10.5 ± 0.02 mA/cm2, 56.0 ± 0.06, and
be attributed to the binding energy of C=O, which was most 3.12% ± 0.01%). The low photovoltaic parameter values of
likely to be partly caused by the water molecule attached to the latter can be attributed to its low charge-collection effi-
the surface of composites. The C 1s core level spectrum, ciency (due to its high thickness), which in turn increases
which can be curve-fitted into four peaks at 284.7, 284.8, charge recombination. The improved PCE of the ­WO3/
286.3, and 289.2 eV, corresponding to s­ p2 hybridized car- PEDOT:PSS composite is attributed to the high electrical
bon, ­sp3 hybridized carbon, C–O, and C=O, respectively conductivity of the PEDOT:PSS. The external quantum effi-
[39]. Regarding the S 2p spectra in Fig. 7e, the two strong ciency (EQE) of each sample cell was measured, as shown
peaks at 162.5 and 163.9 eV are assigned to those of S 2­ p3/2 in Fig. 8c. As expected, the cells with the ­WO3/PEDOT:PSS
and S ­2p1/2, respectively. layer displayed a higher EQE value in the visible and near
infrared regions in comparison with the bare W ­ O3 (which
Photovoltaic Studies is consistent with the increase of J­ SC), benefiting from
increased intensity of the incident light to the p–n junction.
The constructed solar device was shown schematically in In the ultraviolet band, the unexpected lower EQE value was
Fig. 8a. The illuminated current–voltage (J–V) character- caused by parasitic absorption of ­WO3 material. This can
istics of solar cells under 1.5 G irradiation (100 mW/cm2) be largely attributed to the recombination that occurred on
using PEDOT:PSS, ­WO3 and ­WO3/PEDOT:PSS compos- the surface or in the bulk, which reduced the carrier den-
ites layers are shown in Fig. 8b. The photovoltaic param- sity and thus caused the decrease of short-circuit current
eters of all the devices are summarized in Table  1. As density (Fig. 8b). Unlike W ­ O3/PEDOT:PSS displayed a

Fig. 7  XPS spectra of ­WO3/PEDOT:PSS composite a survey, b W 4f, c O 1s, d C 1s and e S 2p

13
P. Gurudevi et al.

Fig. 8  a Schematic representation of the organic solar cell device bility test of PEDOT:PSS, ­WO3, and ­WO3/PEDOT:PSS composite
set up, b J–V characteristics curve of PEDOT:PSS, W­ O3, and ­WO3/ thin film samples
PEDOT:PSS composite thin film samples, c ECE spectra and d sta-

Table 1  Photovoltaic parameters of PEDOT:PSS, W


­ O3, and W
­ O 3/ of the DSSCs was measured with an AM 1.5 G solar simu-
PEDOT:PSS lator (Bunkoh-keiki Co., Ltd., OTENTO-SUN5) with and
Samples JSC (mA/cm2) VOC (V) FF η (%) without a UV cut-off filter (< 420 nm). The test was carried
out at RT with relative humidity of 71%.
PEDOT:PSS 9.35 ± 0.03 0.592 ± 0.03 35 ± 0.05 1.32 ± 0.03 After 500 h continuous testing, the device loss only 2.1%
WO3 10.5 ± 0.02 0.631 ± 0.02 56 ± 0.06 3.12 ± 0.02 of PCE from its initial value, which suggest that fabricated
WO3/ 16.2 ± 0.01 0.714 ± 0.02 58 ± 0.03 7.81 ± 0.01 device showed ultra stability. In order to further study the
PEDOT:PSS
performance of the cells, the dark current density versus
Data for each cell are based on five individual samples with standard voltage characteristic of the hybrid solar cells was meas-
deviation ured, and the results were plotted (Fig. 9a). The dark J–V
curves were simulated according to the thermionic emis-
high-extinction coefficient and parasitic absorption in this sion model [40]. By combining the thermionic emission
ultraviolet wavelength region. In order to explore the stabil- model and the curve we tested, the n, J­ s, and Φbi of the
ity of the prepared OSC in the practical devices, the device solar cells were extracted (see Table 2). The device with
was further tested the J–V parameters in multiple experi- ­WO3/PEDOT:PSS layers exhibited a J­ s value of 1.59 × ­10−7
ments (Fig. 8d). The long-term stability of the performance A/cm2, approximately a half of that of the W ­ O3 device

13
Solution Processed ­WO3 and PEDOT:PSS Composite for Hole Transport Layer in ITO‑Free Organic…

Fig. 9  a Dark J–V curve of electrode sample, b dynamic photocurrent response, c EIS Nyquist plots of PEDOT:PSS, W
­ O3 and ­
WO3/
PEDOT:PSS, and d transient PL spectra

