Next Article in Journal
Analysis of Indoor Air Pollutants and Guidelines for Space and Physical Activities in Multi-Purpose Activity Space of Elementary Schools
Next Article in Special Issue
Influence of Cr Doping on Structural, Optical, and Photovoltaic Properties of BiFeO3 Synthesized by Sol-Gel Method
Previous Article in Journal
Analyzing the Impact of Cybersecurity on Monitoring and Control Systems in the Energy Sector
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Toward Eco-Friendly Dye-Sensitized Solar Cells (DSSCs): Natural Dyes and Aqueous Electrolytes

1
Department of New Energy Engineering, Seoul National University of Science and Technology, 232 Gongneung-ro, Nowon-gu, Seoul 01811, Korea
2
Department of Chemical and Biomolecular Engineering, Seoul National University of Science and Technology, 232 Gongneung-ro, Nowon-gu, Seoul 01811, Korea
3
Material & Component Convergence R&D Department, Korea Institute of Industrial Technology, Ansan 15588, Korea
*
Authors to whom correspondence should be addressed.
Energies 2022, 15(1), 219; https://doi.org/10.3390/en15010219
Submission received: 28 November 2021 / Revised: 12 December 2021 / Accepted: 27 December 2021 / Published: 29 December 2021

Abstract

:
Due to their low cost, facile fabrication, and high-power conversion efficiency (PCE), dye-sensitized solar cells (DSSCs) have attracted much attention. Ruthenium (Ru) complex dyes and organic solvent-based electrolytes are typically used in high-efficiency DSSCs. However, Ru dyes are expensive and require a complex synthesis process. Organic solvents are toxic, environmentally hazardous, and explosive, and can cause leakage problems due to their low surface tension. This review summarizes and discusses previous works to replace them with natural dyes and water-based electrolytes to fabricate low-cost, safe, biocompatible, and environmentally friendly DSSCs. Although the performance of “eco-friendly DSSCs” remains less than 1%, continuous efforts to improve the PCE can accelerate the development of more practical devices, such as designing novel redox couples and photosensitizers, interfacial engineering of photoanodes and electrolytes, and biomimetic approaches inspired by natural systems.

1. Introduction

Since the late 19th century, renewable energy resources, such as solar, hydropower, geothermal, biomass, and biofuel energy, have attracted much attention to substitute fossil fuel, the main cause of global warming [1]. Solar energy is a limitless energy source without the emission of CO2, which accounts for a large portion of greenhouse gas. Solar energy can be converted into other types of energy by photovoltaic or photothermal mechanisms [2,3,4]. A solar cell utilizes solar energy by converting sunlight into electricity based on the photovoltaic mechanism. While a silicon solar cell is the representative one, other types of cells, such as organic solar cells, thin-film solar cells, dye-sensitized solar cells (DSSCs), and perovskite solar cells, have also been studied as economical alternatives [5,6,7]. DSSCs are one of the third-generation photovoltaic devices suggested as an alternative to conventional Si-based solar cells. DSSCs have various advantages, such as a low cost and robust fabrication process, reasonable power conversion efficiency (PCE), and semitransparency [8,9]. The color of the photoanode can be varied using different dyes adsorbed on it. Additionally, DSSCs operate efficiently even under low-light intensity, enabling them to be used indoors.
DSSCs are composed of a transparent electrode, a photoanode, a dye sensitizer, an electrolyte, and a counter electrode. In 1991, Brian O’Regan and Michael Grätzel reported DSSCs with a high PCE of 7.12%, based on a transparent nanoporous film of titanium dioxide (TiO2) and ruthenium (Ru) complex dye [10]. Since then, researchers have developed novel dyes, electrolytes, photoanodes, and counter electrodes, to further improve the efficiency of DSSCs. Recently, a high PCE of 14.34% was attained by Kakiage et al. via the co-sensitization of alkoxysilyl-anchor dye (ADEKA-1) and a carboxy-anchor organic dye (LEG4) using various co-adsorbents with a (Co(phen)3)2+/3+ (phen = 1,10-phenanthroline) redox electrolyte [11]. Concerning tandem-type cells, Eom et al. reported a high PCE value of 14.64% in a tandem cell structure, where alkylated thieno(3,2-b)indole-based organic dye (SGT-137) and Zn(II)-porphyrin dye (SGT-021) with a (Co(bpy)3)2+/3+ redox electrolyte were used [12]. For high-efficiency DSSCs, Ru-complexes and organic solvents are mainly used as dyes and electrolyte solvents, respectively. However, Ru-complex dyes are expensive and synthesized via complicated synthesis processes [13,14]. Ru compounds are treated as moderately toxic, environmentally hazardous, and carcinogenic. Moreover, organic solvents are generally toxic and explosive and cause environmental problems [15,16,17]. Ironically, DSSCs that mimic photosynthesis in natural leaves to produce energy are made of materials that could be harmful to nature. Because of these problems, researchers are making efforts to increase efficiency and develop more environmentally friendly DSSCs.
Herein, we review various efforts to fabricate DSSCs based on eco-friendly components, such as natural photosensitizers and water-based electrolytes. First, the harmful effect of typical high-efficient DSSCs on the environment is discussed. Next, we discuss recent research to employ natural dyes derived from nature and aqueous electrolytes in DSSCs. Finally, the example studies of DSSCs fully based on environmentally benign dyes and electrolytes are discussed.

2. Problems or Harmful Effects

For high-efficiency DSSCs, Ru-complex dyes and organic solvent-based electrolytes are typically used. However, due to some issues in terms of possible human toxicity, potential environmental impact, production cost, stability, and safety, they are unfavorable and may need to be replaced with other materials.
For high-efficiency DSSCs, electrolytes with low viscosity, high dielectric constant, good solubility, and high chemical stability are required [18,19,20]. Various solvents have been used for DSSCs. Table 1 presents the melting point, boiling point, vapor pressure (P), and viscosity (η) of organic solvents popularly used. Nitrile-based solvents, such as acetonitrile (ACN) and 3-methoxypropionitrile (MPN), are considered the most preferred solvents for electrolytes. Ethylene carbonate (EC), propylene carbonate (PC), γ-butyrolactone (GBL), and N-methyl-2-pyrrolidone (NMP) are also used due to their low vapor pressure and volatility. However, such organic solvents are not the best choice in terms of safety [19]. ACN can be metabolized to produce hydrogen cyanide, which is the source of the observed toxic effects in microsomes, especially in the liver [21,22]. EC is converted into ethylene glycol, which is toxic alcohol, causing metabolic acidosis during ingestion [23,24]. GBL is the precursor of γ-hydroxybutyrate (GHB), which can affect the central nervous system [25,26]. Additionally, some organic solvents have low viscosity, resulting in easy electrolyte leakage and high flammability. Regarding health issues, a high volatility at room temperature causes absorption into the human body due to the high exposure possibility to the solvents. Thus, the solvents are unsuitable given the fabrication of DSSCs that are safe for humans and the environment.
In DSSCs, the dye sensitizer plays a crucial role in absorbing light and converting it into electricity. For high PCE, having a wide range of absorption wavelengths in visible light is important. Ru-complex dyes absorb a wide range of absorption wavelength from 300 nm to 600 nm in visible light, resulting in high-efficiency DSSCs with the maximum efficiency of 11.18% [27,28]. The Ru-complex dyes, such as N719, N3, and black dye, are the most widely used due to their long-excited lifetime, wide absorption wavelength, and highly efficient metal-to-ligand charge transfer, despite their low molar extinction coefficients [29,30,31,32,33]. However, Ru-complex dyes are expensive, need sophisticated and complex syntheses processes [13,14], and cause environmental problems, which could be problems to be used in cost-effective and eco-friendly DSSCs. Although Ru is nontoxic, its compounds, such as ruthenium oxide (RuO4), are highly toxic and volatile [34,35,36]. Materials that undergo dye synthesis processes are also harmful to health and cause environmental pollution. For example, ammonium thiocyanate and hydrochloric acid are harmful to health, have high causticity, and generate chlorine [37]. Recently, metal-free organic dyes were developed to replace Ru-based dyes, but organic dyes could be toxic and carcinogenic and produce hazardous pollutants during their synthesis [38,39]. All in all, the typical components in DSSCs, organic solvents, and Ru-based complex dyes may need to be replaced to realize low cost, biocompatible, and environmentally benign devices. Water and natural dyes derived from plants could be excellent alternatives.

