Download as pdf or txt
Download as pdf or txt
You are on page 1of 33

Accepted Manuscript

Facile Control of Intra- and Inter-particle Porosity in Template-Free Synthesis of


Size-Controlled Nanoporous Titanium Dioxides Beads for Efficient Organic-Inorganic
Heterojunction Solar Cells

Ganapathy Veerappan, Sora Yu, Dong Hwan Wang, Wan In Lee, Jong Hyeok Park

PII: S0378-7753(14)02162-4
DOI: 10.1016/j.jpowsour.2014.12.123
Reference: POWER 20402

To appear in: Journal of Power Sources

Received Date: 21 September 2014


Revised Date: 1 December 2014
Accepted Date: 24 December 2014

Please cite this article as: G. Veerappan, S. Yu, D.H. Wang, W.I. Lee, J.H. Park, Facile Control of Intra-
and Inter-particle Porosity in Template-Free Synthesis of Size-Controlled Nanoporous Titanium Dioxides
Beads for Efficient Organic-Inorganic Heterojunction Solar Cells, Journal of Power Sources (2015), doi:
10.1016/j.jpowsour.2014.12.123.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. 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.
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

Facile Control of Intra- and Inter-particle Porosity in Template-Free

Synthesis of Size-Controlled Nanoporous Titanium Dioxides Beads for

Efficient Organic-Inorganic Heterojunction Solar Cells

PT
Ganapathy Veerappan1, Sora Yu2, Dong Hwan Wang3, Wan In Lee2*, Jong Hyeok Park1*

RI
1
SKKU Advanced Institute of Nanotechnology (SAINT) and School of Chemical Engineering,

SC
Sungkyunkwan University, Suwon 440-746, Republic of Korea
2
Department of Chemistry, Inha University, Incheon 402-751, Republic of Korea
3

U
School of Integrative Engineering, Chung-Ang University, 84 Heukseok-Ro, Dongjak-gu,
AN
Seoul 156-756, Republic of Korea.
M
D
TE
C EP

*Email: lutts@skku.edu (JHP) and wanin@inha.ac.kr (WIL)


AC

1
ACCEPTED MANUSCRIPT

ABSTRACT

In thin film solid-state heterojunction solar cells (HSCs), titanium-dioxide (TiO2) electrodes need

to be optimized to have large specific surface area, controllable pore sizes, and superior light

scattering properties. In this study, we synthesize hierarchical nanoporous TiO2 beads with sub-

PT
micron diameters by a template-free, fast, and low-temperature synthetic scheme to satisfy the

RI
aforementioned requirements for HSCs. These nanoporous TiO2 beads are composed of

numerous TiO2 nano crystallites that provide mesopores, and the inter-particle distances of size-

SC
controlled TiO2 beads can provide additional controllable macropores. We report the first

successful application of TiO2 bead films (SP250, SP450) with controllable hierarchical

U
nanostructure to be sensitized with Sb2S3 for all-solid-state heterojunction solar cells (Sb-HSCs).
AN
The Sb-HSCs made using the controlled TiO2 beads as photoanodes exhibit a superior light-to

electricity conversion efficiency of 4.8%, yielding more than 15% enhancement in comparison
M

with that (3.6%) of commercial TiO2 nanoparticle (NP40) electrodes. The well-tailored
D

photoanode with high surface area, fewer grain boundaries, multi-scale pore structure, and
TE

enhanced optical scattering results in much better infiltration of hole-conducting materials,

decreased recombination with increased electron lifetime, and enhanced light scattering, which
EP

result in the enhanced photovoltaic properties.


C
AC

Keywords: Heterojunction solar cells, Antimony sulfide, Titanium dioxides beads, enhanced
electron lifetime, scattering effect

2
ACCEPTED MANUSCRIPT

1. INTRODUCTION

Conventional liquid dye-sensitized solar cells (DSSCs) with a network of interconnected

PT
TiO2 nanoparticles with dyes have been attracting interest as a promising alternative to

conventional silicon photovoltaic devices because of their low cost and high efficiency [1, 2]. At

RI
present, although the reported efficiency of liquid-based DSSCs has exceeded 12.0%, one of the

SC
serious obstacles for their application is the evaporation and leakage of the liquid electrolyte

from the DSSCs [3]. To address these issues, several metal chalcogenides (inorganic

U
semiconductor quantum dots) combined with solid-state electrolytes have been proposed and
AN
successfully realized [4-10]. Metal chalcogenides have various advantages as sensitizers such as

a high absorption coefficient, multiple exciton generation, good stability, and bandgap energy
M

that can be tuned by controlling the size. Thus, as dye replacements, metal chalcogenides,

particularly sulfides, selenides, and tellurides, have been extensively studied as sensitizers in
D

hope of attaining high photovoltaic performance [11-16].


TE

One of the key components in solid-state dye-sensitized solar cells (SDSSCs) is the
EP

nanoporous TiO2 electrode, which transfers electrons from quantum dot sensitizers to the FTO

substrate and concurrently allows the quantum dots to sit on the TiO2 surface and hole
C

conductors to be connected with the anchored quantum dots. In recent years, Sb2S3 has been
AC

successfully applied as a sensitizer by few groups in heterojunction solar cells due to its high

absorption coefficient (1.8 × 105 cm −1 at 450 nm) and mid-range optical bandgap of 1.7 eV [6,

17-21]. However, Sb2S3 sensitized all-solid-state heterojunction solar cells (Sb-HSCs) still suffer

from high recombination possibility mainly due to the non-optimized photoanodes (disordered

TiO2 nanostructure), which have many trap states and grain boundaries [17, 18]. In previous

3
ACCEPTED MANUSCRIPT

studies with DSSCs based on a conventional liquid electrolyte, tailoring of the TiO2

nanostructures strongly influenced the photovoltaic performance [22, 23].