Table 2  Diode ideality factor (n), reverse saturation current density to the improved contact ability of the PEDOT:PSS layer.
­(Js), and Schottky barrier height (Φbi) values of PEDOT:PSS, W ­ O3, Moreover, the PCE obtained in the present work is higher
and ­WO3/PEDOT:PSS
than previous published works [41–44] and the results are
Samples Js (A/cm2) n Φbi (eV) summarized in Table 3.
WO3 2.91 × ­10−7 2.65 0.815
Electrochemical and Photodynamic Response
WO3/PEDOT:PSS 1.59 × ­10−7 2.27 0.836

Electrical impedance spectroscopy (EIS) was performed


to investigate the charge transport dynamics of the OSCs
(which was 2.97 × ­10−7 A/cm2). The smaller n value of the fabricated with PEDOT:PSS, ­WO3 and ­WO3/PEDOT:PSS
PEDOT:PSS incorporated device implied a suppressed (Fig. 9b). This analysis allowed us to observe the current
recombination rate. Moreover, the improved J­ 0 and n param- response by applying alternating current voltage as a func-
eters are also a compelling evidence that the charge carriers tion of frequency; the OSCs with W ­ O3/PEDOT:PSS dem-
were collected by the electrode with higher efficiency thanks onstrated slightly lower charge transfer resistance, revealing

13
P. Gurudevi et al.

Table 3  Comparison results of Device name Type Solar power PCE (%) References
PCE between present work and
already reported works MoO3/Au/MoO3–PEDOT:PSS OSC 1.5 G (100 mW/cm ) 2
3.03 [41]
PEDOT:PSS OSC 1.5 G (100 mW/cm2) 4.41 [42]
MoO3 OSC 1.5 G (100 mW/cm2) 2.96 [42]
TiO2/MWCNT/PEDOT:PSS DSSC 1.5 G (100 mW/cm2) 0.34 [43]
MoS2 in α-MoO3 OSC 1.5 G (100 mW/cm2) 1.34 [44]
MoO3/PEDOT:PSS OSC 1.5 G (100 mW/cm2) 7.81 This work

that the holes were effectively transported from the active ­ O3 is due to boost the efficiency of charge transfer and
W
layer to the anode. The corresponding R ­ S value of cells separation.
employing ­WO3/PEDOT:PSS as HTL was 14.8 Ω∙cm2,
while the reference showed 21.7 Ω∙cm2. Lower ­RS indicates
that better interfacial contact and charge collection efficiency Conclusions
were obtained due to the addition of the conducting ­WO3/
PEDOT:PSS layer. To further investigate the separation In summary, hybrid nanocomposite material was prepared
and transfer efficiency of the photo-generated electrons and by mixing of ­WO3 nanoparticles in PEDOT:PSS conduct-
holes, photocurrents of ­WO3/PEDOT:PSS nanocompos- ing polymer and film of ­WO3/PEDOT:PSS composite suc-
ites under visible light irradiation were detected for four cessfully developed by cost-effective spin coating method.
on–off cycles. The transient photocurrent measurement was The PL intensity was lower for W ­ O3/PEDOT:PSS compos-
performed under the illumination of 380 nm and 520 nm ite layer than compared with ­WO3, indicating that greater
LEDs driven by a pulse generator (AVTECH, AV-1011-B). PL quenching due to exciton dissociation occurred in the
The duration and frequency were 400 μs and 50 Hz respec- active layer of the ­WO3/PEDOT:PSS composite. The OSCs
tively. The rise time and fall time of the light pulse are less using ­WO3/PEDOT:PSS as anodes showed a considerable
than 0.5 μs. Photocurrent was measured by connecting the improvement in PCE of 7.81 ± 0.1%, with an open-circuit
device to a digitizing oscilloscope (TEK, TDS540D) with voltage ­(VOC) of 0.714 ± 0.2%, V, a short-circuit current
the input impedance of 50 Ω. The devices were kept in the ­(JSC) of 16.2 ± 0.1%, mA/cm2 and a FF of 58.1 ± 0.3%. Addi-
dark until the measurement was carried out. Figure 9c shows tionally, the ­WO3/PEDOT:PSS composite anode exhibited
the photocurrent responses of PEDOT:PSS, W ­ O3 and W ­ O 3/ a lower charge-transport r and shorter decay than the ­WO3
PEDOT:PSS under visible light irradiation for four on–off film, as indicated by impedance spectroscopy and transient
cycles. From Fig. 9c, the current intensity rises to a steady PL spectra analysis. Our results suggest that high-quality
value when the light on, and rapidly decreases to the dark ­WO3/PEDOT:PSS hole transparent electrodes can be fab-
current value when the light off. At the moment of both light ricated via the combination of a conductive polymer com-
on and light off, the changes of currents are almost vertical, posite and the novel processing protocol, yielding high-
indicting the charge transport rates in the fabricated sam- performance OSCs.
ples are very fast. Among all of these materials, the ­WO3/
Supplementary Information  The online version contains supplemen-
PEDOT:PSS nanocomposite exhibits the highest photocur- tary material available at https://d​ oi.o​ rg/1​ 0.1​ 007/s​ 10876-0​ 22-0​ 2368-6.
rent response, suggesting the highest separation efficiency
and longest lifetime of the photo-generated electrons and Acknowledgements  The manuscript should not be submitted to more
holes in ­WO3/PEDOT:PSS nanocomposite. The result fur- than one publication for simultaneous consideration. The submitted
work should be original and should not have been published elsewhere
ther provides valid evidence that the homogeneous disper- in any form or language.
sion of ­WO3 on PEDOT:PSS can accelerate the separation
and transfer of charge carriers in the interfacial surface of Author Contributions  PG and PV, study conceptualization and writ-
­WO3/PEDOT:PSS nanocomposites, which is considered to ing (original draft) the manuscript. TS, data curation, formal analysis
and writing (review and editing), and funding acquisition and project
the contribution to the enhancement of the photoconversion administration.
efficiency. The charge transfer could be further affirmed
by transient PL spectra, as displays in Fig. 9d. The ­WO3/ Data Availability  The data that support the findings of this study are
PEDOT:PSS shows a shorter decay time τ value 8.01 ns available from the corresponding author, upon reasonable request.
than ­WO3 (11.52 ns), related to a nonradiative pathway and
rapid interfacial electron injection efficiency at the W ­ O 3/ Declarations 
PEDOT:PSS layer composite. The improved photocurrent Conflict of interest  The authors declare they have no competing of in-
and life time of W ­ O3/PEDOT:PSS compared with bare terest.