3. Natural Dyes Extracted from Nature

Natural photosensitizers are extracted from parts of plants, such as leaves, fruits, and flowers. The dyes contain anthocyanin, chlorophyll, carotenoid, and betalain. The molecular structures of the natural dyes are shown in Figure 1. Anthocyanins are generally obtained from petals of flowers and fruits and absorb 450–580 nm wavelength of visible light with a maximum peak at the 520 nm region [40,41,42,43]. Anthocyanins contain carbonyl (–CO–) and hydroxyl (–OH) groups. The functional groups enable the molecules to be stably bound to the surface of TiO2, which facilitates electron injection from anthocyanin molecules to the conduction band of TiO2 [44,45,46,47]. Chlorophyll, a key molecule in photosynthesis, is a green pigment commonly found in green leaves and plants. Chlorophyll absorbs 400–450 and 640–680 nm wavelength of visible light and has a maximum peak at 430 nm [48,49]. Carotenoids are yellow, orange, and red pigments obtained from colored vegetables, plants, and algae. Betalains are yellow and red pigments obtained from petals of flowers, fruits, leaves, and roots of plants. The process of extracting natural dye is simple and possibly environmentally friendly. Therefore, many studies have been actively performed to adopt the natural dyes as a photosensitizer of DSSCs to realize eco-friendly DSSCs.
Table 2 summarizes the previous works on DSSCs based on natural dyes, including the types and sources of naturally derived dyes, materials for photoanodes, electrolytes, and cathodes, and their PCE. Compared to those based on organic or metal-based dyes, DSSCs based on natural dyes exhibited low PCE (mostly < 1%). One of the main reasons is that natural dyes absorb a narrow range of light, which inhibits the increase in PCE, whereas organic and rare metal-based dyes have a broad absorption spectrum of visible light [27,28].
The condition of natural dyes and concentration of dye extract solutions are affected by the extraction temperature, pH, and the types of solvents. For example, high temperatures can thermally degrade dyes, whereas low temperatures can limit the solubility of dyes in extracting solvents [43]. Moreover, dyes have different solubilities depending on the types of solvents because the molecules of natural dyes have different polarities [71,83,84]. Figure 2a shows the absorbance of anthocyanin extracted from Areca catechu according to different extracting solvents [71]. Wongcharee et al. investigated the effect of the types of extracting solvents on the efficiency of the resulting DSSCs [43]. It has been reported that although anthocyanin is more soluble in ethanol than water, the photocatalytic decomposition by TiO2 occurred in the presence of ethanol, decreasing the efficiency after being exposed to sunlight for some time. It has been concluded that ethanol is unsuitable as an anthocyanin-extracting solvent. Thus, the appropriate choice of extracting solvents is important in the dye extraction process.
The pH adjustment of extracting solvents and the addition of acid can affect the properties of dyes, such as the highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) levels, absorption spectra, polarity, and stability. The addition of acid changes the HOMO and LUMO levels of dyes, which is closely related to the VOC value. Suyitno et al. investigated the HOMO and LUMO levels of chlorophyll extracted from papaya leaves (PLs) depending on the pH of the extracting solvent (Figure 2b) [58]. After the pH changed to 3.5 by acidification, the bandgap lowered from 2.30 to 2.16 eV, and the PCE was improved four times from 0.07% to 0.28%. Acidification also increases the polarity of the extracting solvents, such as ethanol or methanol, which could enhance the dye separation from source materials, thereby increasing the concentration of the dye extract solution and absorption value, which assists in higher harvesting from sunlight [43,57,85]. However, high-pH conditions can decompose natural dyes, thereby degrading the PCE of DSSCs. Therefore, it is important to carefully optimize the acid or alkali treatment conditions depending on the types of natural dyes.
The mixture of multiple natural dyes can absorb a wider range of wavelength of light (Figure 3a). For a high photocurrent, the LUMO and HOMO levels should also be above the conduction and valence bands of the photoanode materials, respectively, thereby increasing the injection of photoelectrons and reducing recombination loss [79,86]. A mixture of multiple dyes can facilitate the photoelectron injection and increase the electron lifetime due to the intermediate energy level of electrons in adjacent dye molecules (Figure 3b) [59,87]. Consequently, it has been reported that the PCE of DSSCs can be improved by employing a dye cocktail (Table 3). The most frequently used natural dye is anthocyanin because its carbonyl and hydroxyl groups form a stable bonding to the photoanode surface. The combination of dyes, the optimal mixing ratios of the dyes, and the resulting photovoltaic performances of the DSSCs are presented in Table 3. The optimal ratio may differ depending on the types of natural dyes and electrolytes. Kumar et al. fabricated a co-sensitized solar cell with a high PCE of 1.139% (Figure 3c) by mixing chlorophyll and anthocyanin in a ratio of 1:1 [59]. In their study, the dyes were extracted from cactus and bermudagrass, respectively. The dye-mixing ratio is important to obtain a maximum PCE. For example, Bashar et al. reported that when betalain and chlorophyll were mixed in an optimized mixing ratio of 4:1, a maximum PCE of 0.99% was obtained. By combining anthocyanin and chlorophyll, a maximum PCE of 1.29% was obtained in the ratio of 1:1 [29]. Instead of employing a simple mixture of multiple dyes, Kumara et al. performed the sequential adsorption of natural dyes for the layered co-sensitization (Figure 3d) [42]. The PCE of DSSCs prepared by the layered co-sensitization was 1.55%, which exceeded that of DSSCs with homogenous adsorption of the dyes (1.13%).

