Вы находитесь на странице: 1из 9

Available online at www.sciencedirect.

com

ScienceDirect
Solar Energy 98 (2013) 203211
www.elsevier.com/locate/solener

Optimization of the dye-sensitized solar cell with anthocyanin


as photosensitizer
Chiang-Yu Chien, Ban-Dar Hsu
Department of Life Science, National Tsing Hua University, Hsinchu 30013, Taiwan
Received 16 January 2013; received in revised form 13 May 2013; accepted 24 September 2013
Available online 6 November 2013
Communicated by: Associate Editor Sam-Shajing Sun

Abstract
Dye-sensitized solar cells (DSSC) were fabricated using anthocyanin extracted from red cabbage (Brassica oleracea var. capitata f.
rubra), and the conditions that could maximize their performance were explored. The best light-to-electricity conversion eciency (g)
was obtained when the pH and the concentration of the anthocyanin extract were at 8.0 and 3 mM, respectively, and when the immersion
time for fabricating sensitized TiO2 lm was 15 min. Fabricating the DSSC in the presence of a coadsorbent, deoxycholic acid, at a molar
ratio of dye:coadsorbent of 1:40 was also found to be able to improve the g signicantly. The highest g reached was over 1.4%, which is
almost three times higher than previously reported conversion eciencies. In addition, the copigments associated with anthocyanin were
found to contribute signicantly to the performance of DSSC, and further purication of the anthocyanin extract led to a lower g.
2013 Elsevier Ltd. All rights reserved.

Keywords: Dye-sensitized solar cell; Anthocyanin; Deoxycholic acid; Copigments

1. Introduction
Dye-sensitized solar cell (DSSC) is an economical photovoltaic device that converts light energy into electricity.
It is only composed of a few components: photosensitizer,
transparent conductive oxide glass, TiO2 lm and electrolyte (ORegan and Gratzel, 1991). DSSC has many advantages over the conventional silicon-based solar cells, such
as having a colorful and transparent appearance, good portability and low cost (Kalowekamo and Baker, 2009). In
Abbreviations: DCA, deoxycholic acid; FF, ll factor; JSC, short-circuit
current density (mA/cm2); PA, puried anthocyanin; UPA, unpuried
anthocyanin; VOC, open-circuit voltage (V); g, light-to-electricity conversion eciency.
Corresponding author. Address: Department of Life Science, National
Tsing Hua University, No. 101, Section 2, Kuang-Fu Road, Hsinchu
30013, Taiwan. Tel.: +886 3 5742749; fax: +886 3 5715934.
E-mail address: bdhsu@life.nthu.edu.tw (B.-D. Hsu).
0038-092X/$ - see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.solener.2013.09.035

recent years, the light-to-electricity conversion eciency


(g) has improved to more than 10% with ruthenium polypyridyl complex (e.g., N719) as photosensitizer (Nazeeruddin et al., 2001; Qin and Peng, 2012). Nevertheless,
ruthenium is a trace element, which can become too expensive and inaccessible if widely used. Besides, ruthenium is a
heavy metal and its threat to the environment is a latent
risk (Gaiddon et al., 2005; Yasbin et al., 1980). Seeking
new photosensitizers is therefore inevitable.
Biological pigments can be decomposed naturally without any pollution. Their environmentally friendly properties, easy accessibility and high absorption in the visible
region make them good candidates. Anthocyanins, a group
of avonoids found in fruits, leaves and owers, are watersoluble plant pigments that usually carry vivid colors ranging from red to blue (Brouillard, 1983). They also have
long been considered to have health benets and widely
used as food colorants (Jackman et al., 1987; Lila, 2004).

