Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 96 (2012) 305–309

Contents lists available at SciVerse ScienceDirect

Spectrochimica Acta Part A: Molecular and


Biomolecular Spectroscopy
journal homepage: www.elsevier.com/locate/saa

Performance of Kerria japonica and Rosa chinensis flower dyes as sensitizers


for dye-sensitized solar cells
K.V. Hemalatha, S.N. Karthick, C. Justin Raj, N.-Y. Hong, S.-K. Kim, H.-J. Kim ⇑
School of Electrical Engineering, Pusan National University, Jangjeon, Geumjeong, Busan 609-735, Republic of Korea

h i g h l i g h t s g r a p h i c a l a b s t r a c t

" The efficiency of Kerria japonica and


Rosa chinensis dye is 0.22%, 0.29%
respectively.
" In the presence of sugar molecule, it
was increased to 0.29% for K. japonica
and decreased to 0.27% for
R. chinensis.
" Implementation of natural dye in
DSSC is cost effective compared to the
metal based dyes.

a r t i c l e i n f o a b s t r a c t

Article history: The natural dyes carotenoid and anthocyanin were extracted from Kerria japonica and Rosa chinensis,
Received 26 December 2011 respectively, using a simple extraction technique without any further purification. They were then used
Received in revised form 23 April 2012 as sensitizers in dye-sensitized solar cells (DSSCs), and their characteristics were studied. The ranges of
Accepted 10 May 2012
short-circuit current (JSC) from 0.559 to 0.801 (mA/cm2), open-circuit voltage (VOC) from 0.537 to 0.584 V,
Available online 24 May 2012
and fill factor from 0.676 to 0.705 were obtained for the DSSCs made using the extracted dyes. Sugar mol-
ecules were added externally to the dye for stabilization and to increase the conversion efficiency. The
Keywords:
efficiencies of the K. japonica and R. chinensis dyes were 0.22% and 0.29%, respectively; after the addition
Anthocyanin
Carotenoid
of sugar, the efficiency increased to 0.29% for K. japonica and decreased to 0.27% for R. chinensis. Thus, the
Dye-sensitized solar cell addition of sugar molecules increased the conversion efficiency slightly with the carotenoid dye of K.
Photocurrent japonica, while there was no considerable change with the anthocyanin of R. chinensis. This paper briefly
discusses the simple extraction technique of these natural dyes and their performance in DSSCs.
Ó 2012 Elsevier B.V. All rights reserved.

Introduction employed to sensitize nanocrystalline TiO2 semiconductors is a


transition-metal coordination compound (ruthenium polypyridyl
Dye-sensitized solar cells (DSSCs) are considered as alternative complex). The ruthenium polypyridyl complexes anchored to
energy sources to conventional inorganic semiconductor photovol- nanocrystalline TiO2 surfaces through carboxylic acid groups [1]
taic devices because of the low fabrication costs involved. Dyes are the most successful sensitizers for applications in regenerative
play key roles in light harvesting and in the conversion of solar en- solar cells. However, the high cost and long-term unavailability of
ergy into electrical energy, and the dye molecules in DSSCs ‘‘sensi- Ru complexes have led to a widening of research on low-cost, effi-
tize’’ wide-bandgap semiconductors to visible radiation. Therefore, cient sensitizers. Consequently, metal-free sensitizers such as nat-
dyes are referred to as ‘‘sensitizers’’. The most efficient sensitizer ural dyes have been investigated as alternative molecules for DSSC
applications [2]. The advantages of natural pigments as photosen-
⇑ Corresponding author. Tel.: +82 51 510 2364; fax: +82 51 513 0212. sitizers are their intramolecular p–p⁄ transitions, the absence of
E-mail address: heeje@pusan.ac.kr (H.-J. Kim). noble metals such as Ru, and their large absorption coefficients,