4. Aqueous Electrolyte

4.1. Aqueous Electrolyte

To develop eco-friendly DSSCs, efforts to replace organic electrolytes with aqueous electrolytes have progressed for several years (Table 4). In 2010, Law et al. replaced MPN with water when preparing an electrolyte (comprising 2.0 M 1-propyl-3-methylimidazolium iodide (PMII), 0.05 M iodine, 0.1 M guanidinium thiocyanate (GuSCN), and 0.5 M 4-tert-butylpyridine (TBP)). When the MPN was completely displaced by water, the PCE decreased from 5.5% to 2.4% (Figure 4a) [88]. Similarly, Vaghasiya et al. fabricated DSSCs based on aqueous electrolytes containing organic ionic liquid. The effect of water content on the PCE was investigated, showing that the PCE reduced from 5.61% (0% water) to 3.46% (100% of water) [89].
Although ionic liquids, such as PMII, 1-butyl-3-methylimidazolium iodide (BMII), and 1-ethyl-3-methylimidazolium iodide (EMII), are typically used in organic electrolytes of DSSCs, they are not completely soluble in water, and surfactants are needed to avoid phase separation in the electrolyte. Instead of using ionic liquids, water-soluble salts, such as KI, NaI, and LiI, are employed in water-based electrolytes. Bella et al. investigated the effect of the iodide/triiodide concentration in the electrolyte and the types of counter-ions, resulting in the KI salt-based electrolyte exceeding the NaI-based one in performance, attaining a PCE value of 0.8% (D131 dye, VOC = 0.488 V, JSC = 2.70 mA/cm2, and FF = 0.62) [90]. The desorption of the dye molecules from the photoanode surface in aqueous electrolytes can also decrease the performance of the DSSCs based on aqueous electrolytes than common DSSCs based on organic electrolytes. Co-absorbents, such as chenodeoxycholic acid (CDCA), which co-grafts with the dye onto the photoanode surface, prevent detachment and aggregation of the dye in water and reduce the charge recombination, resulting in improvement in PCE and stability of DSSCs based on an aqueous electrolyte [91,92,93,94,95]. With a CDCA-to-dye molar ratio of 18:1, a PCE value of 1.25% (D131 dye, VOC = 0.59 V, JSC = 3.86 mA/cm2, and FF = 0.55) was reached using an aqueous electrolyte containing 0.5 M NaI and 10 mM I2 [92].
To minimize the dye desorption in the presence of water and prevent evaporation or leakage of electrolytes, quasi-solid gel electrolytes were introduced by gelation of aqueous electrolytes. Gel electrolytes for DSSCs have been researched for decades for better device stability; however, several components in the electrolytes are still petroleum-derived, harmful, corrosive, or expensive [96]. For more eco-friendly DSSCs, aqueous and bio-derived gels based on xanthan gum (XG) [95,97,98], cellulose [96], and agarose [99] have been developed recently. XG is a water-soluble polysaccharide and a well-known stabilizing agent widely used in the food and cosmetic industries. Additionally, because XG is thixotropic, meaning that its viscosity decreases when an external force is applied, the gel electrolyte based on XG can penetrate the mesoporous TiO2 electrode [100,101]. Park et al. developed half aqueous XG-based gel electrolyte with PMII and MPN solvent for DSSCs. The resulting device reached a PCE value of 4.40% even after 288 h, indicating that the XG-based electrolyte enhanced the long-term stability [101]. Galliano et al. prepared 100% aqueous XG-based electrolyte containing NaI salt, based on which the resulting DSSC device showed only a slightly lower PCE value of 1.93% than that based on a liquid-state electrolyte (2.28%) due to lower diffusion coefficient and UV-vis absorption. Moreover, it exhibited impressive stability after more than 1500 h of the aging test (Figure 4b) [97]. Further study to enhance the PCE was conducted using a cobalt-based redox couple, leading to an overall PCE of 4.47% and stability for five days [98]. The aqueous gel electrolyte based on carboxymethylcellulose (CMC) was prepared and used for DSSCs, which showed a PCE value of 0.72% in optimum CMC concentration (5.5 wt.%), without any additives and surface treatment of the photoanode, such as UVO or TiCl4 [96]. In the study conducted by Haro et al., a bioderived gel electrolyte was developed using lignin, a lignocellulose material, which is the most available material on earth for biofuels. The resulting DSSCs with the lignin-based electrolyte exhibited a PCE value of 1.54% with VOC = 0.63 V, JSC = 3.62 mA/cm2, and FF = 0.67 [102].

4.2. Efforts to Improve Performance

The relatively poor performance of DSSCs using aqueous electrolytes compared to organic electrolytes can be attributed to various causes: (1) less wettability of the photoanode surface [104,107,108,112]; (2) desorption of the dye from the surface of the semiconductor [89,118,119,120]; (3) reduction in the diffusion coefficient [89,104,107,108,112]; (4) recombination derived from a higher concentration of free iodine [91,121]; and (5) negative shift of the conduction band [122,123,124]. To overcome these problems, endeavors to improve performances, such as adding surfactants, developing novel redox couples and hydrophobic sensitizers, and chemical and morphological modification of the photoanode surface have been performed. In this chapter, we will discuss the efforts to improve aqueous DSSCs.

4.2.1. Development of Novel Redox Couples and Photosensitizers

To overcome the low VOC and JSC values of aqueous DSSCs, novel redox couples and sensitizers were developed to have fast kinetics and a high positive redox potential, which is related to high VOC. The radical of 4-hydroxy-2,2,6,6-tetramethlypiperidinoxyl (4-hydroxy-TEMPO or TEMPOL (Figure 5a)), which has 0.7 V of redox potential in water, was developed and was added to an aqueous electrolyte in DSSCs with a D131 dye [109,110], resulting in a PCE value of 1.3% (VOC = 0.81 V, JSC = 3.1 mA/cm2, and FF = 0.56) [110]. Additionally, Kato et al. immobilized TEMPOL on the Nafion layer coated on a counter electrode to enhance the reduction peak current and achieved a PCE value of 2.1% with 1.0 M TEMPOL/TEMPOL+ and D205 dye (VOC = 0.69 V, JSC = 4.5 mA/cm2, and FF = 0.64) [109]. The high VOC compared to that of the previously reported aqueous systems was due to the high positive redox potential of the TEMPO/TEMPO+ redox couple [103]. JSC remained constant or only slightly higher than without TEMPO/TEMPO+, which was attributed to the recombination between TEMPO+ and the use of the highly hydrophobic dyes [108,125,126]. Fayad et al. introduced a new water-soluble redox couple based on a thiolate/disulfide (T/DS) in an aqueous electrolyte and new zwitterionic dye (T169) to improve poor wetting, showing excellent performance with VOC = 0.55 V, JSC = 13.30 mA/cm2, FF = 0.62, and PCE = 4.50% [108].

4.2.2. Interface Engineering of Photoanodes and Aqueous Electrolytes

Surfactants are widely used to impede phase separations in aqueous electrolytes, including organic ionic liquids [88,127], and improve incomplete wettability between hydrophobic semiconductor photoanode and aqueous electrolytes by reducing the interfacial tension [106,128]. For a 100% aqueous electrolyte, Zhang et al. applied ionic surfactants (AOT and FK-1 as anionic surfactants and CTAB and FC-134 as cationic surfactants) to 100% water-electrolyte incorporating NaI and I2. Both the anionic and cationic surfactants could improve the PCE from 2.51% (without surfactant) to 2.98% (with 0.1 wt.% of AOT) and 3.96% (with 0.2 wt.% of FC-134) due to the better wettability of the aqueous electrolyte and dye-coated TiO2 layer [107]. However, since the surfactants could decrease the DSSC performance by imposing a diffusion limitation of the redox couple, caution should be taken to use the appropriate redox couple and surfactants [103].
Chemical and morphological modification of photoanode surfaces could also enhance the performance of DSSCs. Miyasaka et al. treated a TiO2 photoanode with ozone and UV light to increase its hydrophilicity and absorbed dye on the TiO2 surface in the presence of tert-butylpyridine (TBP) to reinforce the dye–TiO2 binding. The PCE was enhanced from 0.6% to 1.1% with an aqueous electrolyte (0.5 M KI and 25 mM I2 in water) [114,115]. Furthermore, an increase in the active surface area is a method for improving the photocurrent and, therefore, the PCE of DSSCs. TiCl4 treatment on TiO2 surface forms a rough nanolayer of TiO2, causing the surface area augmentation and increasing the light-harvesting efficiency and adsorption of dye [129,130]. The TiCl4 treatment on the TiO2 surface also inhibited the charge recombination between the electrons and the oxidized redox couple through the barrier effect [116,131,132]. In fact, in research activities on various types of solar cells, including DSSCs, TiCl4 treatment has been actively utilized to increase efficiency. Bella et al. used the TiCl4 treatment, which improved the PCE of DSSCs based on an aqueous electrolyte comprising NaI and I2 from 1.25% to 2.37% (Figure 5b) [116]. The TiCl4 liquid deposition process could damage the semiconductor film, resulting in flaking off from the fluorine-doped tin oxide (FTO) after a long treatment time (>4 h). Therefore, Pham et al. introduced a shorter (<1 h) and more effective nanoTiO2-layer-coating method on SnO2 film using a (NH4)2TiF6 solution. With this approach, the PCE increased ten times, from 0.067% (no treatment) to 0.66% ((NH4)2TiF6 treatment), which was higher compared to the 0.204% improvement using the TiCl4 solution treatment [117]. S. Castro et al. employed anchoring molecules, trioctylmethyl ammonium dodecanedioate (DTMA) containing carboxyl groups and alkyl chains, to the TiO2 layer before the dye adsorption step. The molecules are anchored onto the TiO2 layer by acting as selective physical barriers that hinder the triiodide molecules from contacting the TiO2 layer [133].