204

C.-Y. Chien, B.-D. Hsu / Solar Energy 98 (2013) 203211

Compared to ruthenium-based dyes, anthocyanins are


metal free, nontoxic and widely available at very low
expense. They also have enough hydroxyl groups to bind
TiO2 nanocrystallites and have been shown to be able to
inject electrons into the TiO2 conduction band at an ultrafast rate when excited with visible light (Cherepy et al.,
1997).
There have been several studies on DSSC using anthocyanins as photosensitizer (Aduloju and Shitta, 2012; Calogero and Di Marco, 2008; Chang and Lo, 2010; Furukawa
et al., 2009; Hao et al., 2006; Luo et al., 2009; Polo and
Murakami Iha, 2006; Wongcharee et al., 2007). The g from
those studies, however, were generally quite low
(0.50.6%). Recently, a report by Calogero et al. (2012),
using sealed solar cells with enhanced electrodes (multilayer TiO2 lm plus a scattering layer) and anthocyanins
contained organic acids, demonstrated an eciency around
1.0%. In this study, the fabrication conditions that can
boost g were explored using anthocyanin from red cabbage. The highest g reached was over 1.4% under standard
AM 1.5 solar conditions.
2. Experimental
2.1. Materials
Fresh red cabbages (Brassica oleracea var. capitata f.
rubra) were purchased from a local market. The electrodes
were supplied by Tripod Technology Corporation, Taiwan.
TiO2 with average particle size of 1520 nm was coated
onto FTO glass (thickness 2.2 mm, resistivity 10 X/square,
average transparency higher than 80% in visible light
range) as a porous layer (thickness 12 lm, active area
28.3 mm2) of the anode, and a scattering layer (thickness
3 lm) consisted of bigger TiO2 particles (150 nm) was
further coated on the porous layer by screen printing
method. The cathode was FTO glass coated with nanoPt. All chemicals were purchased from Sigma Aldrich Co.
Deionized water was used throughout this study.
2.2. Water-based anthocyanin extraction
The leaves of red cabbage were washed thoroughly and
dried by tissue paper. Deveined leaves were then sliced and
homogenized by a blender (7012G, Waring Products, Inc.)
which had been pre-cooled to 0 C. The juice was obtained
by pressing the slurry through a stainless steel sieve, followed by centrifugation at 16,000g for 10 min at 4 C (3
18 K, Sigma, with the rotor 12158-H) to remove cell debris
and other undissolved substances. The supernatant was ltered rst by gravity through a lter paper (Whatman,
grade No. 2), and further through a syringe-driven membrane lter (Millipore, pore size 0.45 lm). The ltrates,
designated as the anthocyanin extract, were stored at
20 C until use. The whole procedure was conducted
under dim light.

2.3. Concentration measurement


The concentration of anthocyanin was determined
based on Wrolstad et al. (2005) using the following
equation:
Anthocyanin concentration mM A  DF  103 =e  l;
where DF is the dilution factor, l is the light-path length
(1 cm), e is the molar extinction coecient (L mol1 cm
1
) which, in this case, equals to 26,900. A is the absorbance, given by:
A Amax  A700 pH

1:0

 Amax  A700 pH

4:5 ;

where Amax and A700 are the absorbance at the maximum


and 700 nm measured at pH 1.0 or pH 4.5, respectively.
The absorbance was measured by using a UVVIS spectrophotometer (U-3210, Hitachi).
2.4. Anthocyanin purication
In some experiments, further puried anthocyanin was
used. Purication was done as described by Ordaz-Galindo
et al. (1999) using the BAKERBONDe spe Octadecyl disposable extraction column (J.T. Baker). The water-based
anthocyanin extract (1 mL) was introduced to the column,
which had been activated rst with 1 mL of methanol and
then 1 mL of 0.01% HCl (v/v). The column was then rinsed
with 2 mL of 0.01% HCl (v/v) to remove sugars, acids, and
other water-soluble compounds (designated as impurities).
Anthocyanin was then collected by eluting the column with
methanol containing 0.01% HCl (v/v). To replace the solvent methanol with water, the eluate was dried at 4 C
and the powder was then dissolved in water.
2.5. Dye coating and assembly of the solar cell
Normally, dye coating was done by soaking TiO2 glass
electrodes in the anthocyanin extract (3 mM, pH 8.0) at
4 C in the dark for 24 h, and dried once with a tissue paper
before drying again in a vacuum desiccator for at least 1 h.
In some experiments, various concentrations (3 lM6 mM)
or pH (212) of the anthocyanin extract, or immersion time
down to 1 s were employed. The pH of the anthocyanin
extract was adjusted by 3 M NaOH and 3 M HCl, and
monitored by a pH meter (SP-701, Suntex). In other experiments, 40 mM deoxycholic acid (DCA) was mixed with
various concentrations of anthocyanin (0.12 mM) when
fabricating sensitized TiO2 lms.
Assembly of the solar cell was done rst by placing an
adhesive tape (ca. 40 lm in thickness) as a spacer layer
and sealing element which dened a 10  10 mm2 aperture
on the conductive side of the cathode. The cathode was
then placed on top so that its conductive side faced the
stained TiO2 lm of the anode. The two electrodes were
held together by binding clips, and the space between them
was lled with liquid electrolyte (0.05 M iodine, 0.5 M