1386-1425/$ - see front matter Ó 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.saa.2012.05.027
306 K.V. Hemalatha et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 96 (2012) 305–309

high light-harvesting efficiencies, easy preparation, environmen- temperature for adequate extraction without exposure to sunlight.
tally friendly nature, high availability, and low cost [3]. After the extraction, the solid residues were filtered out, and the
In our present investigation, DSSCs were assembled using clear solutions were used as prepared, without purification. Fur-
natural dyes extracted from the flowers of Kerria japonica and Rosa ther purification of the extracts was avoided so as to achieve effi-
chinensis as sensitizers. The performances of the raw natural dyes cient sensitization using simple extraction procedures. The
are better than those of the commercial or purified analogs because extractants were properly stored, protected from direct sunlight,
of the presence of natural extracts like alcohols, organic acids, etc., and used further as sensitizers in DSSCs.
which act as a co-adsorbates, suppressing recombination with the
electrolyte, reducing dye aggregation, and favoring charge Solar-cell assembly
injection [4]. Therefore, the dyes were extracted using a minimal
chemical procedure without any purification. The pigment A fluorine-doped tin oxide (FTO) glass plate (13 O/sq), pur-
carotenoid is responsible for the color of the flower of K. japonica. chased from Hartford Glass Co., Inc., was used as the current collec-
The total percentage of carotenoid in the form of carotene and tor. The FTO plate was first cleaned using an ultrasonic bath with
xanthophyll content present in K. japonica is 88% [5]. All carote- acetone, ethanol, and water for about 10 min each. Scotch tape
noids are polyisoprenoids, possessing an extensive system of was used as a spacer to control the film thickness and to provide
conjugated double bonds, as shown in Fig. 1(a). This serves as non-coated areas for electrical contact. The TiO2 paste (Ti-nanoxide
the light-absorbing chromophore responsible for the yellow, T/SP, Solaronix) was coated on the FTO plate using the doctor-
orange, and red colors of the plant sources, and provides the visible blade method. After being coated, the TiO2 films were air-dried
absorption spectrum that is the basis for their identification. The for about 5 min to reduce the surface irregularities. The films were
long system of alternating double and single bonds giving this annealed at 450 °C in air for 30 min to remove the organic loads
distinct light-absorbing property is the most important feature of and facilitate the interconnection of the TiO2 nanoparticles. The
the carotenoids [6]. In the flower of R. chinensis, the anthocyanin thickness of the films was 20 lm, and the active area of the
molecule is the core component, which shows color in the range TiO2 electrodes used was 0.27 cm2. The TiO2 film electrodes were
of visible light from red to blue, so it is considered as a valuable cooled to 80 °C, and were then immersed in an extracted caroten-
natural material to be used as an efficient sensitizer for wide-band- oid and anthocyanin dye solution separately for 24 h. After the dye
gap semiconductors. The presence of carbonyl and hydroxyl groups adsorption was completed, the film was cleaned with pure ethanol
in the anthocyanin molecule (Fig. 1b) bound to the TiO2 helps the and dried with hot air. The Pt electrode was prepared using plati-
photoelectric conversion efficiency. The extracted dyes were trea- num paste (platisol T/SP, Solaronix). The paste was coated on the
ted with sugar molecules, because, according to Filipa Queiroz FTO glass plate using the doctor-blade technique and heated at
et al., 2009 [7], sugar has a positive effect on the stability of the 400 °C for 30 min. The dye-covered TiO2 electrode and Pt counter
anthocyanin pigment. Therefore, we attempted to add sugar exter- electrode were assembled as a sandwich-type cell. A drop of Iodo-
nally to the anthocyanin pigment to stabilize the dye, and then lyte AN-50 (Solaronix) electrolyte solution was injected into the
evaluated the current conversion efficiency. Natural dyes are cell via a hole at the back of the counter electrode. The hole was
promising alternative sensitizers for DSSCs because they are easy then sealed with a hot-melt ionomer film (SX 1170, Solaronix)
to prepare, widely available, cheap, and eco-friendly. and covered with glass. Finally, the edge of each side of the FTO
glass was cleaned and soldered (ultrasonic soldering system, Mod-
Experimental part el-9200) with alloy #143 (Cerasolza) in order to achieve a good
electrical contact for measurement.
Preparation of dye sensitizer solutions
Results and discussion
The flowers of K. japonica and R. chinensis were collected fresh
and kept in a vacuum furnace for about 10 h at 65 °C to remove FTIR spectra
the moisture. After being dried, the flowers were crushed in a mor-
tar to make them into powder. About 1 g of the powdered sample Fig. 2 shows the FTIR spectra of the extracted dyes from
was dissolved in 60 mL of ethanol and kept for two days at ambient K. japonica and R. chinensis. The main pigment of the K. japonica

Fig. 1. Structure of (a) xanthophyll and (b) anthocyanin.