5. Efforts to Obtain Fully Eco-Friendly DSSCs

In previous chapters, we discussed the studies where artificially synthesized dyes and organic solvent-based electrolytes were individually replaced with eco-friendly materials of natural dyes and water-based electrolytes, respectively. Here, studies to realize fully “green” DSSCs by simultaneously using both natural dyes and aqueous electrolytes are introduced (Figure 6a).
Gu et al. attained a PCE value of 0.01% using natural dye from purple cabbage and an aqueous electrolyte, including 0.5 M KI and 50 mM I2 [134]. Kim et al. improved the PCE of DSSCs based on chlorophyll and 100% aqueous electrolyte with KI/I2 via O2 plasma treatment. The treatment enhanced the hydrophilicity of the TiO2 photoanode surface, thereby increasing the PCE from 0.023% (VOC = 0.46 V, JSC = 0.089 mA/cm2, and FF = 0.56) to 0.033% (VOC = 0.46 V, JSC = 0.14 mA/cm2, and FF = 0.52) [118]. Furthermore, Hon et al. used a 35% aqueous ethanol electrolyte, including 0.1 M Ce(NO3)3/0.05 Ce(NO3)4 and Au nanoparticles, which can create a Schottky barrier between the Au nanoparticles and the TiO2 electrode to enhance the photocurrent. They achieved a PCE value of 1.49% [135].
Other examples of approaches for more eco-friendly and low-cost DSSCs, besides using natural dyes and aqueous electrolytes, are to use a CoS-deposited carbon fabric [136] and graphite electrode [134] as a counter electrode or develop the eco-friendly synthesis process of TiO2 using Terminalia arjuna bark extract. As an intriguing approach for eco-friendly DSSCs, Koo et al. reported a biomimetic, regenerable DSSC with microfluidic hydrogels inspired by the vein of a leaf (Figure 6b), which showed a PCE value of 0.21%, with VOC = 0.63 V, JSC = 0.59 mA/cm2, and FF = 0.57. In microfluidic DSSCs with organic eosin Y dye, the dyes and aqueous electrolytes could be repeatedly infused and supplied to the device through the microfluidic hydrogel network, thereby continuously regenerating the DSSCs [99].

6. Conclusions and Future Outlooks

Due to low cost, facile fabrication, and high conversion efficiency, DSSCs have attracted much attention as new renewable energy devices. To replace the expensive and toxic materials in typical DSSCs with less harmful ones, efforts using natural dyes and aqueous electrolytes have been made. The DSSCs based on natural dyes and aqueous electrolytes, however, showed a lower efficiency than conventional DSSCs due to poor wettability, desorption of dye, low-diffusion coefficient of ions, recombination of photoanodes, and negative shift of the conduction band. Various efforts, such as the combination of the dyes, addition of the surfactants, and treatment of the photoanode, have improved the performance of aqueous DSSCs. Today, a few studies on fully eco-friendly DSSCs have been reported, even though the efficiency is still very low. Attaining long-term stability of the eco-friendly DSSCs is another key task. Fortunately, it has been reported that an aqueous electrolyte could be more durable than an organic solvent-based electrolyte, possibly due to the low volatility, high surface tension, high specific heat, and high boiling point of water [19,90]. Research on fully eco-friendly DSSCs with enhanced efficiency and stability should be conducted to develop more practical energy devices with the minimum environmental footprint.