C.-Y. Chien, B.-D. Hsu / Solar Energy 98 (2013) 203211

lithium iodide and 0.5 M 4-tert-butylpreidine (TBP) in 3methoxypropionitrile (MPN)) by capillary action. Any
excess electrolyte was removed from the edge of the solar
cell with a tissue paper. The assembled solar cells were kept
in the dark for 10 min before measurement.
2.6. Characterization and measurement
The currentvoltage (IV) curve of the solar cell was
measured by using a computer-controlled digital source
meter (Keithley 2400) under one sunlight (AM1.5G), which
was simulated by a solar simulator (Oriel Sol3A, Newport)
and calibrated to an intensity of 1 kW m2 by using a silicon reference photodiode. The IV curves were recorded
by sweeping between 0 and 0.8 V at a 0.05 V s1 sweep
rate. The open-circuit voltage (VOC) and short-circuit photocurrent (JSC), given by the IV curve, were used to calculate the light-to-electricity conversion eciency (g) with the
formula:
g V OC  J SC  FF=P in ;
where Pin is the power of incident light per unit area, and
the ll factor (FF) was given by:
FF V max  J max =V OC  J SC ;
where Vmax and Jmax were the maximum voltage and photocurrent measured while the cell was working.
The monochromatic incident photon-to-electron conversion eciency (IPCE) was measured by an IPCE measurement device (model QEX7, PV Measurements, Inc.).
The absorption spectrum of anthocyanin was measured
by using a UVVIS spectrophotometer (Hitachi, U-3210).
3. Results and discussion
3.1. Eects of pH
The aglycone structure of anthocyanin is shown in
Fig. 1a. The major anthocyanin of red cabbage is cyanidin-3-O-diglucoside-5-O-glucoside, in which the R1 position of the aglycone is a hydroxyl group and R2 is a
hydrogen atom (McDougall et al., 2007). In aqueous solutions, the aglycones of anthocyanins can exist as ve molecular species in chemical equilibrium: red avylium cation,
colorless carbinol pseudo base, purple quinoidal base, blue
quinoidal anion, and yellowish chalcone (Brouillard and
Dangles, 1993). Both the equilibrium and its color shift
depending on the pH of the solution (Fig. 1b). The absorption spectra of the anthocyanin extract at various pH values are shown in Fig. 1c. At acidic pH (pH 2.0),
avylium cation is the dominant form and has a strong
absorption at 520 nm, displaying a red color. As the pH
rises, the avylium cation gradually converts to quinoidal
base by losing a proton. The absorption becomes much
weaker and shifts to 540560 nm, yielding a magenta (pH
4.0) or a purple (pH 6.0) color. At neutral pH, the blue quinoidal anion forms by losing another proton, and it has a

205

stronger and red-shifted absorption at about 600 nm.


Under basic conditions (pH > 8.0), the avylium cation
will dehydrate to form colorless carbinol pseudo base and
then chalcone. The chalcone is yellow in color, because
of its absorption in the UV and violet-blue regions. The
green color of the anthocyanin solution at pH 10.0 is a
result of a mixture of blue quinoidal anion and yellow
chalcone.
The performance of DSSC was investigated in relation
to the pH values of the anthocyanin extract, in which the
TiO2 nanocrystalline lm was soaked when fabricating
DSSC. The results are listed in Table 1. Under our experimental conditions, the g reached maximum at pH 8.0
and decreased toward both acidic and basic ends. The
trend of JSC was similar. The VOC increased as conditions
change from acidic to basic. There was no signicant dierence in FF.
The continuous increase of VOC with increasing pH of
our anthocyanin-based DSSC is likely to be due to the interface at the surface of the TiO2 lm being charged more negatively by bound anthocyanin molecules at higher pH, and
thus shifts the conduction band edge and the quasi-Fermi
level upwards. As for JSC, its low value at acidic pH may
be attributed to the decreased amount of dye molecules
bound to the TiO2 surface. The binding of anthocyanin is
through the hydroxyl group in its B ring, which undergoes
deprotonation before the formation of the bond (Cherepy
et al., 1997; Dai and Rabani, 2002). A low pH may suppress
the deprotonation of avylium cations, causing poor binding of anthocyanin (Cherepy et al., 1997). The bound avylium cations may also prevent the binding of other dyes due
to charge repulsion. In addition, a positive charge on anthocyanin molecule may impede electron injection to the conduction band of TiO2, and increase the charge
recombination probability. The lower JSC at basic pH
may be explained by two dierent factors, one is due to
the decreasing driving force for electron injection caused
by the upward shift of the conduction band edge of TiO2
with increasing pH. It may also be attributed to the fact that
the concentration of the chalcone form increases at high
pH. Chalcone absorbs poorly in the peak region of solar
radiation spectrum in compared with the quinoidal from
(see Fig. 1c), and thus contributes less to the photocurrent.
It is worth noting that the VOC of the DSSC employing
anthocyanin (530620 mV) is lower than that of ruthenium-based sensitizers (650850 mV). This is probably
due to a higher rate of charge recombination at the interface when this natural dye was used. It is pointed out that
small and at organic molecules like anthocyanin may not
be able to insulate the underlying TiO2 layer from the oxidized electrolyte well, resulting in an increased dark current
(Burke et al., 2008).
3.2. Eects of anthocyanin concentration
With two hydroxyl groups on the B ring, cyanidin possesses good ability to bind TiO2 (Cherepy et al., 1997). It

206

C.-Y. Chien, B.-D. Hsu / Solar Energy 98 (2013) 203211

Fig. 1. (a) The aglycone structure of anthocyanin. (b) The structure of main anthocyanin forms and their pH-dependent equilibria. (c) The UVVIS
absorption spectra of the anthocyanin extract at various pH values.

has been shown that the stability of the surface complexes


of titanium (IV) with organic ligands are determined by

various factors, including the type of TiO2 crystal plane,


accessibility of titanium ions, the type of ligating sites of

C.-Y. Chien, B.-D. Hsu / Solar Energy 98 (2013) 203211

Table 1
The eects of the pH of the anthocyanin extract on the performance of
DSSC.
pH

VOC (mV)

JSC (mA/cm2)

FF

g (%)

