Open access peer-reviewed chapter

Chlorophyll as Photosensitizer in Dye-Sensitized Solar Cells

Written By

Abdul Kariem Arof and Teo Li Ping

Submitted: 30 October 2016 Reviewed: 17 February 2017 Published: 10 May 2017

DOI: 10.5772/67955

From the Edited Volume

Chlorophyll

Edited by Eduardo Jacob-Lopes, Leila Queiroz Zepka and Maria Isabel Queiroz

Chapter metrics overview

3,309 Chapter Downloads

View Full Metrics

Abstract

Chlorophyll, being the most abundant pigment that commonly found in plants, bacteria, bryophytes and algae, plays a vital role in photosynthesis. Chlorophylls are natural pigments and therefore safe, environmental friendly, easily available and cheap. Chlorophyll has been experimented to function as a photosensitizer in dye-sensitized solar cells (DSSCs) as DSSCs mimic the photosynthesis process in green plants. DSSC was first developed by Gratzel in 1991 and since then has gained tremendous attention as its fabrication is cheap and easy. A DSSC basically comprises a semiconductor that has been soaked in sensitizing dye (chlorophyll), a counter electrode, and an electrolyte containing a redox mediator. The dye absorbs light, which is transformed into electricity. Chlorophyll can be extracted from the leaves of pomegranate, bougainvillea, papaya, Pandanus amaryllifolius, spinach, green grasses, seaweeds, algae and bryophytes. Chlorophyll from these sources has been studied as possible photosensitizers for DSSCs. Most researches done in chlorophyll DSSC use the extracted natural pigments. The type of solvent and pH of the dye solution will also affect the stability of chlorophyll and subsequently the performance of the DSSCs. This chapter will present an inexhaustive overview on DSSCs using chlorophyll as dye.

Keywords

  • chlorophyll
  • photosensitizer
  • light adsorption
  • dye-sensitized solar cells
  • efficiency

1. Introduction

Over billions of years, Mother Nature has been converting light from the sun into energy via photosynthesis. Sunlight is the most abundant and sustainable energy source that is free. The Earth receives energy from the sun at the rate of ~12 × 1017 J s−1 [1]. This has exceeded the yearly worldwide energy consumption rate of ~1.5 × 1013 J s−1 [1]. Therefore, it is a challenge to devise an approach for the effective capture and storage of solar energy for our consumption since fossil fuels such as oil and gas will be depleted in the years to come. In order to imitate the photosynthesis process, Gratzel and coworkers have developed dye-sensitized solar cells (DSSCs) based on the similar working mechanism [2]. Nevertheless, one main difference between photosynthesis of plants and DSSCs is that the energy can be stored in plants for later use but DSSC is unable to store energy. Ever since the birth of DSSCs, they have become the spotlight of attention among scientists and researchers around the world as they are much cheaper, easier to fabricate, and more environmental friendly when compared with conventional silicon solar cells [3, 4]. A DSSC is an electrochemical device that comprises a transparent-conducting oxide (TCO) glass over which is deposited a semiconductor. The semiconductor will be soaked in a dye solution. An electrolyte with reduction-oxidation (redox) mediator and cathode are the other remaining components. The fluorine-doped tin oxide (FTO)/semiconductor/dye assembly is referred to as photoanode. Indium-doped tin oxide (ITO) and FTO are two TCOs used commonly in DSSCs. Titanium dioxide (TiO2) is one of the popular semiconductors used for DSSC since it is cheap, non-toxic, and possesses a large bandgap [5]. TiO2 is deposited on the TCO substrate in the form of TiO2 nanoporous particle network to increase the coverage area for the sensitizing dye. The cathode is made up of another TCO on top of which platinum is deposited. Carbon and conducting polymers can also be employed as counter electrode. If a gel polymer electrolyte is used, it is sandwiched between the photoanode and cathode. The dye, on the other hand, can be categorized into two groups: synthetic and natural. The most frequently used synthetic dye is the ruthenium (Ru)-based dyes but they are not environmental friendly since Ru is a heavy metal [6]. Such dyes are also very expensive due to the scarcity of Ru. By contrast, natural dyes are readily available and thus cheap besides being non-toxic, environmental friendly, biodegradable, easily extracted as well as can be used without any purification [6]. Since DSSC mimics the photosynthesis of green plants, therefore chlorophyll can also function as photosensitizer for DSSC. In fact, report on chlorophyll as photosensitizer on zinc oxide (ZnO) semiconductor was first published by Tributsch in 1972 [7].

Advertisement

2. Basic working principle of chlorophyll-sensitized DSSC

In this chapter, discussion is based on the TiO2 semiconductor photoanode. However, occasionally we refer to zinc oxide (ZnO) and tin dioxide (SnO2). The dye is chlorophyll extracted from various sources including leaves, grasses, flowers, seaweeds, and algae. The electrolyte is generally in the form of liquid and quasi-solid state. The commonly used mediator is the I/I3 redox couple and the counter electrode chosen in the preceding discussion is platinum (Pt) or carbon (C). Upon shining light on the cell, the molecules of the chlorophyll dye (D) will be excited (D*) after absorbing photons () and inject electrons into the semiconductor conduction band as described in the equation below:

hv+DD*E1

The excited chlorophyll molecules (D*) will inject electrons into the TiO2 conduction band and the excited dye will then be oxidized or ionized (D+). The reaction process involved is as follows:

D*+TiO2D++ecb(TiO2)E2

The oxidized chlorophyll dye molecules (D+) will accept electrons from an iodide ion (I) in the electrolyte when the I ions were released to the oxidized molecules and in turn oxidized to triiodide ions (I3) according to the equation below:

2D++3II3+2DE3

The electron in the TiO2 conduction band flows out of the device through the load to reach the counterelectrode and reduced the triiodide ion as follows:

I3+2e3IE4

The iodide ion is now restored, the electron circuit is completed, and the whole system is back to its original state to start a new cycle. These processes will continue as long as there is light and current is produced in the external circuit continuously. Under illumination, the voltage generated is given by the energy difference between the photoanode’s Fermi level and the electrolyte’s redox potential. Figure 1 illustrates the schematic diagram of the chlorophyll-sensitized DSSC and its operating principle.