Author Contributions

Conceptualization, J.-H.K., D.-H.K. and H.-J.K.; investigation, J.-H.K., D.-H.K. and H.-J.K.; data curation, J.-H.K., D.-H.K. and H.-J.K.; writing, J.-H.K., D.-H.K., J.-H.S. and H.-J.K.; review and editing, J.-H.S. and H.-J.K.; supervision, J.-H.S. and H.-J.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Advanced Research Project funded by SeoulTech (Seoul National University of Science and Technology).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ellabban, O.; Abu-Rub, H.; Blaabjerg, F. Renewable energy resources: Current status, future prospects and their enabling technology. Renew. Sustain. Energy Rev. 2014, 39, 748–764. [Google Scholar] [CrossRef]
  2. Mekhilef, S.; Saidur, R.; Safari, A. A review on solar energy use in industries. Renew. Sustain. Energy Rev. 2011, 15, 1777–1790. [Google Scholar] [CrossRef]
  3. Wang, H.; Feng, Y.; Yu, H.; Dong, L.; Zhai, F.; Tang, J.; Ge, J.; Feng, W. Utilisation of photo-thermal energy and bond enthalpy based on optically triggered formation and dissociation of coordination bonds. Nano Energy 2021, 89, 106401. [Google Scholar] [CrossRef]
  4. Green, M.A. Photovoltaic principles. Physica E 2002, 14, 11–17. [Google Scholar] [CrossRef]
  5. Bouich, A.; Ullah, S.; Ullah, H.; Mollar, M.; Marí, B.; Touhami, M.E. Electrodeposited CdZnS/CdS/CIGS/Mo: Characterization and solar cell performance. JOM 2020, 72, 615–620. [Google Scholar] [CrossRef]
  6. Bouich, A.; Ullah, S.; Marí, B.; Atourki, L.; Touhami, M.E. One-step synthesis of FA1-xGAxPbI3 perovskites thin film with enhanced stability of alpha (α) phase. Mater. Chem. Phys. 2021, 258, 123973. [Google Scholar] [CrossRef]
  7. Wu, Q.; Guo, J.; Sun, R.; Guo, J.; Jia, S.; Li, Y.; Wang, J.; Min, J. Slot-die printed non-fullerene organic solar cells with the highest efficiency of 12.9% for low-cost PV-driven water splitting. Nano Energy 2019, 61, 559–566. [Google Scholar] [CrossRef]
  8. Rho, W.-Y.; Jeon, H.; Kim, H.-S.; Chung, W.-J.; Suh, J.S.; Jun, B.-H. Recent progress in dye-sensitized solar cells for improving efficiency: TiO2 nanotube arrays in active layer. J. Nanomater. 2015, 2015, 247689. [Google Scholar] [CrossRef] [Green Version]
  9. Mehmood, U.; Rahman, S.-u.; Harrabi, K.; Hussein, I.A.; Reddy, B. Recent advances in dye sensitized solar cells. Adv. Mater. Sci. 2014, 2014, 974782. [Google Scholar] [CrossRef] [Green Version]
  10. O’Regan, B.; Grätzel, M. A Low-Cost, High-Efficiency Solar Cell Based on Dye-Sensitized Colloidal TiO2 Films. Nature 1991, 353, 737–740. [Google Scholar] [CrossRef]
  11. 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. Chem. Commun. 2015, 51, 15894–15897. [Google Scholar] [CrossRef]
  12. Eom, Y.K.; Kang, S.H.; Choi, I.T.; Yoo, Y.; Kim, J.; Kim, H.K. Significant light absorption enhancement by a single heterocyclic unit change in the π-bridge moiety from thieno [3,2-b] benzothiophene to thieno [3,2-b] indole for high performance dye-sensitized and tandem solar cells. J. Mater. Chem. A 2017, 5, 2297–2308. [Google Scholar] [CrossRef]
  13. Jitchati, R.; Thathong, Y.; Wongkhan, K. Three synthetic routes to a commercial N3 dye. Int. J. Appl. Phys. Math. 2012, 2, 1076. [Google Scholar] [CrossRef]
  14. Rajan, A.K.; Cindrella, L. Studies on new natural dye sensitizers from Indigofera tinctoria in dye-sensitized solar cells. Opt. Mater. 2019, 88, 39–47. [Google Scholar] [CrossRef]
  15. Hemmatzadeh, R.; Mohammadi, A. Improving optical absorptivity of natural dyes for fabrication of efficient dye-sensitized solar cells. J. Theor. Appl. Phys. 2013, 7, 57. [Google Scholar] [CrossRef]
  16. Gong, J.; Liang, J.; Sumathy, K. Review on dye-sensitized solar cells (DSSCs): Fundamental concepts and novel materials. Renew. Sustain. Energy Rev. 2012, 16, 5848–5860. [Google Scholar] [CrossRef]
  17. Buitrago, E.; Novello, A.M.; Meyer, T. Third-Generation Solar Cells: Toxicity and Risk of Exposure. Helv. Chim. Acta 2020, 103, e2000074. [Google Scholar] [CrossRef]
  18. Carella, A.; Borbone, F.; Centore, R. Research progress on photosensitizers for DSSC. Front. Chem. 2018, 6, 481. [Google Scholar] [CrossRef]
  19. Bella, F.; Gerbaldi, C.; Barolo, C.; Grätzel, M. Aqueous dye-sensitized solar cells. Chem. Soc. Rev. 2015, 44, 3431–3473. [Google Scholar] [CrossRef] [Green Version]
  20. Iftikhar, H.; Sonai, G.G.; Hashmi, S.G.; Nogueira, A.F.; Lund, P.D. Progress on electrolytes development in dye-sensitized solar cells. Materials 2019, 12, 1998. [Google Scholar] [CrossRef] [Green Version]
  21. Turchen, S.G.; Manoguerra, A.S.; Whitney, C. Severe cyanide poisoning from the ingestion of an acetonitrile-containing cosmetic. Am. J. Emerg. Med. 1991, 9, 264–267. [Google Scholar] [CrossRef]
  22. Mueller, M.; Borland, C. Delayed cyanide poisoning following acetonitrile ingestion. Postgrad. Med. J. 1997, 73, 299–300. [Google Scholar] [CrossRef] [Green Version]
  23. Brent, J.; McMartin, K.; Phillips, S.; Burkhart, K.K.; Donovan, J.W.; Wells, M.; Kulig, K. Fomepizole for the treatment of ethylene glycol poisoning. N. Engl. J. Med. 1999, 340, 832–838. [Google Scholar] [CrossRef]
  24. Bove, K.E. Ethylene glycol toxicity. Am. J. Clin. Pathol. 1966, 45, 46–50. [Google Scholar] [CrossRef]
  25. Schep, L.J.; Knudsen, K.; Slaughter, R.J.; Vale, J.A.; Mégarbane, B. The clinical toxicology of gamma-hydroxybutyrate, gamma-butyrolactone and 1,4-butanediol. Clin. Toxicol. 2012, 50, 458–470. [Google Scholar] [CrossRef]
  26. Wood, D.M.; Brailsford, A.D.; Dargan, P.I. Acute toxicity and withdrawal syndromes related to gamma-hydroxybutyrate (GHB) and its analogues gamma-butyrolactone (GBL) and 1,4-butanediol (1,4-BD). Drug Test. Anal. 2011, 3, 417–425. [Google Scholar] [CrossRef]
  27. Nazeeruddin, M.K.; De Angelis, F.; Fantacci, S.; Selloni, A.; Viscardi, G.; Liska, P.; Ito, S.; Takeru, B.; Grätzel, M. Combined experimental and DFT-TDDFT computational study of photoelectrochemical cell ruthenium sensitizers. J. Am. Chem. Soc. 2005, 127, 16835–16847. [Google Scholar] [CrossRef]
  28. Wen, P.; Han, Y.; Zhao, W. Influence of TiO2 nanocrystals fabricating dye-sensitized solar cell on the absorption spectra of N719 sensitizer. Int. J. Photoenergy 2012, 2012, 906198. [Google Scholar] [CrossRef] [Green Version]