K.V. Hemalatha et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 96 (2012) 305–309 307

Fig. 2. FT-IR spectra of natural dye extracted from K. japonica and R. chinensis.

flower is carotenoid. The strong absorption of carotenoid in the


blue–green spectral region is caused by the strongly allowed
S0–S2 transition. The peaks at 1080 and 1646 cm 1 are attributed
to the symmetric vibrational modes of C–C and C@C stretching,
respectively. These are important spectroscopic measurements
since they reflect the structural properties of the carotenoid
molecule [8]. The intense bands at 1066 cm 1 are attributed to
the C–O–C vibrational modes of acid and carbohydrate groups.
The strong and broad bands at 3000–3700 cm 1 are due to
the –OH groups of water. The two peaks at 2923 and 2850 cm 1
correspond to the –CH stretching mode [9,10]. The strong
characteristic absorption spectrum observed with R. chinensis
may be due to the superimposition of the absorptions of two main
anthocyanin pigments. The two peaks at 2922 and 2850 cm 1
correspond to the –CH stretching modes. The peak at a wavenum-
ber of 3393 cm 1 corresponds to the OH stretching vibration for
the anthocyanin dye of R. chinensis. The spectrum also contains a
peak at 1726 cm 1, corresponding to the C@O stretching vibration. Fig. 3. UV absorption spectra of natural dye solution extracted from (a) K. japonica,
R. chinensis and (b) TiO2 adsorbed dye of K. japonica and R. chinensis.
This indicates that the anthocyanin dye has a partial quinonoidal
form [11]. absorption spectra of the TiO2-adsorbed carotenoid and anthocya-
nin dyes. Broad absorption peaks in all the samples indicate that
UV–Vis absorption spectra the dyes adsorbed on the TiO2 particles well, increasing the
absorption of light by the TiO2 nanoparticles in the visible range.
Fig. 3(a) shows the absorption spectra of the K. japonica and The intensity of the absorption spectrum is high for R. chinensis
R. chinensis dye solutions with and without sugar molecules. compared to K. japonica. Among the absorption spectrum of K.
Absorption peaks were observed for K. japonica between 400 and japonica the intensity is lower for the carotenoid pigment treated
500 nm without the addition of sugar, coinciding with the absorp- with sugar molecules than for the untreated dye. In the case of
tion spectrum obtained by Liu et al., 2008, for yellow chrysanthe- TiO2-adsorbed anthocyanin also, the intensity is low for the
mum [12]. The spectrum obtained between 400 and 500 nm sugar-treated dye compared to the bare dye (Fig. 3b). Although
confirms the extraction of the carotenoid pigment. According to the sugar has a positive effect on the stabilization of the anthocy-
Gao et al., 2000 [13], light absorption in the range 380–520 nm anin pigment, it does not improve the efficiency. The efficiency is
and molar extinction coefficients larger than 105 make certain slightly higher for the bare dye than for the sugar-treated dye for
carotenoids potential sensitizer materials for solar photovoltaic anthocyanin sensitized device, whereas for the carotenoid
cells and other artificial photochemical devices. Upon the addition sensitized the efficiency is slightly high for the sugar treated dye
of sugar molecules to the extracted dye, the intensity of the peak compared to the bare dye.
increases. This is because, as per the Beer–Lambert law, the
amount of light absorbed by a body is proportional to the number
of absorbing particles in it. The increase in the number of sugar IV characteristics of DSSCs sensitized with natural dyes
molecules in the pigment slightly increases the amount of visible
light absorption of the carotenoid pigment. Absorption peaks were Table 1 shows that the conversion efficiencies of the DSSCs fab-
observed for R. chinensis between 500–600 nm confirms the ricated using carotenoid pigment extracted from the flower petals
extraction of anthocyanin pigment. The intensity of the absorption of K. japonica treated with and without sugar are 0.29% and 0.22%,
spectrum of the dye solution of R. chinensis, increased with the the open-circuit voltages (VOC) are 0.5526 and 0.5839 V, the short-
addition of sugar, and the absorption wavelength was around circuit current densities (JSC) are 0.7509 and 0.5597 mA/cm2
550 nm higher than that of bare dye (Fig. 3a). Fig. 3b shows the (Fig. 4a), and the fill factors (FF) are 0.70 and 0.67, respectively.
308 K.V. Hemalatha et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 96 (2012) 305–309