So far, various TiO2 nanostructures, including nanoparticles, nanowires, nanotubes,

nanosheets, mesoporous nanostructures, three-dimensional (3D) structures, hollow fibers, and

PT
beads have been successfully synthesized and applied as photoanodes for DSSCs, and they have

RI
been obtained using sol-gel synthesis, hydro/solvothermal methods, electrospray, electrospinning,

anodization, and template-assisted methods [24-28]. Among these, mesoporous or nanoporous

SC
sub-micron sized beads offer great potential for the cells due to their bifunctions, providing

sufficient surface area and improved light scattering, which are essential for the photoanode in

U
sensitized solar cells [28-34]. In DSSCs, TiO2 beads size-dependant photovoltaic performance
AN
was reported by several groups in the recent past [28, 30-32]. Depending on the beads size, dye-

loading or QD sensitization, light scattering, pore size (internal and external pores), liquid
M

electrolyte or hole transporting material penetration and electron diffusion rate can be varied. So
D

the as-synthesized photoanode material should satisfy all the above criteria to yield high
TE

efficiency. But, till date, researchers failed to realize the importance of beads size-dependant

photovoltaic property in SDSSCs or Sb-HSCs.


EP

A high surface area in TiO2 electrodes is essential for sufficient loading of Sb2S3, which

will generate electrons after absorbing solar light. Also, the external pores formed via inter-
C

particle distance further act as the loading sites for Sb2S3 while simultaneously helping the
AC

penetration of hole-transporting material (HTM) (e.g. P3HT). The number of grain boundaries

also might be minimized due to the sub-micron size of TiO2 to suppress the charge

recombination and induce strong light scattering within a very thin film (< 1 µm). However, for

Sb-HSCs, there have only been few reports on TiO2 nanostructure-dependent power conversion

4
ACCEPTED MANUSCRIPT

efficiency to date [17, 18]. The performance of these cells has been disappointing compared with

liquid DSSCs, because the commonly used TiO2 nanoparticle electrode for Sb-HSCs might

induce insufficient pore-filling of the hole-conducting material (e.g. P3HT), which has resulted

in poor contact of the polymer with the sensitizer in photoanodes. These factors lead to faster

PT
recombination kinetics in between the interfaces, and thus can have a negative effect on the

RI
open-circuit voltage (VOC) of Sb-HSCs. In most conventional Sb-HSCs, 40 to 50-nm-sized

(NP40) randomly interconnected TiO2 nanoparticle films are employed, which have several

SC
disadvantages, such as high charge recombination due to poor particle connectivity, poor light

scattering, and poor hole-conducting material (HTM) penetration as described above [6, 20, 21,

24].
U
AN
In this study, we report the synthesis of highly crystalline and nanoporous TiO2 beads

with different bead sizes through a two-step synthesis method, and their successful application
M

for the first time in solution-processed and cost-effective Sb-HSCs. First, the TiO2 beads were
D

controlled to have two different sub-micron diameters (250-nm beads: SP250; 450-nm beads:
TE

SP450) to improve light scattering ability and to investigate the dependency of cell performance

on bead size. Second, the individual TiO2 SP250 and SP450 beads are composed of assembled
EP

TiO2 nanocrystals with size around 15 nm to maximize their specific surface area. Third, the

TiO2 beads with different diameters can cause different porosity and inter-particle pore size in
C

TiO2 photoanodes versus conventional TiO2 nanoparticle film, which can also affect the device
AC

performance of Sb-HSCs. As a result, the nanoporous TiO2 beads of SP250 had the best cell

efficiency among various Sb-HSCs with different TiO2 photoanodes due to the diameter and

pore structure, which can optimize the sensitization of Sb2S3 quantum dots and light scattering,

5
ACCEPTED MANUSCRIPT

in addition to reducing charge recombination and internal cell resistance from efficient HTM

penetration and a reduced number of grain boundaries.

PT
2. EXPERIMENTAL SECTION

RI
2.1 Synthesis of TiO2 beads. Nanoporous TiO2 beads were synthesized by a 2-step process

consisting of controlled hydrolysis and subsequent hydrothermal reaction. In preparing 250-nm-

SC
sized beads, 1 mM methylamine and 48 mM distilled water were added to a mixed solution

containing 50 ml of ethanol (98%, Aldrich) and 40 ml of acetonitrile (99.8%, Aldrich), and

U
stirred for 5 min. 6 mM titanium tetra-isopropoxide (TTIP, Aldrich) stabilized in 10 ml of
AN
ethanol was then added to this solution, and the resultant milky solution was gently stirred for 1 h

to obtain amorphous beads. In preparing 450-nm-sized beads, the amount of added water was
M

decreased to 24 mM, while other reaction parameters remained the same. The obtained
D

amorphous beads were collected by centrifugation and then transferred to a 0.3-L titanium bomb
TE

containing a mixed solution of water (70ml) and ethanol (70ml). After a hydrothermal reaction at

230ºC for 5 h, the amorphous beads were converted to highly porous and crystallized spherical
EP

structures. The prepared beads were then collected by centrifugation, washed with ethanol

several times, and dried at 100ºC.