2.0
4.0
6.0
8.0
10.0
12.0

533 20
561 25
571 16
592 18
612 18
624 26

1.449 0.089
1.975 0.104
2.367 0.172
2.756 0.174
1.943 0.169
1.127 0.087

0.711 0.002
0.702 0.005
0.708 0.005
0.718 0.011
0.718 0.026
0.719 0.012

0.555 0.016
0.786 0.056
0.967 0.028
1.184 0.051
0.863 0.076
0.511 0.061

ligands and pH. Several dierent structures of surface complexes are described such as monodentate, bidentate (chelating) and bidentate (bridging) (Macyk et al., 2010). It is
likely that the binding of anthocyanin to the TiO2 surface
may involve dierent structures with dierent anities.
The performance of DSSC was studied in relation to the
concentration of the anthocyanin extract, in which the
TiO2 lm was soaked when fabricating DSSC. The concentration was varied between 3 lM and 6 mM at constant pH
of 8.0. As shown in Table 2, all four parameters in general
increased with increasing anthocyanin concentration, but
the trend was more pronounced in JSC and g. The best performance was reached at 3 mM, and higher concentrations
did not yield better results. With increasing concentration
of anthocyanin, the number of bound dye molecules also
increases, so does the photocurrent. In the meantime, the
increase of bound anthocyanins would increase their surface coverage of TiO2 nanoparticles, which helps to insulate the TiO2 from electrolyte, reduces the dark current
and raises VOC.
It is worth noting that a 300-fold increase in anthocyanin concentration from 10 lM to 3 mM improved the eciency by only about twofold, while the increases were
roughly in proportion under 10 lM. The result indicates
that TiO2 nanoparticles possess very high anity binding
sites with a micromolar dissociation constant for anthocyanin molecules. At a concentration above 10 lM, saturation of these binding sites will force anthocyanins to bind
to the sites with lower anity. The resulting chemical
bonds have less stability and contribute less to the eciency of solar cells. It has been shown that, besides the
hydroxyl groups in the B ring of anthocyanin molecules,
the hydroxyl groups on sugar moieties can also interact

207

with TiO2 (Shkrob et al., 2004). Little photocurrent is


expected to generate from this kind of interaction. The performance of cells with anthocyanin concentration higher
than 3 mM was slightly lowered. This might be caused by
the self-aggregation of anthocyanin molecules on TiO2 surfaces (Leydet et al., 2012).

3.3. Eects of immersion time


In the majority of previous studies, the electrodes were
immersed in the anthocyanin extract for at least 16 h to
ensure TiO2 particles were completely covered with anthocyanin molecules. The eects of the immersion time from
1 s to 16 h at constant pH (8.0) and anthocyanin concentration (3 mM) were investigated. As shown in Table 3, the
VOC and FF of longer immersion time were slightly higher.
On the other hand, JSC and g steadily increased with
immersion time and reached the maximum at 15 min.
Longer immersion time resulted in drops in both parameters, which then became stabilized between 1 and 16 h.
The process of coating TiO2 nanoparticles with anthocyanin involves the diusion of anthocyanin molecules into
the porous TiO2 lm and the ensuing binding reaction of
the two. Anthocyanin is a small molecule with molecular
weight less than 1 kDa. It is estimated that anthocyanin
molecules can diuse over a distance of the thickness of
the TiO2 lm (15 lm) in 1 s (diusion coecient  3  10
6
cm2 s1). The binding reaction of anthocyanin is
reported to occur through a loss of water and the condensation of two hydroxyl groups respectively from the B ring
Table 3
The eects of the time of TiO2 lm immersed in the anthocyanin extract on
the performance of DSSC.
Time

VOC (mV)

JSC (mA/cm2)

FF

g (%)

1s
10 s
30 s
1 min
5 min
10 min
15 min
30 min
1h
16 h

530 11
551 9
555 7
548 20
546 5
552 4
570 19
555 15
556 5
572 18

2.231 0.306
2.572 0.215
2.704 0.135
2.784 0.264
3.054 0.456
3.094 0.319
3.247 0.410
3.061 0.409
3.066 0.783
3.027 0.362

0.640 0.031
0.618 0.018
0.651 0.017
0.684 0.027
0.684 0.028
0.695 0.010
0.666 0.048
0.697 0.026
0.678 0.048
0.671 0.027

0.757 0.042
0.876 0.058
0.977 0.062
1.044 0.044
1.141 0.124
1.187 0.054
1.233 0.076
1.184 0.099
1.156 0.163
1.162 0.045

Table 2
The eects of the concentration of the anthocyanin extract on the performance of DSSC.
Concentration

VOC (mV)

JSC (mA/cm2)

FF

g (%)

6.0 mM
3.0 mM
1.0 mM
300 lM
100 lM
30 lM
10 lM
3 lM

550 4
570 8
547 5
545 7
537 5
520 0
534 11
506 49

2.928 0.319
3.305 0.565
3.017 0.221
2.805 0.590
2.580 0.226
2.288 0.109
1.940 0.067
0.887 0.280