Figure 1.

Operating principle of a chlorophyll-sensitized DSSC.

The light to electricity conversion efficiency (η) of the chlorophyll DSSC can be calculated from the equation below:

η=Jsc×Voc×FFPin×100%E5

Here Jsc is the short circuit current density (unit: mA cm−2), which is obtained without any external applied voltage or potential, Voc is the open circuit voltage (unit: mV) obtained under the condition of open circuit when there is no current, Pin is the input power (total incident light power density), and FF is the fill factor which can be expressed as

FF=Jmax×VmaxJsc×VocE6

Here Jmax and Vmax are the photocurrent density and voltage in the J-V curve at the maximum power output. Each single component in a DSSC is important to ensure good performance. The main focus in this chapter is the chlorophyll dye. The dye is to absorb light, injects electrons into the semiconductor, and receives electrons from the redox mediator in the electrolyte. The cycle continues. An efficient dye sensitizer should display unique characteristics as listed below [8, 9]:

  • Absorb light in the visible region.

  • Good attachment at the surface of photoelectrode to ensure fast electron transfer.

  • Good interfacial properties and high stability to enable good absorption to TiO2.

  • Easily accepting replacement electron from electrolyte.

  • Excited state of dye must be slightly above the TiO2 conduction band and its ground-state level below the redox potential of the electrolyte.

  • Lifetime of the dye must be consistent with device life.

  • Stable enough to sustain about 20 years exposure to natural light.

Advertisement

3. Performance of chlorophyll-sensitized DSSCs

Table 1 summarizes the performance of some DSSCs employing chlorophyll as photosensitizer reported by researchers worldwide. Herein, the illumination of the chlorophyll-sensitized DSSCs was carried out under intensity of 100 mW cm−2 unless stated otherwise.