C

2.2 Device fabrication. TiO2 paste was prepared by mixing ethanol, ethyl cellulose, and alpha-
AC

terpineol with the synthesized nanoporous TiO2 beads powder. The remaining solvent was then

evaporated using a rotary evaporator to obtain a viscous paste. A 70-nm-thick TiO2 blocking

layer (BL-TiO2) was spin coated onto previously etched and cleaned fluorine-doped tin oxide

(FTO) glass substrates. The TiO2 beads paste was then deposited by a doctor blading method

6
ACCEPTED MANUSCRIPT

onto the BL-TiO2/FTO substrate and sintered at 500ºC for 30 min. To improve the contact

between TiO2 particles, 20 mM TiCl4 treatment was carried out at 70ºC for 12 h and sintered at

500ºC for 30 min. For comparison, a commercial nanoparticle TiO2 (NP40) electrode (TiO2 size

= ~ 50nm, EnB Korea) was also prepared, and the associated films were prepared by the doctor

PT
blading method. All the TiO2 (NP40, SP250, and SP450) electrodes were ~1.0µm thick. The

RI
Sb2S3 inorganic sensitizer was deposited by chemical bath deposition (CBD) [6, 19]. Next, the

prepared TiO2 electrodes were dipped into a CBD mixture containing 650 mg of SbCl3 in 2.5 ml

SC
of acetone, 25 ml of 1M Na2S2O3 aqueous solution, and 72.5 ml of deionized water at 6ºC.

Orange-colored amorphous stibnite layers that formed on the TiO2 electrodes were sintered at

U
330ºC in an argon atmosphere for 30 min [6, 19]. A P3HT solution (15mg mL−1 in 1, 2-
AN
dichlorobenzene) was spin coated onto Sb2S3-sensitized TiO2 electrodes at 500 rpm for 20 min

and dried at 90ºC for 30 min. Then, poly (3,4-ethylenedioxythiophene) doped with
M

poly(4stylenesulfonate) (PEDOT:PSS) diluted with five volumes of isopropyl-alcohol (IPA) was


D

introduced onto the TiO2/Sb2S3/P3HT layer by spin coating at 2000 rpm for 30 s. Finally,
TE

devices were completed by depositing a gold (Au) counter electrode by thermal evaporation with

a base pressure of 5×10−6.


EP

2.3 Characterization. X-ray diffraction (XRD) and Brunauer-Emmett-Teller (BET) analysis

were performed to determine the phase, surface area, pore size distribution, and particle size of
C

the TiO2 used in this work. XRD (D8 Discover, Bruker-AXS) diffractometer with CuKα
AC

radiation (λ = 1.54056 Å) were used to collect XRD data. Accelerated Surface Area and

Porosimetry System (ASAP 2010) was adopted to determine the BET surface area and pore

diameters in this work. The multipoint BET method and Langmuir method were used to

calculate the specific surface area of TiO2. N2 gas was used to determine the adsorption

7
ACCEPTED MANUSCRIPT

isotherms. The adsorption isotherm of nitrogen was measured in the range of relative pressure

(p/p0) from 0.05 to 0.3 at 77 K, where p0 is the saturated vapor pressure of the system.The

specimen was out-gassed at 353 K for 1 h before the measurements. The nanostructure

morphology of TiO2 beads and films were investigated by field-emission scanning electron

PT
microscopy (FE-SEM, JEOL JSM-7500F Japan) and high-resolution transmission electron

RI
microscope (HR-TEM, JEOL JEM-2100F Japan). The photocurrent-voltage (I-V) properties of

DSSCs were studied under AM 1.5 simulated solar light irradiation using a 300 W xenon lamp

SC
(Newport, USA) calibrated with a standard Si solar cell. The incident photon-to-current

conversion efficiency (IPCE) was measured as a function of wavelength ranging from 400 to 800

U
nm (PV Measurement, Inc.). UV-vis diffuse reflectance spectra were obtained for reflectance
AN
analysis. Electrochemical impedance spectroscopy (EIS) measurements were carried out under

1-sun conditions in a frequency range of 100 mHz to 100 kHz using an impedance analyzer
M

(Zahner IM6).
D
TE

3. RESULTS AND DISCUSSION

A schematic diagram for synthesizing the nanoporous TiO2 beads with sub-micron size is
EP

shown in Figure 1a. In brief, titanium tetra-isopropoxide (TTIP) was hydrolyzed with

methylamine, DI water, and ethanol to form amorphous TiO2 beads [30]. Then, the resultant
C

TiO2 beads were transferred and treated at 230ºC for 5 h in a hydrothermal reactor to yield
AC

crystalline nanoporous TiO2 beads. Figures 1b, 1c, and 1d show a schematic illustration of the

Sb2S3 sensitized nanoporous TiO2 beads in the Sb-HSCs device, the energy band diagram, and a

schematic depiction representing electron and hole transportation in the Sb-HSCs, respectively.