0.709 0.005
0.710 0.006
0.702 0.006
0.689 0.007
0.684 0.007
0.678 0.009
0.685 0.016
0.643 0.009

1.163 0.032
1.337 0.045
1.159 0.089
1.053 0.014
0.948 0.070
0.807 0.181
0.710 0.127
0.289 0.164

208

C.-Y. Chien, B.-D. Hsu / Solar Energy 98 (2013) 203211

of anthocyanin and the surface of TiO2 nanostructure. This


is considered to be a rapid reaction (Cherepy et al., 1997).
In fact, the solar cells produced with an immersion time of
1 min could achieve an eciency that is 85% of the maximum (Table 3). The results clearly show that the adsorption of anthocyanin by TiO2 is faster than many other
pigments.
3.4. Eects of coadsorbent (DCA)
Anthocyanin is a planar molecule with aromatic rings
(see Fig. 1a). They are prone to pp stacking interactions
with themselves (self-association), organizing into aggregations from simple to complex (Hoshino, 1991; Houbiers

Fig. 2. (a) A proposed structure of two self-associated cyanidin molecules


which are stacked in a counterclockwise screw manner. (b) Higher order
anthocyanin aggregation (shown as orange sticks) and monomeric
molecules (pink sticks) are attached to the TiO2 surface. (c) DCA (thick
grey sticks) can function as spacers that break up the aggregation of
anthocyanin molecules (blue sticks). (For interpretation of the references
to colour in this gure legend, the reader is referred to the web version of
this article.)

et al., 1998; Leydet et al., 2012). Fig. 2a shows a proposed


structure of two self-associated cyanidin molecules which
are stacked in a counterclockwise screw manner (Hoshino,
1991). Although aggregation of dye molecules improves
light harvesting and stability (Houbiers et al., 1998; Wang
et al., 2007), many studies have shown that the aggregation
is a troublesome problem to DSSC performance due to
inecient electron injection and intermolecular energy
quenching (Ehret et al., 2001; Pastore and Angelis, 2010;
Tatay et al., 2007). In addition, higher order dye aggregation attached to the TiO2 surface, as shown in Fig. 2b,
may shade the monomeric molecules nearby and prevent
the access to the later (in oxidized form) from electrolyte.
One remedy is to introduce coadsorbents that can act as
the spacers between dye molecules (Hara et al., 2004; Lim
et al., 2011; Sharma et al., 2012; Wang et al., 2007; Xia and
Yanagida, 2011). Deoxycholic acid (DCA), a planar,
amphipathic molecule with a steroid backbone, is believed
to be able to break up the aggregation of anthocyanin molecules (see Fig. 2c), thus minimizing the intermolecular
quenching and making the later a more ecient electron
injector. In this study, 40 mM DCA was mixed with various concentrations of anthocyanin when fabricating sensitized TiO2 lms. As shown in Table 4, DCA indeed
improved the performance of DSSC. The highest g was
raised to 1.426% from 1.130% with anthocyanin at a concentration of 1 mM (molar ratio of dye:coadsorbent = 1:40), an augment of 26%. The largest extent of
improvement occurred at 0.1 mM anthocyanin (molar
ratio = 1:400), reaching 79%. There were parallel changes
in JSC and g. There was no change in VOC upon the addition of DCA at the concentration of anthocyanin of 2 mM,
but the increases in VOC at lower concentrations were signicant. The FF remained virtually unchanged whether
DCA was added or not. The increase in photocurrent could
be explained by the dispersion of anthocyanin aggregation
by DCA, leading to the suppression of quenching process
between molecules. It could also due to the adsorption of
DCA on the TiO2 surface, which would positively shift
the conduction band edge, resulting in a larger driving

Table 4
The eects of DCA on the performance of DSSC.
Anthocyanin concentration

VOC (mV)

JSC (mA/cm2)

FF

g (%)

2 mM
2 mM + DCAa
1 mM
1 mM + DCA
0.6 mM
0.6 mM + DCA
0.3 mM
0.3 mM + DCA
0.2 mM
0.2 mM + DCA
0.1 mM
0.1 mM + DCA

590 9
590 15
595 10
624 33
584 5
606 23
562 2
578 17
546 8
591 26
538 18
566 15

2.932 0.325
3.472 0.078
2.775 0.472
3.159 0.001
2.378 0.474
2.925 0.195
2.184 0.498
3.083 0.256
1.718 0.211
2.638 0.152
1.522 0.122
2.591 0.315

0.693 0.031
0.693 0.012
0.684 0.063
0.724 0.050
0.710 0.042
0.706 0.029
0.690 0.003
0.666 0.022
0.673 0.034
0.697 0.058
0.675 0.024
0.673 0.008

1.199 0.042
1.420 0.027
1.130 0.054
1.426 0.079
0.986 0.168
1.251 0.075
0.846 0.148
1.188 0.061
0.631 0.011
1.086 0.086
0.552 0.126
0.988 0.094

The concentration of DCA was 40 mM.