DyePhotoanodeElectrolyteCounter electrodeJsc (mA cm−2)Voc (V)FFη (%)Ref.
Anethum graveolens leaves (fresh)TiO2/FTOI/I3 LEPt/FTO0.9650.5790.4000.220[10]
Anethum graveolens leaves (dried)TiO2/FTOI/I3 LEPt/FTO0.4540.5620.3200.080[10]
Arugula leaves (fresh)TiO2/FTOI/I3 LEPt/FTO0.7880.5990.4200.200[10]
Arugula leaves (dried)TiO2/FTOI/I3 LEPt/FTO0.7130.5940.4300.180[10]
Parsley leaves (fresh)TiO2/FTOI/I3 LEPt/FTO0.5350.4450.3400.070[10]
Parsley leaves (dried)TiO2/FTOI/I3 LEPt/FTO0.4480.5530.4000.090[10]
Bougainvillea spectabilis flowerTiO2/ITOI/I3 GPEPEDOT/FTO1.1100.5000.5860.325[11]
Amaranthus caudatus flowerTiO2/ITOI/I3 GPEPEDOT/FTO1.8200.5500.6100.610[11]
Cordyline fruticosa leavesTiO2Not statedNot stated1.3000.6160.6020.500[12]
Pawpaw leavesTiO2/FTOLENot stated0.6490.5040.6050.200[13]
Pomegranate leavesTiO2/ITOI/I3 LEPt/FTO2.0500.5600.5200.597[14]
Platanus orientalis L. (Chinar leaves)TiO2/ITOI/I3 LEPt/ITO0.0120.4680.0040.550[15]
Shiso leavesTiO2/FTOp-CuI-3.5200.4320.3900.590[16]
Bougainvillea leavesAu/TiO2/FTOI/I3 LEPt/FTO3.2300.5000.4100.618[17]
Ocimum Gratissimum (scent leaves)TiO2/FTOI/I3 LEPt/FTO0.0440.4660.4000.021[18]
Spinach oleraceaTiO2/ITOI/I3 LEPt/ITO0.4670.5500.5100.131[19]
Spinach oleracea (fresh)TiO2/FTOI/I3 LEPt/FTO0.3320.5900.4200.080[10]
Spinach oleracea (dried)TiO2/FTOI/I3 LEPt/FTO1.1100.5830.4600.290[10]
ZnO/FTOI/I3 LEPt/FTO0.1230.2260.2000.008[10]
Red spinach leavesTiO2/ITOI/I3 LEC/FTO1.0000.5050.5780.583[20]
TiO2/ITOI/I3 LEC/FTO0.7000.5590.4550.357[20]
TiO2/ITOI/I3 LEC/FTO0.5000.7500.3940.296[20]
Green spinach leavesZnO/ITOI/I3 LEC/ITO0.0520.5900.5300.016[21]
Papaya leavesTiO2/FTOI/I3 LEPt/FTO0.3600.3250.5600.070[22]
TiO2/ITOI/I3 LEC/FTO0.060 mA0.3940.250-[23]
Jatropha leavesTiO2/ITOI/I3 LEC/FTO0.042 mA0.3500.250-[23]
Ipomoea leaves extractTiO2/ITOI/I3 LEPt/ITO0.8500.4950.5360.233[19]
TiO2/ITOI/I3 LEPt/ITO0.9140.5400.5630.278[19]
TiO2/ITOI/I3 LEPt/ITO0.8250.5330.5480.259[19]
TiO2/ITOI/I3 LEPt/ITO1.1200.5650.5920.318[19]
TiO2/ITOI/I3 LEPt/ITO0.9820.5430.5640.292[19]
TiO2/ITOI/I3 LEPt/ITO0.9150.5100.5520.253[19]
Azadirachta indica (Neem) leavesTiO2/FTOI/I3 LEC/FTO0.4300.4040.4010.720[24]
TiO2/FTOI/I3 LEPt/FTO0.2300.4670.3920.050[25]
Ziziphus jujuba leaves (dried)TiO2/FTOI/I3 LEPt/FTO3.1800.6520.5191.077[26]
Basil leaves (dried)TiO2/FTOI/I3 LEPt/FTO1.3980.5810.4990.409[26]
Basil flowerTiO2/FTOI/I3 LEPt/FTO1.1200.6000.4000.270[27]
Mint flowerTiO2/FTOI/I3 LEPt/FTO0.4500.5600.3800.090[27]
Mint leaves (dried)TiO2/FTOI/I3 LEPt/FTO0.9800.5790.4000.227[26]
Lemon leavesaTiO2/FTOI/I3 LEC/FTO1.0800.5920.1000.036[28]
Morula leavesaTiO2/FTOI/I3 LEC/FTO0.0590.4720.0500.001[28]
Fig leaves (dried)TiO2/FTOI/I3 LEPt/FTO2.0910.5960.5150.642[26]
Berry leaves (dried)TiO2/FTOI/I3 LEPt/FTO3.5730.5950.4410.939[26]
Pandanus amaryllifolius leavesTiO2/ITOGPEPt/ITO1.6100.3600.4100.240[29]
TiO2/FTOGPEPt/FTO1.1900.4900.6300.390[30]
TiO2/FTOGPEPt/FTO1.9100.4800.5600.510[31]
Banana leaves (dried)TiO2/FTOI/I3 LEPt/FTO1.7700.5960.4920.522[26]
Peach leaves (dried)TiO2/FTOI/I3 LEPt/FTO2.5550.6110.4220.659[26]
Black tea leavesTiO2/FTOI/I3 LEPt/FTO0.3900.5500.4000.080[27]
Coccinia indica leavesSnO2/FTOI/I3 LENot stated0.7000.5400.6100.260[32]
La-SnO2/FTOI/I3 LENot stated0.8200.5400.5400.290[32]
La-Cu-SnO2/FTOI/I3 LENot stated1.0100.5600.5100.310[32]
Ficus retusa Linn.Au-TiO2/ITOCe4+/3+ LEPt/ITO7.8500.5200.2891.180[33]
Garcinia suubellipticaAu-TiO2/ITOCe4+/3+ LEPt/ITO6.4800.3220.3310.691[33]
PerillaTiO2/FTOI/I3 LEPt/FTO1.3600.5220.6960.500[34]
PetuniaTiO2/FTOI/I3 LEPt/FTO0.8500.6160.6050.320[34]
Eggplant pulpTiO2/FTOI/I3 LEPt/FTO0.3500.6300.3900.090[27]
Festuca ovina grassTiO2/ITOI/I3 LEC/ITO1.1890.5480.6990.460[35]
Hierochloe odorata grassTiO2/FTOI/I3 LEPt/FTO2.1990.5940.3550.460[36]
Torulinium odoratum grassTiO2/FTOI/I3 LEPt/FTO1.0040.6540.4830.320[36]
Dactyloctenium aegyptium grassTiO2/FTOI/I3 LEPt/FTO0.6980.7190.4810.240[36]
Moss bryophyte (hyophila involuta)TiO2/FTOGPEPt/FTO5.7800.6000.5701.970[37]
TiO2/FTOGPEPt/FTO4.5900.6100.6401.770[38]
TiO2/FTOGPEPt/FTO5.9600.5800.5802.000[38]
TiO2/FTOGPEPt/FTO3.7100.6400.7201.690[38]
TiO2/FTOGPEPt/FTO5.3700.5500.7302.170[38]
TiO2/FTOGPEPt/FTO8.4400.5400.5802.620[38]
Rhoeo spathacea (Sw.) StearnAu-TiO2/ITOCe4+/3+ LEPt/ITO10.9000.4960.2741.490[33]
Sargassum wightii (marine seaweed)bZnO/FTOI/I3 LEPt/FTO0.2030.3300.4600.070[39]
Kelp (brown algae)TiO2/TCOI/I3 LEPt/TCO0.4330.4410.620-[40]
Undaria pinnatifida (brown seaweed)TiO2/FTOI/I3 LEPt/FTO0.8000.3600.6900.178[41]
TiO2/FTOI/I3 LEPt/FTO10.7000.5300.6003.400[42]
TiO2/FTOI/I3 LEPt/FTO13.8000.5700.5804.600[42]
TiO2/FTOI/I3 LEPt/FTO8.6000.4700.6002.500[42]
TiO2/FTOI/I3 LEPt/FTO9.0000.4700.6102.600[42]
Cladophora sp. (green algae)TiO2/FTOI/I3 LEPt/FTO0.1450.5850.5900.055[43]
Green algae (fresh)TiO2/FTOI/I3 LEPt/FTO0.1340.4160.2100.010[10]
Green algae (dried)TiO2/FTOI/I3 LEPt/FTO0.3970.5590.4400.100[10]
Chlorella vulgaris (microalgae)TiO2/FTOI/I3 LEPt/FTO2.5300.5510.6500.900[44]

Table 1.

The photovoltaic performance of some chlorophyll-sensitized DSSCs.

aIntensity 80 mW cm−2.

bIntensity 45 mW cm−2.

LE: liquid electrolyte; GPE: gel polymer electrolyte; PEDOT: poly(3,4-ethylenedioxythiophene).

It is evident that the condition of leaves whether fresh or dried affects the adsorption of chlorophyll onto the photoanode surface and consequently the performance. Taya et al. [10] observed that DSSCs having chlorophyll extracted from fresh leaves of Anethum graveolens (Indian traditional medicinal herb and spice) and arugula (arugula salad leaves) exhibited better performance than DSSCs with A. graveolens and arugula leaves that have been dried for 1 week. On the other hand, higher efficiencies were detected in DSSCs with dried parsley, spinach, and green algae as compared to fresh ones. However, the authors did not discuss the reason behind this. Among parsley, arugula, A. graveolens, Spinach oleracea, and green algae, chlorophyll extracted from spinach produced the best efficiency of 0.290 % [10]. The efficiency of the DSSCs depends on the soaking temperature and time. The 0.290% efficiency was obtained when the TiO2 photoanode was soaked in the spinach extract solution at 60°C for 12 h [10]. Decreasing the temperature yielded low efficiency with η = ~0.0380% at 30°C. Beyond the optimum temperature, the efficiency decreased to ~0.175% (70°C) and ~0.0190% (80°C), respectively. Reducing the TiO2 soaking time in the chlorophyll spinach solution to 2 h gave poor efficiency of ~0.021% [10]. Beyond 12 h of soaking, no obvious change in efficiency was observed from the spinach chlorophyll-sensitized DSSC [10]. Therefore, it can be inferred that the freshness of chlorophyll leaves, the soaking temperature of TiO2 in chlorophyll solution, and its duration influenced the DSSC performance. From Table 1, comparison has been made between TiO2 and ZnO photoanodes for chlorophyll spinach DSSCs where the former gave better performance than the latter [10].