8
ACCEPTED MANUSCRIPT

Figure 2a shows the XRD spectra of the synthesized nanoporous TiO2 beads. XRD

analysis revealed that the synthesized TiO2 beads consisted of a highly pure anatase TiO2 phase

with no other impurity peaks. The crystal sizes of the TiO2 beads were estimated from the full

width half maximum (FWHM) of the anatase peak (101) using the Scherrer equation. From the

PT
(101) peaks, it was confirmed that the crystal sizes of both the SP250 and SP450 were almost

RI
identical (14 - 16 nm), which is in good agreement with the TEM analysis (Figures 2d, e). The

morphologies and shapes of the TiO2 beads are shown in Figures 2b-e. By controlling the water-

SC
to-TTIP ratio as introduced in the experimental section, we controlled the bead sizes between

250 and 450 nm (Figures 2b and c) [30]. The SP250 and SP450 beads have uniform diameters of

U
250 and 450 nm, respectively. In addition, the films prepared from SP250 and SP450 beads
AN
might possess multi-scale porous nanostructures with both nanopores (from intra-bead space)

and macropores (from inter-bead space), which was deemed helpful for high loading of Sb2S3
M

and efficient HTM infiltration inside the photoanodes. Figure S1 shows the SEM surface and
D

cross-sectional images of the TiO2 films on fluorine-doped tin oxide (FTO) substrate with SP250
TE

(Figures S1 a, b) and SP450 (Figures S1 c, d), respectively. The BET surface area and pore size

distribution of the nanoporous TiO2 beads and commercial TiO2 nanoparticles (hereafter NP40)
EP

were characterized by nitrogen-sorption experiments. The specific surface areas of SP250,

SP450, and NP40 were 121.3, 114.6, and 72.0 m2 g−1, respectively. The specific surface areas,
C

pore size distribution, and crystallite size of the beads and NP40 are summarized in Table 1.
AC

The efficiency obtained with the conventional NP40 is promising but still not high

enough compared to the reported values. As discussed previously, the cell efficiencies of Sb-

HSCs devices depend upon many factors, including the nanostructural morphologies of the

photoanodes [6, 17, 18]. To improve upon conventional device nanostructures with random TiO2

9
ACCEPTED MANUSCRIPT

networks (NP40 photoanodes), we applied nanoporous TiO2 beads as alternative photoanodes.

The photocurrent-voltage (I-V) characteristics of the Sb-HSCs devices (with NP40, SP250, and

SP450) were measured under AM 1.5G 100 mW cm−2 illumination and are shown in Figure 4a.

The corresponding photovoltaic parameters are summarized in Table 2. The Sb-HSCs device

PT
prepared with SP250 beads gives the highest efficiency.

RI
Interestingly, SP250 and SP450 had similar BET surface areas and internal pore sizes,

even though the beads had different diameters. Due to the large bead diameters of SP450 over

SC
SP250, they possess much bigger external pore size than the SP250. To investigate Sb2S3

distribution in the photoanodes, HR-TEM and EDX elemental mapping of the Sb2S3 sensitized

U
NP40 and SP250 samples were obtained, as shown in Figure 3. Due to the high surface area
AN
combined with uniform intra and inter pore structure of the SP250 electrodes, Sb2S3 sensitizers

were uniformly distributed in the TiO2 electrode with minimal aggregation (Figure 3b) compared
M

to the NP40/Sb2S3 electrode. The porosity and surface area of the photoanodes can affect the
D

Sb2S3 sensitization in two ways. During the chemical bath deposition of Sb2S3, the antimony and
TE

sulfide precursors need to infiltrate into the TiO2 photoanodes for uniform deposition, but this

was limited by the narrow and disordered pore structures of NP40, leading to Sb2S3 aggregation
EP

in NP40 photoanodes.

To investigate the positive effects of utilizing nanoporous TiO2 beads for photoanodes in
C

Sb-HSCs, two variables were optimized to obtain maximum photovoltaic performance for each
AC

TiO2 material (SP250, SP450 and NP40): TiO2 film thickness and Sb2S3 sensitization time [6].

Interestingly, all devices showed the best performance when the TiO2 film thickness and

sensitization were around 1 µm and 150 min, respectively (Figures S2a and Table S1). The cell

efficiencies of the NP40 increased with Sb2S3 sensitization time (30 min: 2.3%, 150 min: 3.2%).

10
ACCEPTED MANUSCRIPT

The increase in efficiency is mainly associated with the thicker Sb2S3 layers, which can improve

the light harvesting. However, when we increased the Sb2S3 sensitization time from 150 to 270

min, a sudden decrease in efficiency and VOC was observed. This decrease in efficiency and VOC

is mainly attributed to poor electron injection from the thick Sb2S3 layer into the TiO2 layer, as

PT
well as poor hole extraction from the thick Sb2S3 layer into the hole transporting layer. The

RI
optimized Sb2S3 sensitization time (150 min) and TiO2 film thickness (1 µm) were utilized in all

the TiO2 nanostructures used in this work. Additionally, digital photographic images of the

SC
photoanodes are shown in Figure S2b for 1) TiO2 on FTO glass substrate, 2) Sb2S3 sensitized on

TiO2 before sintering, and 3) Sb2S3 sensitized on TiO2 after sintering.