C.-Y. Chien, B.-D. Hsu / Solar Energy 98 (2013) 203211

209

force for electron ejection from the excited anthocyanin


molecules (Kay and Gratzel, 1993). The improvement of
VOC with addition of DCA could be explained by the
reduction of dark current due to the shielding provided
by DCA on the unoccupied Ti (IV) (Sharma et al., 2012).
This was especially true for the cells that were fabricated
using lower concentrations of anthocyanin. The increase
due to this eect could compensate for the loss due to positive shift of the conduction band edge.
It was found that a high molar ratio of dye:coadsorbent
was required for the best performance of DSSC. Similar
phenomenon has been observed in many previous studies
(Kay and Gratzel, 1993; Nuay et al., 2009; Ono et al.,
2010). As mentioned earlier, anthocyanins are small and
at organic molecules. They are not able to insulate the
underlying TiO2 layer from the oxidized electrolyte well
(Burke et al., 2008), and thus a large quantity of DCA
was needed to enlarge the surface coverage. In addition,
DCA is an ionic detergent. A large amount is required to
dissolve aggregated anthocyanin when fabricating the sensitized TiO2 anodes.
However, it is worthy to note that the color of sensitized
TiO2 lm became lighter when DCA was introduced. This
was because that DCA, as a coadsorbent, also occupied
binding sites on TiO2, resulting in a decrease in the amount
of anthocyanin that could be adsorbed on the TiO2 lm.

The monochromatic incident photon-to-electron conversion eciency (IPCE) spectra of DSSC sensitized with
anthocyanin and anthocyanin plus DCA are shown in
Fig. 3. The enhancement eect of DCA was found in all
the visible region of the spectrum.
In the IPCE spectra, the peak at 350 nm was from TiO2,
whereas that at 610 nm was attributed to anthocyanin. The
peak at 420 nm, which could be from the copigments such
as avonoids and polyphenols, had an even higher conversion eciency (16%) compared to the peak at 610 nm of
anthocyanin (9%).

Fig. 3. The IPCE spectra of DSSCs sensitized with anthocyanin and


anthocyanin plus DCA as coadsorbant. The anthocyanin concentration
was 0.6 mM. The molar ratio of dye:coadsorbant = 1:66.7.

Fig. 4. The UVVIS absorption spectra of the PA and UPA solutions at


pH 8.0. The spectrum in the lower panel represents the dierence between
the upper two spectra.

3.5. Eects of purication


The anthocyanin extract was further puried by octadecyl extraction columns. Table 5 shows the characteristics
of DSSC sensitized with unpuried (UPA) and puried
anthocyanin (PA), which were adjusted to 2 mM before
fabrication. Unexpectedly, the performance of the DSSC
with PA was worse than that with UPA in all parameters.
If the impurities (the fraction of the eluate that left the

Table 5
The eects of anthocyanin purication on the performance of DSSC.

UPA
PA
PA + impurities
UPA + H2O
PA + H2O

VOC (mV)

JSC (mA/cm2)

FF

g (%)

590 17
517 31
585 7
585 7
550 10

2.696 0.017
1.266 0.126
2.401 0.022
2.474 0.149
1.610 0.043

0.683 0.003
0.646 0.033
0.701 0.004
0.685 0.012
0.672 0.011

1.098 0.029
0.422 0.015
0.994 0.016
1.003 0.089
0.602 0.017

210

C.-Y. Chien, B.-D. Hsu / Solar Energy 98 (2013) 203211

column before the PA fraction) was added back to the PA


solution before constructing sensitized TiO2 lms (nal
anthocyanin concentration 0.6 mM), DSSC regained most
of its original performance, clearly indicating that there
were yet unknown components in the crude extract that
could help anthocyanin molecules in the conversion of light
into electricity. There was a slight improvement when the
same amount of water as the impurities was added to the
PA solution, which is probably due to the dilution of the
dye.
Fig. 4 shows the UVVIS absorption spectra of the PA
and UPA solutions at pH 8.0, and the dierence spectrum
between the two (lower panel). The impurities removed by
column chromatography showed a major peak around
380 nm. This could correspond to a peak at 420 nm in
the IPCE spectra shown in Fig. 3, arising from the copigments associated with anthocyanin. The shift of the wavelength of the peak from 380 to 420 nm might be caused by
the binding with TiO2. The removal of the impurities,
which also contributed signicantly to the process of photon to electron conversion as shown in Fig. 3, would surely
bring down the performance of DSSC. It is worth to note
that the common copigments are avonols like quercetin
and kaempferol, which show absorption peaks in the same
spectral range (300400 nm).
4. Conclusion
Anthocyanin from red cabbage has a high anity for
TiO2 nanoparticles, with binding taking place at low
anthocyanin concentrations and in relatively short time.
The DSSC sensitized with this natural dye shows a better
performance if fabrication conditions such as the pH and
concentration of anthocyanin extract, immersion time of
TiO2 electrode, as well as the molar ratio of dye/coadsorbent (deoxycholic acid) are optimized. The highest lightto-electricity conversion eciency reached was over 1.4%
under standard AM 1.5 solar conditions, almost three
times the previously reported conversion eciencies
(0.50.6%). In addition, crude extract performs better than
further puried anthocyanin. This is attributed to the presence of natural copigments in the crude extract.
With the advantages such as ready availability, quick
and easy preparation procedures, environmental friendliness and reasonable high eciency, the DSSC sensitized
with natural anthocyanin certainly deserves further
exploration.
References
Aduloju, K.A., Shitta, M.B., 2012. Dye sensitized solar cell using natural
dyes extracted from red leave onion. Int. J. Phys. Sci. 7, 709712.
Brouillard, R., 1983. The in vivo expression of anthocyanin colour in
plants. Phytochemistry 22, 13111323.
Brouillard, R., Dangles, O., 1993. Flavonoids and ower colour. In:
Harborne, J.B. (Ed.), The Flavonoids: Advances in Research Since
1986. Chapman and Hall, London, pp. 565588.