  29. Chang, H.; Kao, M.-J.; Chen, T.-L.; Chen, C.-H.; Cho, K.-C.; Lai, X.-R. Characterization of natural dye extracted from wormwood and purple cabbage for dye-sensitized solar cells. Int. J. Photoenergy 2013, 2013, 159502. [Google Scholar] [CrossRef]
  30. Vougioukalakis, G.C.; Philippopoulos, A.I.; Stergiopoulos, T.; Falaras, P. Contributions to the development of ruthenium-based sensitizers for dye-sensitized solar cells. Coord. Chem. Rev. 2011, 255, 2602–2621. [Google Scholar] [CrossRef]
  31. Sharma, G.; Singh, S.P.; Kurchania, R.; Ball, R. Cosensitization of dye sensitized solar cells with a thiocyanate free Ru dye and a metal free dye containing thienylfluorene conjugation. RSC Adv. 2013, 3, 6036–6043. [Google Scholar] [CrossRef]
  32. Sharma, K.; Sharma, V.; Sharma, S. Dye-sensitized solar cells: Fundamentals and current status. Nanoscale Res. Lett. 2018, 13, 381. [Google Scholar] [CrossRef] [PubMed]
  33. Nam, S.-H.; Lee, K.H.; Yu, J.-H.; Boo, J.-H. Review of the development of dyes for dye-sensitized solar cells. Appl. Sci. Converg. Technol. 2019, 28, 194–206. [Google Scholar] [CrossRef] [Green Version]
  34. De Freitas, E.S.; Da Silva, P.B.; Chorilli, M.; Batista, A.A.; de Oliveira Lopes, É.; Silva, M.M.d.; Leite, C.Q.F.; Pavan, F.R. Nanostructured lipid systems as a strategy to improve the in vitro cytotoxicity of ruthenium (II) compounds. Molecules 2014, 19, 5999–6008. [Google Scholar] [CrossRef] [Green Version]
  35. Dragutan, I.; Dragutan, V.; Demonceau, A. Special Issue on Ruthenium Complexes. Molecules 2017, 2, 255. [Google Scholar] [CrossRef] [Green Version]
  36. Tojo, G.; Fernández, M. Ruthenium tetroxide and other ruthenium compounds. In Oxidation of Primary Alcohols to Carboxylic Acids, 1st ed.; Springer: New York, NY, USA, 2007; pp. 61–78. [Google Scholar]
  37. Elnagar, M.M.; Samir, S.; Shaker, Y.M.; Abdel-Shafi, A.A.; Sharmoukh, W.; Abdel-Aziz, M.S.; Abou-El-Sherbini, K.S. Synthesis, characterization, and evaluation of biological activities of new 4′-substituted ruthenium (II) terpyridine complexes: Prospective anti-inflammatory properties. Appl. Organomet. Chem. 2021, 35, e6024. [Google Scholar] [CrossRef]
  38. Ismail, M.; Akhtar, K.; Khan, M.; Kamal, T.; Khan, M.A.; Asiri, A.M.; Seo, J.; Khan, S.B. Pollution, toxicity and carcinogenicity of organic dyes and their catalytic bio-remediation. Curr. Pharm. Des. 2019, 25, 3645–3663. [Google Scholar] [CrossRef]
  39. Mariotti, N.; Bonomo, M.; Fagiolari, L.; Barbero, N.; Gerbaldi, C.; Bella, F.; Barolo, C. Recent advances in eco-friendly and cost-effective materials towards sustainable dye-sensitized solar cells. Green Chem. 2020, 22, 7168–7218. [Google Scholar] [CrossRef]
  40. Pratiwi, D.; Nurosyid, F.; Supriyanto, A.; Suryana, R. Efficiency enhancement of dye-sensitized solar cells (DSSC) by addition of synthetic dye into natural dye (anthocyanin). IOP Conf. Ser. Mater. Sci. Eng. 2017, 176, 012012. [Google Scholar] [CrossRef] [Green Version]
  41. Alhamed, M.; Issa, A.S.; Doubal, A.W. Studying of natural dyes properties as photo-sensitizer for dye sensitized solar cells (DSSC). J. Electr. Devices 2012, 16, 1370–1383. [Google Scholar]
  42. Kumara, N.; Ekanayake, P.; Lim, A.; Liew, L.Y.C.; Iskandar, M.; Ming, L.C.; Senadeera, G. Layered co-sensitization for enhancement of conversion efficiency of natural dye sensitized solar cells. J. Alloys Compd. 2013, 581, 186–191. [Google Scholar] [CrossRef]
  43. Wongcharee, K.; Meeyoo, V.; Chavadej, S. Dye-sensitized solar cell using natural dyes extracted from rosella and blue pea flowers. Sol. Energy Mater. Sol. Cells 2007, 91, 566–571. [Google Scholar] [CrossRef]
  44. Sakata, K.; Saito, N.; Honda, T. Ab initio study of molecular structures and excited states in anthocyanidins. Tetrahedron 2006, 62, 3721–3731. [Google Scholar] [CrossRef]
  45. Mozaffari, S.A.; Saeidi, M.; Rahmanian, R. Photoelectric characterization of fabricated dye-sensitized solar cell using dye extracted from red Siahkooti fruit as natural sensitizer. Spectrochim. Acta Part A 2015, 142, 226–231. [Google Scholar] [CrossRef]
  46. Luque, A.; Hegedus, S. Handbook of Photovoltaic Science and Engineering; John Wiley & Sons: Hoboken, NJ, USA, 2011. [Google Scholar]
  47. Hao, S.; Wu, J.; Huang, Y.; Lin, J. Natural dyes as photosensitizers for dye-sensitized solar cell. Sol. Energy 2006, 80, 209–214. [Google Scholar] [CrossRef]
  48. Al-Alwani, M.A.; Mohamad, A.B.; Kadhum, A.A.H.; Ludin, N.A. Effect of solvents on the extraction of natural pigments and adsorption onto TiO2 for dye-sensitized solar cell applications. Spectrochim. Acta Part A 2015, 138, 130–137. [Google Scholar] [CrossRef]
  49. Goswami, D.; Sinha, D.; De, D. Nanostructured ZnO and natural dye based DSSC for efficiency enhancement. In Proceedings of the 2017 Third International Conference on Science Technology Engineering & Management, Chennai, India, 23–24 March 2017; pp. 556–560. [Google Scholar]
  50. Marhuenda-Munoz, M.; Hurtado-Barroso, S.; Tresserra-Rimbau, A.; Lamuela-Raventos, R.M. A review of factors that affect carotenoid concentrations in human plasma: Differences between Mediterranean and Northern diets. Eur. J. Clin. Nutr. 2019, 72, 18–25. [Google Scholar] [CrossRef]
  51. Bashar, H.; Bhuiyan, M.; Hossain, M.; Kabir, F.; Rahaman, M.; Manir, M.; Ikegami, T. Study on combination of natural red and green dyes to improve the power conversion efficiency of dye sensitized solar cells. Optik 2019, 185, 620–625. [Google Scholar] [CrossRef]
  52. Kabir, F.; Sakib, S.N.; Matin, N. Stability study of natural green dye based DSSC. Optik 2019, 181, 458–464. [Google Scholar] [CrossRef]
  53. Syafinar, R.; Gomesh, N.; Irwanto, M.; Fareq, M.; Irwan, Y. Chlorophyll pigments as nature based dye for dye-sensitized solar cell (DSSC). Energy Procedia 2015, 79, 896–902. [Google Scholar] [CrossRef] [Green Version]
  54. Taya, S.A.; El-Agez, T.M.; El-Ghamri, H.S.; Abdel-Latif, M.S. Dye-sensitized solar cells using fresh and dried natural dyes. Int. J. Mater. Sci. Appl. 2013, 2, 37–42. [Google Scholar] [CrossRef]
  55. Ammar, A.M.; Mohamed, H.S.; Yousef, M.M.; Abdel-Hafez, G.M.; Hassanien, A.S.; Khalil, A.S. Dye-sensitized solar cells (DSSCs) based on extracted natural dyes. J. Nanomater. 2019, 2019, 1867271. [Google Scholar] [CrossRef] [Green Version]