U
Porosity of the photoanode has a vital role in the performance of the heterojunction solar
AN
cells [18]. Since, when the photoanode material has poor intra and inter pores then the precursors
M

(antimony and sulfur precursor) used for sensitizing the photoanodes will be difficult to infiltrate

and thus can affect the Sb2S3 sensitization (shown in Figure 3c-g); which in turn affects the
D

photovoltaic performance of the Sb-HSCs. In the case of TiO2 beads, the high surface area
TE

combined with multi-scale porous structures (intra and inter pore) can facilitate the precursor

infiltration and uniform distribution in the TiO2 electrode (shown in Figure 3h-l). Along with the
EP

uniform Sb2S3 sensitization, the formation of macropores from inter-and intra-bead distance

could enhance the pore filling of the HTM. Pore filling of HTM in SDSSCs or Sb-HSCs is a key
C

problem in enhancing the device performance. But in this work, the successful use of intra-and
AC

inter porous beads distance have led to better infiltration of HTM into the photoanodes, resulting

in better charge transfer between stibnite/HTM interfaces [18]. The two nanoporous TiO2 bead-

based Sb-HSCs devices exhibited much higher FF and Voc values than the NP40 device, since

SP250 and SP450 possess larger surface area, fewer interparticle grain boundaries, and unique

11
ACCEPTED MANUSCRIPT

pore structures. Particularly, the enhanced VOC originated from the decrease of electron

recombination in the TiO2 photoanodes, as well as the improved interface between the TiO2

electrode and HTM, leading to enhanced electron lifetime [32]. In general, VOC is decided by the

gap between the Fermi level of TiO2 (EF, TiO2) and the formal potential of the HTM, so less

PT
recombination resulted in higher VOC. The reduced recombination was confirmed by the

RI
impedance analysis (which will be discussed later). In addition, the enhanced short-circuit

current density (Jsc) values of the SP250 bead-based Sb-HSCs over the NP40-based Sb-HSCs

SC
were attributed to the efficient light scattering from the particle diameter with sub-micron size

[17, 18]. Figure 4b shows the diffuse reflectance spectra for the SP250, SP450 and NP40

U
electrodes. All the samples showed good diffuse reflectance in the visible range of around 400-
AN
500 nm, but a sudden decrease was identified in the reflectance spectra for NP40 at 500-800 nm

[30]. Since the NP40 films are semi-transparent with poor light scattering function, electrodes
M

made with TiO2 beads can utilize the light scattered in the range of 500-800 nm and thus improve
D

the photon-to-electron conversion efficiency. Although SP450 electrodes showed better diffuse
TE

reflectance around 500-800nm when compared with the SP250 electrodes; the device made with

SP450 showed poor photovoltaic property due to the poor Sb2S3 sensitization.
EP

The IPCE spectra of the TiO2 bead-based Sb-HSCs are given in Figure 4d, and the

increment of the Jsc value was also confirmed by IPCE analysis. The bigger particle size of
C

SP450 induced big interparticle distance, and the low volumetric surface area of the TiO2 film
AC

might lead to poor Sb2S3 sensitization, which resulted in poor charge generation efficiency even

though SP450 has similar intra-bead porosity. To further support the beneficial effects of our

nanoporous beads and its advantages in Sb-HSCs, we obtained cross-sectional SEM images of

our device (Figure 4c). The combination of nanopores and macropores in the SP250 film induces

12
ACCEPTED MANUSCRIPT

ideal nanostructure for the uniform sensitization of Sb2S3 and easy penetration of P3HT for the

solid interface.

To investigate the morphological effects of controlling individual TiO2 nanostructures

(NP40, SP250 and SP450) on the photovoltaic performance, EIS was measured under 1-sun

PT
illumination conditions, and the results are given in Figure 5. Figure 5a shows the Nyquist plot of

RI
three different Sb-HSCs. The high frequency region denotes the charge transfer resistance at the

Au/P3HT or Sb2S3/P3HT interfaces, while the low frequency region is associated with the

SC
electron transport kinetics at the TiO2/Sb2S3/P3HT interfaces [35-40]. As shown in Figure 5a, the

SP250 sample exhibited the lowest charge transfer resistance at both the high frequency and low

U
frequency regions, indicating it had much faster electron transport kinetics. Importantly, the
AN
decrease in charge transfer resistance of the SP250 photoanode sample implied an efficient

electron transport at the TiO2 beads/Sb2S3/P3HT interface and smaller electron recombination
M

due to interconnected nanostructures, which resulted in smaller electron transport resistance.


D

This was further confirmed by the lifetime measurement from EIS [35, 40].
TE

From the impedance bode phase plots, the electron lifetime (τ) for the three different

TiO2 photoanodes can be calculated (Figure 5b) [37]. The different low frequency peaks for the
EP

three TiO2 photoanodes seems to be due to the difference in the charge transfer process, which

indirectly gives the electron lifetime for all the photoanodes used in this work. The electron
C

lifetime was calculated from the low frequency peaks as 0.14 ms for SP250 and 0.11 ms for
AC

SP450. Devices prepared with NP40 showed much smaller lifetime (0.07 ms) than the TiO2

beads photoanodes. The higher electron lifetime explains the decrease in charge recombination

reaction in the photoanodes. This result might have arisen from the low transport resistance and

reduction of electron loss in the grain boundaries caused by the properly interconnected

13
ACCEPTED MANUSCRIPT

nanoporous TiO2 beads, allowing for more effective capture of electrons. Lastly, the high surface

area and nanoporous pores accessible for HTM penetration also play an important role in the

improved electron lifetime and photovoltaic performance [41-43].