Burke, A., Ito, S., Snaith, H., Bach, U., Kwiatkowski, J., Gratzel, M.,
2008. The function of a TiO2 compact layer in dye-sensitized solar cells
incorporating planar organic dyes. Nano Lett. 8, 977981.
Calogero, G., Di Marco, G., 2008. Red Sicilian orange and purple
eggplant fruits as natural sensitizers for dye-sensitized solar cells. Sol.
Energy Mater. Sol. C 92, 13411346.
Calogero, G., Yum, J.H., Sinopoli, A., Di Marco, G., Gratzel, M.,
Nazeeruddin, M.K., 2012. Anthocyanins and betalains as lightharvesting pigments for dye-sensitized solar cells. Sol. Energy 86,
15631575.
Chang, H., Lo, Y.J., 2010. Pomegranate leaves and mulberry fruit as
natural sensitizers for dye-sensitized solar cells. Sol. Energy 84, 1833
1837.
Cherepy, N.J., Smestad, G.P., Gratzel, M., Zhang, J.Z., 1997. Ultrafast
electron injection: Implications for a photoelectrochemical cell utilizing an anthocyanin dye-sensitized TiO2 nanocrystalline electrode. J.
Phys. Chem. B 101, 93429351.
Dai, Q., Rabani, J., 2002. Photosensitization of nanocrystalline TiO2 lms
by anthocyanin dyes. J. Photochem. Photobiol. A 148, 1724.
Ehret, A., Stuhl, L., Spitler, M.T., 2001. Spectral sensitization of TiO2
nanocrytalline electrodes with aggregated cyanine dyes. J. Phys. Chem.
B 105, 99609965.
Furukawa, S., Iino, H., Iwamoto, T., Kukita, K., Yamauchi, S., 2009.
Characteristics of dye-sensitized solar cells using natural dye. Thin
Solid Films 518, 526529.
Gaiddon, C., Jeannequin, P., Bischo, P., Pfeer, M., Sirlin, C., Loeer,
J.P., 2005. Ruthenium (II)-derived organometallic compounds induce
cytostatic and cytotoxic eects on mammalian cancer cell lines through
p53-dependent and p53-independent mechanisms. J. Pharmacol. Exp.
Ther. 315, 14031411.
Hao, S., Wu, J., Huang, Y., Lin, J., 2006. Natural dyes as photosensitizers
for dye-sensitized solar cell. Sol. Energy 80, 209214.
Hara, K., Dan-oh, Y., Kasada, C., Ohga, Y., Shinpo, A., Suga, S.,
Sayama, K., Arakawa, H., 2004. Eect of additives on the photovoltaic performance of coumarin-dye-sensitized nanocrystalline TiO2
solar cells. Langmuir 20, 42054210.
Hoshino, T., 1991. An approximate estimate of self-association constants
and the self-stacking conformation of malvin quinonoidal bases
studied by 1H NMR. Phytochemistry 30, 20492055.
Houbiers, C., Lima, J.C., Macanita, A.L., Santos, H., 1998. J. Phys.
Chem. B 102, 35783585.
Jackman, R.L., Yada, R.Y., Tung, M.A., Speers, R.A., 1987. Anthocyanins as food colorants a review. J. Food Biochem. 11, 201247.
Kalowekamo, J., Baker, E., 2009. Estimating the manufacturing cost of
purely organic solar cells. Sol. Energy 83, 12241231.
Kay, A., Gratzel, M., 1993. Articial photosynthesis. 1. Photosensitization of TiO2 solar cells with chlorophyll derivatives and related natural
porphyrins. J. Phys. Chem. 97, 62726277.
Leydet, Y., Gavara, R., Petrov, V., Diniz, A.M., Parola, A.J., Lima, J.C.,
Pina, F., 2012. The eect of self-aggregation on the determination of
the kinetic and thermodynamic constants of the network of chemical
reactions in 3-glucoside anthocyanins. Phytochemistry 83, 125135.
Lila, M.A., 2004. Anthocyanins and human health: an in vitro investigative approach. J. Biomed. Biotechnol. 5, 306313.
Lim, J., Kwon, Y.S., Park, T., 2011. Eect of coadsorbent properties on
the photovoltaic performance of dye-sensitized solar cells. Chem.
Commun. 47, 41474149.
Luo, P., Niu, H., Zheng, G., Bai, X., Zhang, M., Wang, W., 2009. From
salmon pink to blue natural sensitizers for solar cells: Canna indica L.,
Salvia splendens, cowberry and Solanum nigrum L. Spectrochim. Acta
A 74, 936942.
Macyk, W., Szaciowski, K., Stochel, G., Buchalska, M., Kuncewicz, J.,
abuz, P., 2010. Titanium (IV) complexes as direct TiO2 photosensitizers. Coord. Chem. Rev. 254, 26872701.
McDougall, G.J., Fye, S., Dobson, P., Stewart, D., 2007. Anthocyanins
from red cabbage stability to simulated gastrointestinal digestion.
Phytochemistry 68, 12851294.