In the case of chlorophyll extract from ipomoea leaves (leaves of morning glory flower), 50°C is the optimum temperature for TiO2 immersion with efficiency of 0.278% [19]. Lower efficiencies of 0.233 and 0.259% were obtained when the TiO2-soaking temperature in ipomoea leaves extract solution were at 30 and 80°C, respectively for 24 h [19]. Other than temperature, pH of the dye solution is another factor influencing the efficiency. Maintaining the soaking temperature at 50°C, the pH of ipomoea dye solutions was adjusted to pH 1, 2, and 3 [19]. However, there was no mention on the type of acid used. Thus, it is not known whether the anion of acid had any influence on the DSSC performance. Nevertheless, improvement in efficiency can be seen when the acidity of the dye solution was adjusted to pH 1 and 2 with efficiencies of 0.318 and 0.292%, respectively. However, further increasing the pH to 3 decreased the efficiency (η = 0.253%) [19].

The type of solvent used for pigments extraction can also give different results in the absorption spectrum [4547]. From the work of Al-Alwani et al. [45], it has been reported that the UV-vis absorption spectra of chlorophyll extracted from Pandanus amaryllifolius (screwpine leaves) and Cordyline fruticosa (commonly known as Ti plant or cabbage palm) in ethanol and methanol solution displayed highest intensity absorption peaks among other solvents such as n-butyl alcohol, ethyl-acetate, n-hexane, chloroform, acetonitrile, ethyl-ether, and petroleum ether. Then, 1 g TiO2 powder was added in chlorophyll extracted from P. amaryllifolius and C. fruticosa in respective solvents (ethanol for Pandanus leaves and methanol for Cordyline leaves) and water at different ratios. The addition of an appropriate quantity of water into the respective alcohol solvent increased the polarity of solution for better dye adsorption on the TiO2 surface [45]. For solution containing TiO2 and P. amaryllifolius chlorophyll, the best absorption spectrum was obtained at 2:1 of ethanol to water ratio, whereas the optimum ratio for TiO2-C. fruticosa solution with mixture solvents of methanol and water was 3:1. Better dye absorption onto TiO2 surface is said to improve the performance of DSSCs but there is no DSSC results in Ref. [45]. Nonetheless, in recent publication [12], the same group of authors have turned to response surface methodology (RSM) approach to investigate the various parameters involved in the chlorophyll extraction process from C. fruticosa simultaneously and then predict their response in order to obtain the optimized condition for its extraction. After taking three factors into consideration, that is, boiling temperature for organic solvents (acetonitrile, ethanol, and methanol), different pH ranging from 4 to 8 and temperature for chlorophyll extraction from 50 to 90°C, it was found that chlorophyll can be best extracted from the cordyline leaves under the condition where the solvent was ethanol (boiling point 78°C), pH of 7.99, and at temperature of 78.33°C [12]. As a result, the efficiency of 0.500% was obtained for DSSC with C. fruticosa chlorophyll as listed in Table 1.

It should be noted that both betalain and chlorophyll pigments can be extracted from the flowers of Amaranthus caudatus (common name: love-lies-bleeding, velvet flower) and Bougainvillea spectabilis using different solvents. Chlorophyll pigments have been obtained when 0.1 mol L−1 hydrochloric acid (HCl) was used to dissolve the amaranthus and bougainvillea flowers, whereas ethanol as solvent will yield betalain pigments from the same flowers. Surprisingly, DSSCs having chlorophyll from amaranthus and bougainvillea flowers in Table 1 demonstrated better performance as compared to DSSCs with betalain from the same flowers. The Jsc, Voc, FF, and η values of betalain DSSC from B. spectabilis were 0.081 mA cm−2, 0.450 V, 0.483, and 0.018%, respectively [11]. DSSCs using betalain extracted from A. caudatus flower gave the Jsc of 0.102 mA cm−2, Voc of 0.530 V, FF of 0.610, and η of 0.033% [11]. The efficiencies exhibited by chlorophyll-sensitized DSSCs from amaranthus (η = 0.610 %) and bougainvillea (η = 0.  325 %) flowers were surprisingly high considering that gel polymer electrolyte and PEDOT counterelectrode were used instead of conventional liquid electrolyte and Pt electrode [11].

Khan and coworkers [20] have examined the effect of acid treatment on TiO2 nanoparticles in the making of TiO2 paste to be coated on ITO glass substrate via the doctor blade method. It is found that chlorophyll from red spinach leaves-sensitized DSSC without any acid treatment on TiO2 photoanode exhibited the efficiency of 0.296% which is lower than the TiO2 acid-treated DSSC with chlorophyll extracted from the same source under intensity of 50 mW cm−2 [20]. The presence of acid can prevent agglomeration of TiO2 nanoparticles and results in better TiO2 dispersion and thereby offer more adsorption sites for the dye molecules [20]. Khan et al. [20] used citric acid (organic acid) and nitric acid (inorganic acid) to prevent TiO2 agglomeration in the DSSC fabricated with chlorophyll from red spinach leaves, and TiO2 treated with citric acid gave higher efficiency of 0.583% compared to that using nitric acid treatment on TiO2 electrode (η = 0.357%). The lower efficiency yielded by nitric acid treatment could be due to nitric acid being a strong oxidizing acid and its corrosive nature may ruin the TiO2 surface [48]. From Table 1, the Jsc and FF of the DSSC with chlorophyll extracted from red spinach leaves can also be seen to increase with efficiency following the order η (TiO2) < η (nitric acid treated TiO2) < η (citric acid treated TiO2), whereas the Voc values decreased in the same order.