PT
4. CONCLUSION
In summary, hierarchical nanoporous TiO2 beads were prepared by a hydrothermal

RI
method and successfully utilized as a photoanode in Sb2S3-sensitized all-solid-state

SC
heterojunction solar cells (Sb-HSCs) and compared with conventional commercial TiO2

nanoparticle (NP40) electrodes. The devices prepared with nanoporous SP250 electrodes showed

U
considerable improvements in all the photovoltaic parameters (from JSC: 15.2 mA cm−2, VOC:
AN
0.530V, FF: 58.4, η: 4.8% to JSC: 14.8 mA cm−2, VOC: 0.498V, fill factor(FF): 49.6, η: 3.6% )

compared to the NP40 photoanode. This enhancement is mainly attributed to the larger active
M

surface area (121.3 m2 g−1 for SP250 vs. 72.0 m2 g−1 for NP40), reduced recombination, longer

electron lifetime, and light scattering ability. Notably, by using the template-free synthesis
D

method, it is possible to prepare a well-crystallized nanoporous TiO2 nanostructure with different


TE

diameters in large scale, which makes it beneficial for both commercial and other energy
EP

applications.
C

ACKNOWLEDGMENTS
AC

This work was supported by a grant from the NCRC program (2011-0006268) of National

Research Foundation (NRF) funded by the Korea Ministry of Education, Science and

Technology (MEST).

14
ACCEPTED MANUSCRIPT

APPENDIX A: Supplementary data

Supplementary results associated with the additional SEM images showing surface and cross-

sectional view of the TiO2 spheres (SP250, SP450), I-V characteristics of the nanoparticle TiO2

photoanodes with different stibnite sensitization time, photographic images of the stibnite

PT
sensitized TiO2 and their corresponding summarized photovoltaic data in a Table.

RI
REFERENCES

SC
(1) B. O. Regan, M. Gratzel, Nature 353 (1991)737-740.

(2) A. Hagfeldt, G. Boschloo, L. Sun, L. Kloo, H. Pettersson, Chem. Rev. 110 (2010) 6595-

6663.
U
AN
(3) A. Yella, H. W. Lee, H. N. Tsao, C. Yi, A. K. Chandiran, M. K. Nazeeruddin, E.W. G.

Diau. C. Y. Yeh, S. M. Zakeeruddin, M. Gratzel, Science 334 (2011) 629-634.


M

(4) P. V. Kamat, Acc. Chem. Res. 45 (2012) 1906-1915.


D

(5) A. J. Nozik, M. C. Beard, J. M. Luther, M. Law, R. J. Ellingson, J. C. Johnson, Chem.


TE

Rev. 110 (2010) 6873-6890.

(6) J. A. Chang, J. H. Rhee, S. H. Im, Y. H. Lee, H. J. Kim, S. I. Seok, M. K. Nazeeruddin,


EP

M. Grätzel, Nano. Lett. 10 (2010) 2609-2612.

(7) I. Chung, B. Lee, J. He, R. P. H. Chang, M. G. Kanatzidis, Nature 485 (2012) 486-489.
C

(8) M. M Lee, J. Teuscher, T. Miyasaka, T. N. Murakami, H. J. Snaith, Science 338 (2012)


AC

643-647

(9) U. Bach, D. Lupo, P. Comte, J. E. Moser, F. Weissortel, J. Salbeck, H. Spreitzer, Nature

395 (1998) 583-585.

15
ACCEPTED MANUSCRIPT

(10) J. H. Heo, S. H. Im, J. H. Noh, T. N. Mandal, C. S. Lim, J. A. Chang, Y. H. Lee, H. J.

Kim, A. Sarkar, M. K. Nazeeruddin, M. Grätzel, Nature Photo. 7 (2013) 486-491.

(11) P. V. Kamat, J. Phys. Chem. C 112 (2008) 18737-18753.

(12) E. Barea, M. Shalom, S. Gimenez, I. Hod, I. M. Sero, A. Zaban, J. Bisquert, J. Am.Chem.

PT
Soc. 132 (2010) 6834-683.

RI
(13) H. Yang, W. Fan, A. Vaneski, A. S. Susha, W. Y. Teoh, A. L. Rogach, Adv. Func. Mat.

22 (2012) 2821-2829.

SC
(14) N. Tetreault, E. Arsenault, L. P. Heiniger, N. Soheilnia, J. Brillet, T. Moehl, S.

Zakeeruddin, G. A. Ozin, M. Grätzel, Nano. Lett. 11 (2011) 4579-4584.

U
(15) L. Etgar, D. Yanover, R. K. Capek, R. Vaxenburg, Z. Xue, B. Liu, M. K. Nazeeruddin,
AN
M. Grätzel, Adv.Funct. Mater. 23 (2013) 2736-2741.

(16) Christians, J. A. Kamat, P. V. ACS Nano 7 (2013) 7967-7974.


M

(17) J. C. Cardoso, C. A. Grimes, X. Feng, X. Zhang, S. Komarneni, M. V. B. Zanoni, N. Bao,


D

Chem. Commun. 48 (2012) 2818-2820.


TE

(18) E. L. Gui, A. M. Kang, S. S. Pramana, N. Yantara, N. Mathews, S. Mhaisalkar, J.

Electrochem. Soc. 159 (2012) B247-B250.


EP

(19) J. K. Kim, G. Veerappan, N. Heo, D. H. Wang, J. H. Park, J. Phys. Chem. C 118 (2014)

22672-22677.
C

(20) Y. Itzhaik, O. Niitsoo, M. Page, G. Hodes, J. Phys. Chem. C Lett. 113 (2009) 4254-4256.
AC

(21) N. Bansal, F. T. O'Mahony, T. Lutz, S. A. Haque, Adv.Energy Mater. 3 (2013) 986-990.