13
Solution Processed ­WO3 and PEDOT:PSS Composite for Hole Transport Layer in ITO‑Free Organic…

References 26. V. Shrotriya, G. Li, Y. Yao, C.-W. Chu, and Y. Yang (2006). Appl.
Phys. Lett. 88.
27. G. Li, C.-W. Chu, V. Shrotriya, J. Huang, and Y. Yang (2006).
1. N. S. Sariciftci, L. Smilowitz, A. J. Heeger, and F. Wudl (1992).
Appl. Phys. Lett. 88.
Science 258, 1474.
28. M. Y. Chan, C. S. Lee, S. L. Lai, M. K. Fung, F. L. Wong, H. Y.
2. G. Yu, J. Gao, J. C. Hummelen, F. Wudl, and A. J. Heeger (1995).
Sun, K. M. Lau, and S. T. Lee (2006). J. Appl. Phys. 100.
Science 270, 1789.
29. H. Choi, B. S. Kim, M. J. Ko, D.-K. Lee, H. Kim, S. H. Kim, and
3. C. J. Brabec, N. S. Sariciftci, and J. C. Hummelen (2001). Adv.
K. Kim (2012). Org. Electron. 13, 959.
Funct. Mater. 11, 15.
30. C. Tao, S. Ruan, G. Xie, X. Kong, L. Shen, F. Meng, C. Liu, X.
4. G. Li, R. Zhu, and Y. Yang (2012). Nat. Photonics 6, 153.
Zhang, W. Dong, and W. Chen (2009). Appl. Phys. Lett. 94.
5. W. Xu, X. Zhu, X. Ma, H. Zhou, X. Li, S. Y. Jeong, H. Y. Woo,
31. Z. Hu, J. Zhang, X. Chen, S. Ren, Z. Hao, X. Geng, and Y. Zhao
and Z. Zhou (2022). J. Mater. Chem. A 10, 13492.
(2011). Sol. Energy Mater. Sol. Cells 95, 2173.
6. W. Wang, F. Qin, X. Jiang, X. Zhu, L. Hu, C. Xie, L. Sun, W.
32. Y. Wang, Q. Luo, N. Wu, Q. Wang, H. Zhu, L. Chen, et al. (2015).
Zeng, and Y. Zhou (2022). Org. Electron. 87.
ACS Appl. Mater. Interfaces 7, 7170.
7. R. Hermi, M. Mahdouani, R. Bourguiga, and S. Mahato (2022).
33. J. Ram, R. G. Singh, R. Gupta, V. Kumar, F. Singh, and R. Kumar
Micro–Nanostruct. 168.
(2019). J. Electron. Mater. 48, 1174.
8. W. Zeng, R. Ye, C. Yuan, Y. Shi, Q. Niu, W. Tan, J. Huang, R.
34. J. Ram, R. G. Singh, F. Singh, V. Kumar, V. Chauhan, R. Gupta,
Xia, and Y. Min (2022). Synth. Met. 290.
U. Kumar, B. C. Yadav, and R. Kumar (2019). J. Mater. Sci.
9. A. A. Lima, G. TaquesTractz, A. G. Macedo, F. Thomazi, P. R. P.
Mater. Electron. 30, 01728.
Rodrigues, and C. A. Dartora (2022). Opt. Mater. 132.
35. Y. G. Xu, H. Xu, L. Wang, J. Yan, H. M. Li, Y. H. Song, L. Y.
10. Y. Wang, N. Li, M. Cui, Y. Li, X. Tian, X. Xu, Q. Rong, D. Yuan,
Huang, and G. B. T. Cai (2013). Dalton Trans. 42, 7604.
G. Zhou, and L. Nian (2021). Org. Electron. 99.
36. H. Hwan Jung, D. Ho Kim, C. Su Kim, T.-S. Bae, K. Bum Chung,
11. D. Alemu, H.-Y. Wei, K.-C. Ho, and C.-W. Chu (2012). Energy
and S. Yoon Ryu (2013). Appl. Phys. Lett. 102.
Environ. Sci. 5, 9662.
37. M. L. Kaplan, S. R. Forrest, P. H. Schmidt, and T. Venkatesan
12. J. Saghaei, A. Fallahzadeh, and T. Saghaei (2015). Org. Electron.