It can be noted from Table 1 that SnO2 was employed as photoanode instead of TiO2 in the cell having chlorophyll extracted from Coccinia indica (ivy gourd) leaves [32]. Comparison with TiO2 revealed that SnO2 is chemically stable and has larger bandgap of 3.6 eV and higher electron mobility [49, 50]. Due to its wide bandgap, it is less sensitive to UV degradation and thus possesses better stability as compared to TiO2 [51]. Nonetheless, its large bandgap will also cause SnO2 to have lower open circuit voltage. SnO2 having high electron mobility can yield fast electron transport and therefore electron recombination can be decreased. From the table, the cells with chlorophyll from C. indica leaves exhibited the efficiencies of 0.260, 0.290, and 0.310% using three photoanodes, that is, SnO2, La-doped SnO2, and La-Cu-doped SnO2, respectively [32]. Doping elements into DSSC photoanode improved the performance since more dye molecules can be adsorbed in the working electrode due to larger surface area owing to increased roughness and pores after doping [52].

Chang et al. [17] have investigated the plasmonic effect of gold (Au) nanoparticles with an average size of 27 nm in TiO2 DSSC using chlorophyll from bougainvillea leaves. An efficiency of 0.618% was obtained. The Au nanoparticles showed localized surface plasmon resonance behavior when the frequency of the incident light came close to the surface plasmon frequency of Au and consequently improved light absorption leading to a considerably high efficiency of 0.618% as listed in Table 1. Also, the interface between Au and TiO2 formed a Schottky barrier where electrons will be blocked from re-entering the dye or electrolyte, which decreased electron recombination and improved the DSSC performance. Earlier report on TiO2 loaded with Au nanoparticles prior to chlorophyll sensitization was published by Lai and coworkers [33]. Instead of using I/I3 LE, the authors employed water-based electrolyte at ethanol:water ratio of 7:13 with Ce4+/3+ as redox couple since water-based DSSC can be totally free from toxic and is biologically friendly. The chlorophyll was extracted from herbal plant Rhoeo spathacea (Sw.) Stearn. Unexpectedly, an efficiency of more than 1% was produced by the chlorophyll R. spathacea water-based DSSC which is higher than that of a similar cell but with one of the earliest synthetic dyes, that is, crystal violet (η = 0.010%) [33]. The authors attributed the aggregation of crystal violet dyes as the culprit behind this based on its photocurrent density value (Jsc = 2.040 mA cm−2), which is lower than that of R. spathacea cell (Jsc = 10.900 mA cm−2) [33]. In fact, the other two water-based DSSCs using chlorophyll extracted from Ficus retusa Linn. (common name: bonsai plant) and Garcinia subelliptica (common name: happiness tree) have also demonstrated higher efficiency than that of crystal violet-sensitized DSSC with efficiencies of 1.180 and 0.691%, respectively [33].

From Table 1, it can be observed that most of the DSSCs employ liquid electrolytes based on I/I3 redox mediator. The maximum power conversion efficiencies of liquid electrolyte-based DSSCs using synthetic dyes have reached around 14% [53]. Liquid-based electrolytes are desired since they can infiltrate into the TiO2 nanopores network to make contact with the dye molecules for dye regeneration. Still, liquid electrolyte-based cells have limited durability due to the possibility of leakage and volatility of solvents. Some of the solvents are flammable as well. In an attempt to avoid these complications, researchers worldwide are focusing on developing polymer-based electrolytes for DSSCs. Solid polymer electrolytes can exhibit reasonable ionic conductivities but have poor interfacial contact with electrodes. Hence, gel-type polymer electrolytes (GPEs) are being developed. GPEs, which are basically liquid electrolyte trapped in the polymer matrix, have good flexibility and conductivities comparable to those of liquid electrolytes. It can be clearly seen from Table 1 that the DSSC using chlorophyll extracted from moss bryophyte and gel polymer electrolyte exhibited exceptionally high efficiencies of ~2–2.620% under different conditions. The bryophyte cell with η = 1.970% was obtained with GPE having polyacrylonitrile (PAN) as polymer host, tetrapropylammonium iodide (TPAI) salt, iodine, ethylene carbonate (EC), and propylene carbonate (PC) as solvent and plasticizer [37]. Efficiencies of 1.770 and 2.000% were attained for bryophyte cells using GPE based on poly(vinyl alcohol) (PVA) with single [potassium iodide (KI)] and double salts (KI and TPAI), respectively [38]. Higher efficiency and Jsc values observed in the cell having binary salts GPE could be most probably due to the higher number of iodide ions contained in the GPE. As for the second best performing chlorophyll bryophyte DSSC (η = 2.170 %), it is acquired using GPE based on poly(vinyl alcohol) (PVA) and double salts of potassium iodide (KI) and TPAI with the addition of 0.7 M tert-butylpyridine (TBP) [38]. TBP can be used either by incorporating it in electrolyte or photoanode in order to improve the Voc and subsequently η. However, in this case where TBP was added into GPE, the TBP effect is insignificant as compared to the bryophyte cell when the working electrode was immersed in TBP for 1 h (η = 1.690%) [38]. Nonetheless, the efficiency of the latter is lower than the former owing to its lower Jsc value which might be due to lesser photon harvesting as a result of reflection and light scattering by TBP [38]. The most efficient chlorophyll bryophyte DSSC having η of 2.620% was attained when a co-adsorbent, that is, chenodeoxycholic acid (CDCA), was added in the moss bryophyte [38]. The GPE used was PVA-based double salt without TBP. CDCA served as spacer to prevent the self-aggregation of chlorophyll molecules, diminish electron recombination, and stabilize the chlorophyll, thereby improving the efficiency.