(22) N. Tetreault, M. Grätzel, Energy& Environ. Sci. 5 (2012) 8506-8516.

(23) E. J. W. Crossland, N. Noel, V. Sivaram, T. Leijtens, J. A. Alexander-Webber, H. Snaith,

J. Nature 495 (2013) 215-219.

16
ACCEPTED MANUSCRIPT

(24) T. Fukumoto, T. Moehl, Y. Niwa, M. K. Nazeeruddin, M. Grätzel, L. Etgar, Adv.Energy

Mater. 3 (2013) 29-33.

(25) L. Etgar, W. Zhang, S. Gabriel, S. G. Hickey, M. K. Nazeeruddin, A. Eychmuller, B. Liu,

M. Grätzel, Adv.Mater. 24 (2012) 2202-2206.

PT
(26) C. Y. Cho, J. H. Moon, Adv. Mater. 23 (2011) 2971-2975.

RI
(27) S. H. Ahn, D. J. Kim, W. S. Chi, J. H. Kim, Adv. Mater. 25 (2013) 4893-4897.

(28) H. J. Koo, Y. J. Kim, Y. H. Lee, W. I. Lee, K. Kim, N. G. Park, Adv. Mat. 20 (2008)195-

SC
199.

(29) W. Guo, Y. Shen, M. Wu, T. Ma, Chem. Commun. 48 (2012) 6133-6135.

U
(30) I. G. Yu, Y. J.Kim, H. J. Kim, C. Lee, W. I. Lee, J. Mater. Chem. 21 (2011) 532-538.
AN
(31) D. Chen, F. Huang, Y. B. Cheng, R. A. Caruso, Adv. Mat. 21 (2009) 2206-2210.

(32) Y. Chen, F. Huang, D. Chen, L. Cao, X. L. Zhang, R. A. Caruso, Y. B. Cheng,


M

ChemSusChem 4 (2011) 1498-1503.


D

(33) V. Ganapathy, E. H. Kong, Y. C. Park, H. M. Jang, S. W. Rhee, Nanoscale 6 (2014)


TE

3296-3301.

(34) G. Veerappan, D. W. Jung, J. Kwon, J. Choi, N. Heo, G. R. Yi, J. H. Park, Langmuir 30


EP

(2014) 3010-3018.

(35) C. S. Lim, S. H. Im, H. J. Kim, J. A. Chang, S. I. Seok, Phys. Chem. Chem. Phys. 14
C

(2012) 3622-3626.
AC

(36) S. H. Im, H. J. Kim, J. H. Rhee, C. S. Lim, S. I. Seok, Energy. Environ. Sci. 4 (2011)

2799-2802.

(37) M. Adachi, M. Sakamoto, J. Jiu, Y. Ogata, S. Isoda, J. Phys. Chem. B 114 (2006) 13872-

13880.

17
ACCEPTED MANUSCRIPT

(38) I. M. Sero, S. Gimenez, F. F. Santiago, R. Gomez, Q. Shen, T. Toyoda, J. Bisquert, Acc.

Chem. Res. 42 (2009) 1848-1857.

(39) P. Boix, Y. H. Lee, F. F. Santiago, S. H. Im, I. M. Sero, J. Bisquert, S. I. Seok, ACS

Nano 6 (2012) 873-880.

PT
(40) A. Dualeh, T. Moehl, M. K. Nazeeruddin, M. Gratzel, ACS Nano 7 (2013) 2292-2301.

RI
(41) Q. Zhang, C. S. Dandeneau, X. Zhou, G. Cao, Adv. Mat. 21 (2009) 4087-4108

(42) Y. J. Kim, M. H. Lee, H. J. Kim, G. Lim, Y. S. Choi, N. G. Park, K. Kim, W. I. Lee,

SC
Adv. Mat. 21 (2009) 3668-3673.

(43) M. Adachi, Y. Murata, J. Takao, J. Jiu, M. Sakamoto, F. Wang, J. Am. Chem. Soc. 126

(2004) 14943-14949.
U
AN
M
D
TE
C EP
AC

18
ACCEPTED MANUSCRIPT

Table 1. Surface area, pore size and crystallite sizes of all the photoanodes used in this work.

PT
BET Average pore diameter
Crystallite
surface (nm)
Sample names size

RI
area Internal External
(nm)
( m2 g−1 ) pore a pore a
NP40 46.2 72.0 - -

SC
SP250 14.2 121.3 10.5 90
SP450 15.3 114.6 11.8 190

U
AN
a
The internal pore indicates the confined spaces formed inside the beads whereas the external

pore is those generated among the neighboring beads or particles in the fabricated films. NP40:
M

nanoparticle TiO2, SP250: 250nm TiO2 beads, SP450: 450nm TiO2 beads.
D
TE
C EP
AC

19
ACCEPTED MANUSCRIPT

Table 2. Photovoltaic performances of all solid-state heterojunction solar cells with TiO2 beads

and nanoparticle TiO2 as a photoanodes.