(2012). J. Appl. Phys. 732, 120.
24, 188.
38. J. Di, J. X. Xia, Y. P. Ge, H. P. Li, H. Y. Ji, H. Xu, Q. Zhang, H.
13. Y. Xia and J. Ouyang (2012). Org. Electron. 13, 1785.
M. Li, and M. N. Li (2015). Appl. Catal. B 168, 51.
14. S. De, T. M. Higgins, P. E. Lyons, E. M. Doherty, P. N. Nirmalraj,
39. X. Xu, J. Shen, N. Li, and M. Ye (2014). J. Alloy Compd. 616, 58.
W. J. Blau, J. J. Boland, and J. N. Coleman (2009). ACS Nano 3,
40. Y. Sun, Z. Yang, P. Gao, J. He, X. Yang, J. Sheng, S. Wu, Y.
1767.
Xiang, and J. Ye (2016). Nanoscale Res. Lett. 11, 56.
15. H. Yuk, B. Lu, and X. Zhao (2019). Chem. Soc. Rev. 48, 1642.
41. Md. Maniruzzaman, M. A. Rahman, K. Jeong, and J. Lee (2014).
16. B. Lu, H. Yuk, S. Lin, N. Jian, K. Qu, J. Xu, and X. Zhao (2019).
Mater. Sci. Semicond. Process. 27, 114.
Nat. Commun. 10, 1043.
42. W. Zhang, W. Lan, M. H. Lee, J. Singh, F. Zhu, C. Dwivedi, T.
17. D.-W. Wang, F. Li, J. Zhao, W. Ren, Z.-G. Chen, J. Tan, Z.-S. Wu,
Mohammad, V. Bharti, A. Patra, S. Pathak, and V. Dutta (2018).
I. Gentle, G. Q. Lu, and H.-M. Cheng (2009). ACS Nano 3, 1745.
Sol. Energy 162, 78.
18. G. A. Snook, P. Kao, and A. S. Best (2011). J. Power Sources 196,
43. Q. A. Yousif, K. M. Mahdi, and H. A. Alshamsi (2020). Optik Int.
1.
J. Light Electron. 219.
19. R. R. Søndergaard, M. Hösel, and F. C. Krebs (2013). J. Polym.
44. A. A. Bortoti, A. de Freitas Gavanski, Y. R. Velazquez, A. Galli,
Sci. B 51, 16.
and E. G. de Castro (2017). J. Solid State Chem. 252, 111.
20. S. J. Kwon, J. H. Kang, S. J. Kim, W.-G. Koh, H. J. Song, and S.
Lee (2018). Macromol. Res. 26, 410.
Publisher's Note Springer Nature remains neutral with regard to
21. W. Jang, S. Ahn, S. Park, J. H. Park, and D. H. Wang (2016).
jurisdictional claims in published maps and institutional affiliations.
Nanoscale 8, 19557.
22. Y. Xia, K. Sun, and J. Ouyang (2012). Adv. Mater. 24, 2436.
Springer Nature or its licensor (e.g. a society or other partner) holds
23. J. J. Lee, S. H. Lee, F. S. Kim, H. H. Choi, and J. H. Kim (2015).
exclusive rights to this article under a publishing agreement with the
Org. Electron. 26, 191.
author(s) or other rightsholder(s); author self-archiving of the accepted
24. A. W. Hains and T. J. Marks (2008). Appl. Phys. Lett. 92.
manuscript version of this article is solely governed by the terms of
25. M. D. Irwin, D. B. Buchholz, A. W. Hains, R. P. H. Chang, and T.
such publishing agreement and applicable law.
J. Marks (2008). Proc. Natl Acad. Sci. USA 105, 2783.

13

You might also like