It is worth mentioning from Table 1 that the cell having the efficiency of 0.590% with chlorophyll extracted from shiso leaves used copper iodide (CuI) as hole transport material (HTM) instead of conventional liquid electrolyte [16]. Therefore, the DSSC has the configuration of FTO/TiO2/chlorophyll dye/CuI. CuI, a p-type semiconductor, has bandgap of 3.1 eV and good optical transparency [54, 55]. The p-CuI was coated onto the chlorophyll/TiO2/FTO using dip- and spray-coating technique as this method involves low calcination temperature and thus the degradation of dye will not occur [16, 54]. The p-CuI solid-state DSSC has similar working principle with conventional DSSC except that after photon absorption, the dye molecules will be excited and then inject electrons and holes into TiO2 and p-CuI, respectively. This indicates that the dye at ground state must be positioned below CuI valence band and the dye-excited state should be above the TiO2 conduction band in order to ensure proper functioning of chlorophyll CuI DSSC. With the usage of HTM, there will be no issue on pigment deterioration since natural pigment is unstable against the oxidized species in electrolyte with iodine as redox mediator [16].

Most of the reports on chlorophyll-sensitized DSSCs summarized in Table 1 do not contain information on the type of chlorophyll used. Among the six chlorophylls, chlorophyll a, which plays vital role in photosynthesis process, shows poor adsorption and sensitization on TiO2 due to its structure that contains phytyl and alkyl groups causing steric hindrance that obstruct the chlorophyll molecules to bind efficiently with TiO2 molecules [40, 56, 57]. The structure of chlorophyll b only differs from chlorophyll a by the aldehyde group (–CHO) rather than methyl group (–CH3). On the other hand, chlorophyll c, which consists of chlorophyll c1 and chlorophyll c2, has carboxyl group (–COOH) that can effectively attach to TiO2 as reported in Ref. [41]. Chlorophyll c, which is the main pigment in Undaria pinnatifida (brown seaweed) yielded the efficiency of 0.178% when applied in DSSC as listed in Table 1. Wang et al. [42] have purified the pigments in U. pinnatifida to obtain chlorophyll c and remove chlorophyll a and carotenoids. Then, the purified chlorophyll c was subjected to polyethylene column chromatography to isolate chlorophyll c1 and chlorophyll c2. Using the same method, chlorophyll c1 and chlorophyll c2 in oxidized form can also be obtained. As a result, efficiencies of 3.400 and 4.600% have been obtained from the chlorophyll c1- and chlorophyll c2-sensitized DSSCs with liquid electrolyte [42]. Decrement in efficiency can be seen in DSSCs employing oxidized chlorophyll c1 (designated as chlorophyll c’1) and oxidized chlorophyll c2 (chlorophyll c’2) with the values of 2.500 and 2.600%, respectively [42]. Nonetheless, to the best of our knowledge, the chlorophyll-sensitized DSSC utilizing chlorophyll c2 extracted from U. pinnatifida exhibited the highest efficiency among other chlorophyll-sensitized DSSCs till date. However, there remains the stability issue encountered by chlorophyll-sensitized DSSCs. Therefore, further work must be done to enhance the stability as well as improving the chlorophyll-sensitized DSSCs performance before they can be put into practical usage.

Advertisement

4. Summary

It has been shown that chlorophyll has good potential to serve as photosensitizer in dye-sensitized solar cells. Moreover, they are cheap, non-toxic, biodegradable, easily found, and easy to use as sensitizer. Although the efficiency is still considerably low with highest efficiency to date being only 4.600% from DSSC with U. pinnatifida chlorophyll c2, there remains the possibility and room for improvement to further enhance the performance and improve stability of chlorophyll-sensitized DSSCs for practical applications.

Advertisement

Acknowledgments

The authors thank the University of Malaya and Malaysian Ministry of Higher Education (MOHE) for the UMRG grant no. RP024C-14AFR and PRGS grant no. PR001-2014A.