PT
JSC VOC FF
Sample names (mA cm−2) (V) (%) (%)
NP40 14.8 0.498 49.6 3.6

RI
SP250 15.2 0.539 58.4 4.8
SP450 13.7 0.545 54.7 4.1

SC
NP40: nanoparticle TiO2, SP250: 250nm TiO2 beads, SP450: 450nm TiO2 beads

U
AN
M
D
TE
C EP
AC

20
ACCEPTED MANUSCRIPT

Figure captions

Figure 1. (a) Schematic illustrations of the nanoporous TiO2 beads synthesis, (b) Schematic

illustration of the stibnite sensitized nanoporous TiO2 beads device, (c) Energy level positions

and charge transfer process in the stibnite sensitized all-solid-state heterojunction solar cells, and

PT
(d) Schematic diagram showing the electron and HTM transferring route in the TiO2 beads and

RI
nanoparticles TiO2 electrodes in solid-state heterojunction solar cells.

SC
Figure 2. (a) XRD patterns of the nanoporous TiO2 beads, and FE-SEM images of (b, d) 450-nm

TiO2 beads (SP450) and (c, e) 250-nm TiO2 beads (SP250).

U
AN
Figure 3. HR-TEM images of Sb2S3 sensitized (a) NP40, and (b) SP250 TiO2 samples. EDX

elemental mapping acquired by STEM, where distribution of Ti, O, Sb, S atoms is given in (c-l)
M

for both the NP40 and SP250, respectively.


D

Figure 4. (a) Current-voltage (I-V) characteristics of all-solid-state heterojunction solar cells


TE

with TiO2 beads (SP250, SP450) and nanoparticle TiO2(NP40) photoanodes, (b) UV-vis diffuse
EP

reflectance spectra of the SP250, SP450 and NP40 electrodes, (inset showing the photographic

images of SP250, and NP40 electrodes), (c) FE-SEM cross-sectional images of the SP250
C

TiO2/Sb2S3/P3HT/Au device, and (d) IPCE spectra for both the TiO2 beads and NP40
AC

photoanodes all solid-state heterojunction solar cell device.

Figure 5 Electrochemical impedance spectra of Sb2S3-sensitized all solid-state heterojunction

solar cells with TiO2 beads (SP250, SP450) and nanoparticle TiO2 (NP40) photoanodes (a)

Nyquist plot, and (b) Bode-phase plot.

21
ACCEPTED MANUSCRIPT

PT
Figure 1.

RI
U SC
AN
M
D
TE
C EP
AC

22
ACCEPTED MANUSCRIPT

Figure 2.

PT
RI
U SC
AN
M
D
TE
C EP
AC

23
ACCEPTED MANUSCRIPT

Figure 3.

PT
RI
U SC
AN
M
D
TE
C EP
AC

24
ACCEPTED MANUSCRIPT

Figure 4.

PT
RI
U SC
AN
M
D
TE
C EP
AC

25
ACCEPTED MANUSCRIPT

Figure 5.

PT
RI
U SC
AN
M
D
TE
C EP
AC

26
ACCEPTED MANUSCRIPT

 Nanoporous TiO2 beads with uniform sizes were synthesized by hydrothermal method.

 TiO2 beads are composed of numerous TiO2 nanoparticles that provides large surface

area.

 TiO2 beads big pores are favorable for uniform QD sensitization and HTM penetration.

PT
 First successful application of TiO2 beads (SP250, SP450) for Sb2S3 sensitized all solid-

RI
state heterojunction solar cells.

 Devices made with TiO2 beads showed superior light-to electricity conversion efficiency

SC
(4.8%).

U
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT
Facile Control of Intra- and Inter-particle Porosity in Template-Free

Synthesis of Size-Controlled Nanoporous Titanium Dioxides Beads

for Efficient Organic-Inorganic Heterojunction Solar Cells

PT
Ganapathy Veerappan1, Sora Yu2, Dong Hwan Wang3, Wan In Lee2*, Jong Hyeok Park1*

RI
1
SKKU Advanced Institute of Nanotechnology (SAINT) and School of Chemical

Engineering, Sungkyunkwan University, Suwon 440-746, Republic of Korea

SC
2
Department of Chemistry, Inha University, Incheon 402-751, Republic of Korea
3

U
School of Integrative Engineering, Chung-Ang University, 84 Heukseok-Ro, Dongjak-gu,
AN
Seoul 156-756, Republic of Korea.
M

*Email: lutts@skku.edu (JHP) and wanin@inha.ac.kr (WIL)


D
TE
C EP
AC

1
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D

Figure S1. FE-SEM surface and cross-sectional images of (a, b) SP250 electrodes on FTO
TE

substrates and (c, d) SP450 electrodes on FTO substrates.


C EP
AC

2
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP

Figure S2. (a) Current-voltage (I-V) characteristics of solid-state heterojunction solar cells as

a function of Sb2S3 chemical bath deposition time on nanoparticle TiO2 (NP40) and (b)
C

Corresponding Digital photographic images of the TiO2 photoanodes, 1. TiO2 on FTO glass
AC

substrate, 2. Sb2S3 sensitized on TiO2 before sintering, and 3. Sb2S3 sensitized on TiO2 after

sintering.

3
ACCEPTED MANUSCRIPT

JSC VOC FF η
Sample names (mA cm−2) (V) (%) (%)
30 minSb2S3/ NP40 8.7 0.506 53.4 2.3
150min Sb2S3/ NP40 11.8 0.505 54.2 3.2
270min Sb2S3/NP40 12.0 0.478 48.9 2.8

PT
NP40: nanoparticle TiO2

RI
Table S1 Photovoltaic performance of all-solid-state heterojunction solar cells as a function
of Sb2S3 chemical bath deposition time on NP40 electrodes.

U SC
AN
M
D
TE
C EP
AC

You might also like