Table 1 performance, whereas they enhance the performance of the carot-


Characteristics of the dye sensitized solar cell fabricated by the carotenoid and enoid sensitized cell.
anthocyanin pigment of K. japonica and R. chinensis respectively with and without
sugar molecule.
The cells sensitized with extracts of Blue pea (0.05%), Annatto
(0.19%), Norbixin (0.13%), Spinach (0.13%), Frutus lyciia (0.17%),
Sample VOC (V) JSC (mA/cm2) Fill factor Efficiency (%) Lily (0.17%) [14–17], and so on showed poor performances (less
K. japonica 0.5839 0.5597 0.6775 0.22 than 0.2%) compared to our dyes. Some of the pigments extracted
K. Japonica + sugar 0.5526 0.7509 0.7045 0.29 from Ipomoea (0.27%), Begonia (0.24%), Tangerine peel (0.28%),
R. chinensis 0.5433 0.8017 0.664 0.29
R. chinensis + sugar 0.5373 0.7025 0.6938 0.27
Marigold (0.23%), yellow rose(0.26%), China loropetal (0.27%),
Chinese rose (0.27%), Dragon fruit(0.22%) [16–18] and so on
showed similar efficiencies to our pigments. The fill factors of
natural dyes range from 20 to 70%, the short-circuit currents from
0.2 to 0.8 mA/cm2, and the open-circuit voltages from 0.2 to 0.6 V.
With our dye, we obtained a better fill factor (up to 70%) compared
with other natural dyes.
From the above results, it is seen that most natural-dye-based
DSSCs produce efficiencies of less than 1%. The main reasons for
the low efficiencies of natural dyes are the structure of the pigment,
the anchoring group in the natural dye (interaction of dye with
TiO2), and the stability. If the natural dye structure has long chains
with R groups such as xanthophylls, they hinder the bonding of
the pigment with the oxide surface of the TiO2, thus preventing
the dye molecules from becoming arrayed effectively on the TiO2
film. This results in less electron transfer from the dye molecules
to the conduction band of TiO2. In the present study, the carotenoid
pigments have long alkyl groups that hinder the arraying of dye
molecules on the TiO2 compared to the anthocyanin, so the caroten-
oid pigments showed low efficiency. Most natural dyes have OH and
oxygen ligands and lack carboxylic groups. The carboxylic group is
mainly responsible for anchoring, combining with the hydroxyl of
the TiO2 particles to produce ester groups and boosting the coupling
effect of the electrons in the TiO2 conduction band to give a rapid
electron-transport rate [19]. Carotenoid pigments do not have the
carboxylic group for effective anchoring, and this is the main reason
for their lower efficiency compared to the anthocyanin dye. Another
reason is their stability. Natural dyes are less stable than inorganic
dyes (N719). The low stabilities of both carotenoid and anthocyanin
lead to their very low efficiencies as sensitizers in DSSCs.

Conclusion

We have demonstrated the generation of photovoltaic currents


upon the illumination of FTO electrodes coated with TiO2 nano-
crystals and sensitized with the pigments carotenoid and anthocy-
anin, which were extracted from the flowers of K. japonica and
R. chinensis, respectively, using a simple extraction technique.
The peaks at 1080 and 1646 cm 1 for the carotenoid molecule
Fig. 4. Photocurrent–voltage characteristics of the DSSC using the natural dye of (a)
K. japonica and (b) R. chinensis.
are attributed to the symmetric vibrational modes of the C–C and
C@C stretching. The partial appearance of the quinonoidal form
Here, the efficiency, short-circuit current density, and fill factor of of the anthocyanin molecule is confirmed by the peak at
the sugar-treated dye were increased slightly compared to those of 1726 cm 1, which corresponds to the C@O stretching vibration.
the bare dye, but the open-circuit voltages were similar. The con- The broad absorption peak of both the pigments in the visible
version efficiency of the DSSC fabricated from the anthocyanin range indicates the adsorption of dye molecules to the TiO2 nano-
dyes of R. chinensis extract is 0.29%, VOC is 0.5433 V, JSC is particles. The addition of sugar molecules to the carotenoid pig-
0.8017 mA/cm2, and FF is 0.66. Furthermore, on treating the ex- ment enhanced the short-circuit current from 0.559 to
tract with sugar, the conversion efficiency, VOC, and JSC are reduced 0.750 (mA/cm2) and the fill factor from 0.677 to 0.705, whereas
to 0.27%, 0.5373 V, and 0.7025 mA/cm2, respectively. The fill factor the addition of sugar molecules to the anthocyanin pigment sup-
is similar to that of the bare dye (0.69) (Fig. 4b). pressed the current density, and thereby caused a decrease in effi-
It can be seen from Table 1 that the anthocyanin dye of R. chin- ciency. Although the fill factors for the extracted pigments
ensis gives a better photoelectric conversion efficiency in the DSSC remained high, the current conversion efficiencies were very low
without the sugar treatment. Since the extracted dye was a crude for both K. japonica and R. chinensis compared to N719 dye. The
mixture with co-adsorbents that were involved in enhancing the efficiency could be improved further by the addition of different
performance of the dye, little variation is evident in the conversion stabilizing agents or binders to the pigments in which the research
efficiency on treating the pigment with sugar. Usually, co-sensitiz- is under progress. The simple extraction procedure, low cost, wide
ers are used to stimulate the cell performance. However, as far as availability, and environmentally friendly nature make such natu-
anthocyanin is concerned, the sugar molecules suppress the cell ral dyes promising alternative sources of sensitizers for DSSCs.
K.V. Hemalatha et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 96 (2012) 305–309 309