C.-Y. Chien, B.-D. Hsu / Solar Energy 98 (2013) 203211


Nazeeruddin, M.K., Pechy, P., Renouard, T., Zakeeruddin, S.M.,
Humphry-Baker, R., Comte, P., Liska, P., Cevey, L., Costa, E.,
Shklover, V., Spiccia, L., Deacon, G.B., Bignozzi, C.A., Gratzel, M.,
2001. Engineering of ecient panchromatic sensitizers for nanocrystalline TiO2-based solar cells. J. Am. Chem. Soc. 123, 16131624.
Nuay, V.A., Kim, D.H., Lee, S.H., Ko, J., 2009. Thiophene linked
porphyrin derivatives for dye sensitized solar cell. Bull. Korean Chem.
Soc. 30, 28712872.
Ono, T., Yamaguchi, T., Arakawa, H., 2010. Inuence of dye adsorption
solvent on the performance of a mesoporous TiO2 dye-sensitized solar
cell using infrared organic dye. J. Sol. Energy Eng. 132, 021101-1
021101-7.
Ordaz-Galindo, A., Wesche-Ebeling, P., Wrolstad, R.E., RodriguezSaona, L., Argaiz-Jamet, A., 1999. Purication and identication of
Capulin (Prunus serotina Ehrh) anthocyanins. Food Chem. 65, 201
206.
ORegan, B., Gratzel, M., 1991. A low-cost, high-eciency solar cell based
on dye-sensitized colloidal TiO2 lms. Nature 353, 737740.
Pastore, M., Angelis, F.D., 2010. Aggregation of organic dyes on TiO2 in
dye-sensitized solar cells models: an ab initio investigation. ACS Nano
4, 556562.
Polo, A.S., Murakami Iha, N.Y., 2006. Blue sensitizers for solar cells:
natural dyes from Calafate and Jaboticaba. Sol. Energy Mater. Sol. C
90, 19361944.
Qin, Y., Peng, Q., 2012. Ruthenium sensitizers and their applications in
dye-sensitized solar cells. Int. J. Photoenergy, <http://dx.doi.org/
10.1155/2012/291579>.

211

Sharma, G.D., Kurchania, R., Ball, R.J., Roy, M.S., Mikroyannidis, J.A.,
2012. Eect of deoxycholic acid on the performance of liquid
electrolyte dye-sensitized solar cells using a perylene monoimide
derivative. Int. J. Photoenergy, <http://dx.doi.org/10.1155/2012/
983081>.
Shkrob, I.A., Sauer Jr., M.C., Gosztola, D., 2004. Ecient, rapid oneelectron photooxidation of chemisorbed polyhydroxyl alcohols and
carbohydrates by TiO2 nanoparticles in an aqueous solution. J. Phys.
Chem. B 108, 1251212517.
Tatay, S., Haque, S.A., ORegan, B., Durrant, J.R., Verhees, W.J.H.,
Kroon, J.M., Vidal-Ferran, A., Gavina, P., Palomares, E., 2007.
Kinetic competition in liquid electrolyte and solid-state cyanine dye
sensitized solar cells. J. Mater. Chem. 17, 30373044.
Wang, Z.S., Cui, Y., Dan-oh, Y., Kasada, C., Shinpo, A., Hara, K., 2007.
Thiophene-functionalized coumarin dye for ecient dye-sensitized
solar cells: electron lifetime improved by coadsorption of deoxycholic
acid. J. Phys. Chem. C 111, 72247230.
Wongcharee, K., Meeyoo, V., Chavadej, S., 2007. Dye-sensitized solar cell
using natural dyes extracted from rosella and blue pea owers. Sol.
Energy Mater. Sol. C 91, 566571.
Wrolstad, R.E., Durst, R.W., Lee, J., 2005. Tracking color and pigment
changes in anthocyanin products. Trends Food Sci. Technol. 16, 423428.
Xia, J., Yanagida, S., 2011. Strategy to improve the performance of dyesensitized solar cells: interface engineering principle. Sol. Energy 85,
31433159.
Yasbin, R.E., Matthews, C.R., Clarke, M.J., 1980. Mutagenic and toxic
eects of ruthenium. Chem.-Biol. Interact. 31, 355365.

Вам также может понравиться