References

  1. 1. Blankenship RE, Tiede DM, Barber J, Brudvig GW, Fleming G, Ghirardi M, Gunner MR, Junge W, Kramer DM, Melis A, Moore TA, Moser CC, Nocera DG, Nozik AJ, Ort DR, Parson WW, Prince RC, Sayre RT. Comparing photosynthetic and photovoltaic efficiencies and recognizing the potential for improvement. Science 2011;332:805–809.
  2. 2. O’ Regan B, Grätzel M. A low-cost, high efficiency solar cell based on dye-sensitized colloidal TiO2 films. Nature 1991;353:737–740.
  3. 3. Grätzel M. Photoelectrochemical cells. Nature 2001;414:338–344.
  4. 4. Bella F, Bongiovanni R. Photoinduced polymerization: An innovative, powerful and environmentally friendly technique for the preparation of polymer electrolytes for dye sensitized solar. Review. Journal Photochemistry and Photobiology C 2013;16:1–21.
  5. 5. Karruppuchamy S, Jeong, JM, Amalnerkar DP, Minoura H. Photoinduced hydrophilicity of titanium dioxide thin films prepared by cathodic electrodeposition. Vacuum 2006;80:494–498.
  6. 6. Fernando JMRC and Senadeera GKR. Natural anthocyanins as photosensitizers for dye-sensitized solar devices. Current Science 2008;95:663–666.
  7. 7. Tributsch H. Reaction of excited chlorophyll molecules at electrodes and in photosynthesis. Journal of Photochemistry and Photobiology 1972;16:261–269.
  8. 8. Grätzel M. Dye-sensitized solar cells. Journal of Photochemistry and Photobiology C: Photochemistry Reviews 2003;4:145–153.
  9. 9. Longo C and De Paoli M-A. Dye-sensitized solar cells: A successful combination of materials. Journal of Brazilian Chemistry Society 2003;14:889–901.
  10. 10. Taya SA, El-Agez TM, El-Ghamri HS, Abdel-Latif MS. Dye-sensitized solar cells using fresh and dried natural dyes. International Journal of Materials Science and Applications. 2013;2:37–42.
  11. 11. Godibo DJ, Anshebo ST, Anshebo TY. Dye sensitized solar cells using natural pigments from five plants and quasi-solid state electrolyte. Journal of Brazilian Chemical Society 2015;26:92–101.
  12. 12. Al-Alwani MAM, Ludin NA, Mohamad AB, Kadhum AAH, Baabbad MM, Sopian K. Optimization of dye extraction from Cordyline fruticosa via response surface methodology to produce a natural sensitizer for dye-sensitized solar cells. Results in Physics 2016;6:520–529.
  13. 13. Kimpa MI, Momoh M, Isah KU, Yahya HN, Ndamitso MM. Photoelectric characterization of dye sensitized solar cells using natural dye from pawpaw leaf and flame tree flower as sensitizers. Materials Science and Applications 2012;3:281–286.
  14. 14. Chang H, Lo Y-J. Pomegranate leaves and mulberry fruit as natural sensitizers for dye-sensitized solar cells. Solar Energy 2010;84:1833–1837.
  15. 15. Liu B-Q, Zhao X-P, Luo W. The synergistic effect of two photosynthetic pigments in dye-sensitized mesoporous TiO2 solar cells. Dyes and Pigments 2008;76:327–331.
  16. 16. Kumara GRA, Kaneko S, Okuya M, Onwona-Agyeman B, Konno A, Tennakone K. Shiso leaf pigments for dye-sensitized solid-state solar cell. Solar Energy and Solar Cells 2006;90:1220–1226.
  17. 17. Chang H, Kao M-J, Chen, T-L, Kuo H-G. Natural sensitizer for dye-sensitized solar cells using a three layers of photoelectrode thin films with a Schottky barrier. American Journal of Engineering and Applied Science 2011;4:214–222.
  18. 18. Eli D, Musa GP, Ezra D. Chlorophyll and betalain as light-harvesting pigments for nanostructured TiO2 based dye-sensitized solar cells. Journal of Energy and Natural Resources. 2016;5:53–58.
  19. 19. Chang H, Wu HM, Chen TL, Huang KD, Jwo CS, Lo YJ. Dye-sensitized solar cell using natural dyes extracted from spinach and ipomoea. Journal of Alloys and Compounds 2010;495:606–610.
  20. 20. Khan MA, Khan SMM, Mohammed MA, Sultana S, Islam JMM, Uddin J. Sensitization of nanocrystalline titanium dioxide solar cells using natural dyes: Influence of acids medium on coating formulation. American Academic & Scholarly Research Journal. 2012;4:10.
  21. 21. Hasoon SA, Al-Haddad RMS, Shakir OT, Ibrahim IM. Natural dye-sensitized solar cell based on zinc oxide. International Journal of Scientific & Engineering Research. 2015;6:137–142.
  22. 22. Suyitno S, Saputra TJ, Supriyanto A, Arifin Z. Stability and efficiency of dye-sensitized solar cells based on papaya-leaf dye. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 2015;148:99–104.
  23. 23. Pramono SH, Maulana E, Prayogo AF, Djatmika R. Characterization of dye-sensitized solar cell (DSSC) based on chlorophyll dye. International Journal of Applied Engineering Research 2015;10:193–205.
  24. 24. Swarnkar AK, Sahare S, Chander N, Gangwar RK, Bhoraskar SV, Bhave TM. Nanocrystalline titanium dioxide sensitised with natural dyes for eco-friendly solar cell application. Journal of Experimental Nanoscience 2015;10:1001–1011.
  25. 25. Rosana NTM, Amarnath DJ, Anandan S, Saritha G. Environmental friendly photosensitizing materials for harvesting solar energy. Journal of Materials and Environmental Science 2015;6:2053–2059.
  26. 26. Taya SA, El-Agez TM, Elrefi KS, Abdel-Latif MS. Dye-sensitized solar cells based on dyes extracted from dried plant leaves. Turkish Journal of Physics 2015;39:24–30
  27. 27. Abdel-Latif MS, Abuiriban MB, Dahoudi NA, Al-Kahlout AM, Taya SA, El-Agez TM, El-Ghamri HS. Dye-sensitized solar cells using fifteen natural dyes as sensitizers of nanocrystalline TiO2. Science, Technology and Development 2015;34:135–139.
  28. 28. Maabong K, Muiva CM, Monowe P, Sathiaraj TS, Hopkins M, Nguyen L, Malungwa K, Thobega M. Natural pigments as photosensitizers for dye-sensitized solar cells with TiO2 thin films. International Journal of Renewable Energy Research. 2015;5:54–60.
  29. 29. Noor MM, Buraidah MH, Yusuf SNF, Careem MA, Majid SR, Arof AK. Performance of dye-sensitized solar cells with (PVDF-HFP)-KI-EC-PC electrolyte and different dye materials. International Journal of Photoenergy 2011;2011:5 pages, Article ID 960487.
  30. 30. Shah S, Buraidah MH, Teo LP, Careem MA, Arof AK. Dye-sensitized solar cells with sequentially deposited anthocyanin and chlorophyll as sensitizers. Optical and Quantum Electronics 2016;46:219, 8 pages.