  56. Dinesh, V.; Sukhananazerin, A.; Sneha, J.M.; Kumar, P.M.; Biji, P. Novel stainless steel based, eco-friendly dye-sensitized solar cells using electrospun porous ZnO nanofibers. Nano Struct. Nano Objects 2019, 19, 100311. [Google Scholar] [CrossRef]
  57. Chang, H.; Wu, H.M.; Chen, T.L.; Huang, K.D.; Jwo, C.S.; Lo, Y.J. Dye-sensitized solar cell using natural dyes extracted from spinach and ipomoea. J. Alloys Compd. 2010, 495, 606–610. [Google Scholar] [CrossRef]
  58. Suyitno, S.; Saputra, T.J.; Supriyanto, A.; Arifin, Z. Stability and efficiency of dye-sensitized solar cells based on papaya-leaf dye. Spectrochim. Acta Part A 2015, 148, 99–104. [Google Scholar] [CrossRef]
  59. Kumar, K.A.; Subalakshmi, K.; Senthilselvan, J. Effect of mixed valence state of titanium on reduced recombination for natural dye-sensitized solar cell applications. J. Solid State Electrochem. 2016, 20, 1921–1932. [Google Scholar] [CrossRef]
  60. Adedokun, O.; Awodele, M.K.; Sanusi, Y.K.; Awodugba, A.O. Natural dye extracts from fruit peels as sensitizer in ZnO-based dye-sensitized solar cells. IOP Conf. Ser. Earth Environ. Sci. 2018, 173, 012040. [Google Scholar] [CrossRef]
  61. Noor, M.; Buraidah, M.; Careem, M.; Majid, S.; Arof, A. An optimized poly (vinylidene fluoride-hexafluoropropylene)–NaI gel polymer electrolyte and its application in natural dye sensitized solar cells. Electrochim. Acta 2014, 121, 159–167. [Google Scholar] [CrossRef]
  62. Diantoro, M.; Maftuha, D.; Suprayogi, T.; Iqbal, M.R.; Mufti, N.; Taufiq, A.; Hidayat, A.; Suryana, R.; Hidayat, R. Performance of pterocarpus indicus willd leaf extract as natural dye TiO2-Dye/ITO DSSC. Mater. Today Proc. 2019, 17, 1268–1276. [Google Scholar] [CrossRef]
  63. Cari, C.; Khairuddin; Septiawan, T.; Suciatmoko, P.; Kurniawan, D.; Supriyanto, A. The preparation of natural dye for dye-sensitized solar cell (DSSC). AIP Conf. Proc. 2018, 2014, 020106. [Google Scholar]
  64. Sinha, D.; De, D.; Goswami, D.; Ayaz, A. Fabrication of DSSC with nanostructured ZnO photo anode and natural dye sensitizer. Mater. Today Proc. 2018, 5, 2056–2063. [Google Scholar] [CrossRef]
  65. Maiaugree, W.; Lowpa, S.; Towannang, M.; Rutphonsan, P.; Tangtrakarn, A.; Pimanpang, S.; Maiaugree, P.; Ratchapolthavisin, N.; Sang-Aroon, W.; Jarernboon, W. A dye sensitized solar cell using natural counter electrode and natural dye derived from mangosteen peel waste. Sci. Rep. 2015, 5, 15230. [Google Scholar] [CrossRef] [Green Version]
  66. Ali, R.A.M.; Nayan, N. Fabrication and analysis of dye-sensitized solar cell using natural dye extracted from dragon fruit. Int. J. Integr. Eng. 2010, 2, 55–62. [Google Scholar]
  67. Surana, K.; Idris, M.G.; Bhattacharya, B. Natural dye extraction from Syzygium Cumini and its potential photovoltaic application as economical sensitizer. Appl. Nanosci. 2020, 10, 3819–3825. [Google Scholar] [CrossRef]
  68. Singh, S.; Singh, P.K.; Kakroo, S.; Hachim, D.M.; Dhapola, P.S.; Khan, Z.H. Eco-friendly dye sensitized solar cell using natural dye with solid polymer electrolyte as hole transport material. Mater. Today Proc. 2021, 34, 760–766. [Google Scholar]
  69. Maurya, I.C.; Singh, S.; Srivastava, P.; Maiti, B.; Bahadur, L. Natural dye extract from Cassia fistula and its application in dye-sensitized solar cell: Experimental and density functional theory studies. Opt. Mater. 2019, 90, 273–280. [Google Scholar] [CrossRef]
  70. Kim, H.; Bin, Y.; Karthick, S.; Hemalatha, K.; Raj, C.J.; Venkatesan, S.; Park, S.; Vijayakumar, G. Natural dye extracted from Rhododendron species flowers as a photosensitizer in dye sensitized solar cell. Int. J. Electrochem. Sci. 2013, 8, 6734–6743. [Google Scholar]
  71. Al-Alwani, M.A.; Hasan, H.A.; Al-Shorgani, N.K.N.; Al-Mashaan, A.B.S. Natural dye extracted from Areca catechu fruits as a new sensitiser for dye-sensitised solar cell fabrication: Optimisation using D-Optimal design. Mater. Chem. Phys. 2020, 240, 122204. [Google Scholar] [CrossRef]
  72. Chawla, P.; Srivastava, A.; Tripathi, M. Performance of chitosan based polymer electrolyte for natural dye sensitized solar cell. Environ. Prog. Sustain. Energy 2019, 38, 630–634. [Google Scholar] [CrossRef]
  73. Munawaroh, H.; Saputri, L.; Hanif, Q.; Hidayat, R.; Wahyuningsih, S. The co-pigmentation of anthocyanin isolated from mangosteen pericarp (Garcinia mangostana L.) as Natural Dye for Dye-Sensitized Solar Cells (DSSC). IOP Conf. Ser. Mater. Sci. Eng. 2016, 107, 012061. [Google Scholar] [CrossRef]
  74. Ayalew, W.A.; Ayele, D.W. Dye-sensitized solar cells using natural dye as light-harvesting materials extracted from Acanthus sennii chiovenda flower and Euphorbia cotinifolia leaf. J. Sci. Adv. Mater. Devices 2016, 1, 488–494. [Google Scholar] [CrossRef] [Green Version]
  75. Hamadanian, M.; Safaei-Ghomi, J.; Hosseinpour, M.; Masoomi, R.; Jabbari, V. Uses of new natural dye photosensitizers in fabrication of high potential dye-sensitized solar cells (DSSCs). Mater. Sci. Semicond. Process. 2014, 27, 733–739. [Google Scholar] [CrossRef]
  76. Lim, A.; Ekanayake, P.; Lim, L.B.L.; Bandara, J.S. Co-dominant effect of selected natural dye sensitizers in DSSC performance. Spectrochim. Acta. Part A 2016, 167, 26–31. [Google Scholar] [CrossRef]
  77. Teoli, F.; Lucioli, S.; Nota, P.; Frattarelli, A.; Matteocci, F.; Di Carlo, A.; Caboni, E.; Forni, C. Role of pH and pigment concentration for natural dye-sensitized solar cells treated with anthocyanin extracts of common fruits. J. Photochem. Photobiol. A 2016, 316, 24–30. [Google Scholar] [CrossRef]
  78. Sathyajothi, S.; Jayavel, R.; Dhanemozhi, A.C. The fabrication of natural dye sensitized solar cell (DSSC) based on TiO2 using henna and beetroot dye extracts. Mater. Today Proc. 2017, 4, 668–676. [Google Scholar] [CrossRef]
  79. Sinha, D.; De, D.; Ayaz, A. Photo sensitizing and electrochemical performance analysis of mixed natural dye and nanostructured ZnO based DSSC. Sādhanā 2020, 45, 175. [Google Scholar] [CrossRef]
  80. Hossain, M.K.; Pervez, M.F.; Mia, M.; Mortuza, A.; Rahaman, M.; Karim, M.; Islam, J.M.; Ahmed, F.; Khan, M.A. Effect of dye extracting solvents and sensitization time on photovoltaic performance of natural dye sensitized solar cells. Results Phys. 2017, 7, 1516–1523. [Google Scholar] [CrossRef]