Acknowledgment [7] F. Queiroz, C. Oliveira, O. Pinho, I.M. Ferreira, J. Agric. Food Chem. 57 (2009)
0712–10717.
[8] T. Polvka, V. Sundstrom, Chem. Rev. 104 (2004) 2021.
The first author would like to thank the ‘‘2011 Post-Doc. Devel- [9] I.R. Bunghez, M. Raduly, S. Doncea, I. Aksahin, R.M. Ion, Dig. J. Nanomater. Bios.
opment Program’’ of Pusan National University, Busan, South Korea 6 (2011) 1349–1356.
[10] T.L. Rand, J.Z. Deli, P. Molnar, G. Toth, Helvetica Chim. Acta 85 (2002) 1691–
for its financial support.
1697.
[11] P. Luo, H. Niu, G. Zheng, X. Bai, M. Zhang, W. Wang, Spectrochim. Acta, Part A
74 (2009) 936–942.
References [12] B.Q. Liu, X.P. Zhao, W. Luo, Dyes pigments 76 (2008) 327–331.
[13] F.G. Gao, A.J. Bard, L.D. Kispert, J. Photochem. Photobiol. A. 130 (2000) 49–56.
[1] G.K.R. Senadeera, T. Kitamura, Y. Wada, S. Yanagida, Sol. Energy Mater. Sol. [14] K. Wongcharee, V. Meeyoo, S. Chavadej, Sol. Energy Mater. Sol. Cells 91 (2007)
Cells 88 (2005) 315–322. 566–571.
[2] J. Preat, D. Jacquemin, E.A. Perpète, Environ. Sci. Technol. 44 (2010) 5666–5671. [15] N.M. Gomez-Ortiz, I.A. Vazquez-Maldonado, A.R. Perez-Espadas, G.J. Mena-
[3] S. Furukawa, H. Iino, T. Iwamoto, K. Kukita, S. Yamauchi, Thin Solid Films 518 Rejon, J.A. Azamar-Barrios, G. Oskam, Sol. Energy Mater. Sol. Cells 94 (2009)
(2009) 526–529. 40–44.
[4] G. Calogero, G.D. Marco, S. Cazzanti, S. Caramori, R. Argazzi, A.D. Carlo, C.A. [16] H. Chang, H.M. Wu, T.L. Chen, K.D. Huang, C.S. Jwo, Y.J. Lo, J. Alloys Compd. 495
Bignozzi, Int. J. Mol. Sci. 11 (2010) 254–267. (2010) 606–610.
[5] O. Emter, H. Falk, P. Sitte, Protoplasma 157 (1990) 128–135. [17] H. Zhou, L. Wu, Y. Gao, T. Ma, J. Photochem. Photobiol. A. 219 (2011) 188–194.
[6] U.G. Chandrika, in: T. Bechtold, R. Mussak (Eds.), Handbook of Natural [18] A. Ram, N. Nayan, Int. J. Integr. Eng. 2 (2010) 55–62.
Colourants, John Wiley & Sons, Ltd., United Kingdom, 2009, pp. 221–236. [19] M.R. Narayan, Renew. Sust. Energ. Rev. 16 (2012) 208–215.

You might also like