  31. 31. Noor MM, Buraidah MH, Careem MA, Majid SR, Arof AK. An optimized poly(vinylidene fluoride-hexafluoropropylene)-NaI gel polymer electrolyte and its application in natural dye sensitized solar cells. Electrochimica Acta 2014;121:159–167.
  32. 32. Priyadharsini CI, Prakasam A, Anbarasan PM. Enhanced photoelectrochemical performance of DSSC by co-doped SnO2 nanoparticles. International Letters of Chemistry, Physics and Astronomy 2013;17:82–93.
  33. 33. Lai WH, Su YH, Teoh LG, Hon MH. Commercial and natural dyes as photosensitizers for a water-based dye-sensitized solar cell loaded with gold nanoparticles. Journal of Photochemistry and Photobiology A: Chemistry 2008;195:307–313.
  34. 34. Zhou H, Wu L, Gao Y, Ma T. Dye-sensitized solar cells using 20 natural dyes as sensitizers. Journal of Photochemistry and Photobiology A: Chemistry 2011;219:188–194.
  35. 35. Hernandez-Martinez AR, Estevez M, Vargas S, Quintanilla F, Rodriguez R. Natural pigments-based dye-sensitized solar cells. Journal of Applied Research and Technology 2012;10:38–47.
  36. 36. Shanmugan V, Manoharan S, Sharafali A, Anandan S, Murugan R. Green grasses as light harvesters in dye sensitized solar cells. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 2015;135:947–952.
  37. 37. Hassan HC, Abidin ZHZ, Careem MA, Arof AK. Chlorophyll as sensitizer in I/I3-based solar cells with quasi-solid-state electrolytes. High Performance Polymers 2014;26:647–652.
  38. 38. Hassan HC, Abidin ZHZ, Chowdhury FI, Arof AK. A high efficiency chlorophyll sensitized solar cell with quasi solid PVA based electrolyte. International Journal of Photoenergy 2016;2016:9 pages, Article ID 3685210.
  39. 39. Anand M, Suresh S. Marine seaweed Sargassum wightii extract as a low-cost sensitizer for ZnO photoanode based dye-sensitized solar cell. Advances in Natural Sciences: Nanoscience and Nanotechnology 2015;6:035008, 8 pages.
  40. 40. Hao S, Wu J, Huang Y, Lin J. Natural dyes as photosensitizers for dye-sensitized solar cells. Solar Energy 2006;80:209-214.
  41. 41. Calogero G, Citro I, Di Marco G, Minicante SA, Morabito M, Genovese G. Brown seaweed pigment as a dye source for photoelectrochemical solar cells. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 2014;117:702–706.
  42. 42. Wang X-F, Zhan C-H, Maoka T, Wada Y, Koyama Y. Fabrication of dye-sensitized solar cells using chlorophylls c1 and c2 and their oxidized forms c1 and c2 from Undaria pinnatifida (Wakame). Chemical Physics Letters 2007;447:79–85.
  43. 43. Lim A, Manaf NH, Tennakoon K, Chandrakanthi RLN, Lim LBL, Bandara JMRS, Ekanayake P. Higher performance of DSSC with dyes from Cladophora sp. As mixed cosensitizer through synergistic effect. Journal of Biophysics 2015;2015:8 pages, Article ID 510467.
  44. 44. Mohammadpour R, Janfaza S, Abbaspour-Aghdam F. Light harvesting and photocurrent generation by nanostructured photoelectrodes sensitized with a photosynthetic pigment: A new application for microalgae. Bioresource Technology 2014;163:1–5.
  45. 45. Al-Alwani MAM, Mohamad AB, Kadhum AAH, Ludin NA. Effect of solvents on the extraction of natural pigments and adsorption onto TiO2 for dye-sensitized solar cell applications. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 2015;138:130–137.
  46. 46. Ayalew WA, Ayele DW. Dye-sensitized solar cells using natural dye as light-harvesting materials extracted from Acanthus sennii chiovenda flower and Euphorbia cotinifolia leaf. Journal of Science: Advanced Materials and Devices 2016;1:488–494.
  47. 47. Mohammed IK, Kasim UI, Yabagi JA, Taufiq S. The effect on extracting solvents using natural dye extracts from Hyphaene thebaica for dye-sensitized solar cells. Journal of Material Science & Engineering 2015;4:208, 3 pages.
  48. 48. Hao S, Wu J, Fan L, Huang Y, Lin J, Wei Y. The influence of acid treatment of TiO2 porous film electrode on photoelectric performance of dye-sensitized solar cell. Solar Energy 2004;76:745–750.
  49. 49. Dou X, Sabba D, Mathews N, Wong LH, Lam YM, Mhaisalkar S. Hydrothermal synthesis of high electron mobility Zn-doped SnO2 nanoflowers as photoanode material for efficient dye-sensitized solar cells. Chemistry of Materials 2011;23:3938–3945.
  50. 50. Mao X, Zhou R, Zhang S, Ding L, Wan L, Qin S, Chen Z, Xu J, Miao S. High efficiency dye-sensitized solar cells constructed with composites of TiO2 and the hot-bubbling synthesized ultra-small SnO2 nanocrystals. Scientific Reports 2016;6:19390. doi: 10.1038/srep19390 (2016)
  51. 51. Bhande SS, Shinde DV, Tehare KK, Patil SA, Mane RS, Naushad M, Alothman ZA, Hui KN, Han SH. DSSCs synergic effect in thin metal oxide layer-functionalized SnO2 photoanodes. Journal of Photochemistry and Photobiology A: Chemistry 2014;295:64–69.
  52. 52. Arunachalam A, Dhanapandian S, Manoharan C. Effect of Sn doping on the structural, optical and electrical properties of TiO2 films prepared by spray pyrolysis. Physica E 2016;76:35–46.
  53. 53. Kakiage K, Aoyama Y, Yano T, Oya K, Fujisawa J-i, Hanaya M. Highly-efficient dye-sensitized solar cells with collaborative sensitization by silyl-anchor and carboxy-anchor dyes. Chemical Communications 2015;51:15894–15897.
  54. 54. Tennakone K, Kumara GRRA, Kumarasinghe ARK, Wijayantha KGU, Sirimanne PM. Dye-sensitized nano-porous solid-state photovoltaic cell. Semiconductor Science and Technology 1995;10:1689–1693.
  55. 55. Amalina MN, Rasheid NA, Rusop M. The properties of sprayed nanostructured p-type CuI films for dye-sensitized solar cells application. Journal of Nanomaterials 2012:2012:6 pages, Article ID 637637.
  56. 56. Kay A, Grätzel M. Artificial photosynthesis I. Photosensitization of TiO2 solar cells with chlorophyll derivatives and related natural porphyrins. The Journal of Physical Chemistry 1993;97:6272–6277.
  57. 57. Calogero G, Bartolotta A, Di Marco G, Di Carlo A, Bonaccorso F. Vegetable-based dye-sensitized solar cells. Chemical Society Reviews 2015;44:3244–3294.

Written By

Abdul Kariem Arof and Teo Li Ping

Submitted: 30 October 2016 Reviewed: 17 February 2017 Published: 10 May 2017