  81. Ananth, S.; Vivek, P.; Arumanayagam, T.; Murugakoothan, P. Natural dye extract of lawsonia inermis seed as photo sensitizer for titanium dioxide based dye sensitized solar cells. Spectrochim. Acta Part A 2014, 128, 420–426. [Google Scholar] [CrossRef]
  82. Ruhane, T.; Islam, M.T.; Rahaman, M.S.; Bhuiyan, M.; Islam, J.M.; Bhuiyan, T.; Khan, K.; Khan, M.A. Impact of photo electrode thickness and annealing temperature on natural dye sensitized solar cell. Sustain. Energy Technol. Assess. 2017, 20, 72–77. [Google Scholar] [CrossRef]
  83. Sultana, B.; Anwar, F.; Ashraf, M. Effect of extraction solvent/technique on the antioxidant activity of selected medicinal plant extracts. Molecules 2009, 14, 2167–2180. [Google Scholar] [CrossRef]
  84. Euterpio, M.A.; Cavaliere, C.; Capriotti, A.L.; Crescenzi, C. Extending the applicability of pressurized hot water extraction to compounds exhibiting limited water solubility by pH control: Curcumin from the turmeric rhizome. Anal. Bioanal. Chem. 2011, 401, 2977–2985. [Google Scholar] [CrossRef]
  85. Ruhane, T.A.; Islam, M.T.; Rahaman, M.S.; Bhuiyan, M.M.H.; Islam, J.M.M.; Newaz, M.K.; Khan, K.A.; Khan, M.A. Photo current enhancement of natural dye sensitized solar cell by optimizing dye extraction and its loading period. Optik 2017, 149, 174–183. [Google Scholar] [CrossRef]
  86. Bredas, J.L.; Silbey, R.; Boudreaux, D.S.; Chance, R.R. Chain-length dependence of electronic and electrochemical properties of conjugated systems: Polyacetylene, polyphenylene, polythiophene, and polypyrrole. J. Am. Chem. Soc. 1983, 105, 6555–6559. [Google Scholar] [CrossRef]
  87. Ashok Kumar, K.; Manonmani, J.; Senthilselvan, J. Effect on interfacial charge transfer resistance by hybrid co-sensitization in DSSC applications. J. Mater. Sci. Mater. Electron. 2014, 25, 5296–5301. [Google Scholar] [CrossRef]
  88. Law, C.; Pathirana, S.C.; Li, X.; Anderson, A.Y.; Barnes, P.R.F.; Listorti, A.; Ghaddar, T.H.; O′Regan, B.C. Water-Based Electrolytes for Dye-Sensitized Solar Cells. Adv. Mater. 2010, 22, 4505–4509. [Google Scholar] [CrossRef]
  89. Vaghasiya, J.V.; Nandakumar, D.K.; Yaoxin, Z.; Tan, S.C. Low toxicity environmentally friendly single component aqueous organic ionic conductors for high efficiency photoelectrochemical solar cells. J. Mater. Chem. A 2018, 6, 1009–1016. [Google Scholar] [CrossRef]
  90. Bella, F.; Galliano, S.; Falco, M.; Viscardi, G.; Barolo, C.; Grätzel, M.; Gerbaldi, C. Unveiling iodine-based electrolytes chemistry in aqueous dye-sensitized solar cells. Chem. Sci. 2016, 7, 4880–4890. [Google Scholar] [CrossRef] [Green Version]
  91. Click, K.A.; Schockman, B.M.; Dilenschneider, J.T.; McCulloch, W.D.; Garrett, B.R.; Yu, Y.; He, M.; Curtze, A.E.; Wu, Y. Bilayer dye protected aqueous photocathodes for tandem dye-sensitized solar cells. J. Phys. Chem. C 2017, 121, 8787–8795. [Google Scholar] [CrossRef]
  92. Boldrini, C.L.; Manfredi, N.; Perna, F.M.; Trifiletti, V.; Capriati, V.; Abbotto, A. Dye-Sensitized Solar Cells that use an Aqueous Choline Chloride-Based Deep Eutectic Solvent as Effective Electrolyte Solution. Energy Technol. 2017, 5, 345–353. [Google Scholar] [CrossRef]
  93. Galliano, S.; Bella, F.; Piana, G.; Giacona, G.; Viscardi, G.; Gerbaldi, C.; Grätzel, M.; Barolo, C. Finely tuning electrolytes and photoanodes in aqueous solar cells by experimental design. Sol. Energy 2018, 163, 251–255. [Google Scholar] [CrossRef]
  94. Galliano, S.; Bella, F.; Gerbaldi, C.; Falco, M.; Viscardi, G.; Grätzel, M.; Barolo, C. Photoanode/electrolyte interface stability in aqueous dye-sensitized solar cells. Energy Technol. 2017, 5, 300–311. [Google Scholar] [CrossRef]
  95. Fagiolari, L.; Bonomo, M.; Cognetti, A.; Meligrana, G.; Gerbaldi, C.; Barolo, C.; Bella, F. Photoanodes for Aqueous Solar Cells: Exploring Additives and Formulations Starting from a Commercial TiO2 Paste. ChemSusChem 2020, 13, 6562–6573. [Google Scholar] [CrossRef]
  96. Bella, F.; Galliano, S.; Falco, M.; Viscardi, G.; Barolo, C.; Grätzel, M.; Gerbaldi, C. Approaching truly sustainable solar cells by the use of water and cellulose derivatives. Green Chem. 2017, 19, 1043–1051. [Google Scholar] [CrossRef]
  97. Galliano, S.; Bella, F.; Bonomo, M.; Viscardi, G.; Gerbaldi, C.; Boschloo, G.; Barolo, C. Hydrogel Electrolytes Based on Xanthan Gum: Green Route towards Stable Dye-Sensitized Solar Cells. Nanomaterials 2020, 10, 1585. [Google Scholar] [CrossRef]
  98. Galliano, S.; Bella, F.; Bonomo, M.; Giordano, F.; Grätzel, M.; Viscardi, G.; Hagfeldt, A.; Gerbaldi, C.; Barolo, C. Xanthan-Based Hydrogel for Stable and Efficient Quasi-Solid Truly Aqueous Dye-Sensitized Solar Cell with Cobalt Mediator. Sol. RRL 2021, 5, 2000823. [Google Scholar] [CrossRef]
  99. Koo, H.-J.; Velev, O.D. Regenerable Photovoltaic Devices with a Hydrogel-Embedded Microvascular Network. Sci. Rep. 2013, 3, 2357. [Google Scholar] [CrossRef] [Green Version]
  100. Kim, J.Y.; Kim, T.H.; Kim, D.Y.; Park, N.-G.; Ahn, K.-D. Novel thixotropic gel electrolytes based on dicationic bis-imidazolium salts for quasi-solid-state dye-sensitized solar cells. J. Power Sources 2008, 175, 692–697. [Google Scholar] [CrossRef]
  101. Park, S.J.; Yoo, K.; Kim, J.-Y.; Kim, J.Y.; Lee, D.-K.; Kim, B.; Kim, H.; Kim, J.H.; Cho, J.; Ko, M.J. Water-Based Thixotropic Polymer Gel Electrolyte for Dye-Sensitized Solar Cells. ACS Nano 2013, 7, 4050–4056. [Google Scholar] [CrossRef]
  102. De Haro, J.C.; Tatsi, E.; Fagiolari, L.; Bonomo, M.; Barolo, C.; Turri, S.; Bella, F.; Griffini, G. Lignin-Based Polymer Electrolyte Membranes for Sustainable Aqueous Dye-Sensitized Solar Cells. ACS Sustain. Chem. Eng. 2021, 9, 8550–8560. [Google Scholar] [CrossRef]
  103. Yang, W.; Söderberg, M.; Eriksson, A.I.; Boschloo, G. Efficient aqueous dye-sensitized solar cell electrolytes based on a TEMPO/TEMPO+ redox couple. RSC Adv. 2015, 5, 26706–26709. [Google Scholar] [CrossRef]
  104. Lee, Y.-L.; Chang, C.-H. Efficient polysulfide electrolyte for CdS quantum dot-sensitized solar cells. J. Power Sources 2008, 185, 584–588. [Google Scholar] [CrossRef]
  105. Kamenan, K.A.; Jagadeesh, A.; Assanvo, E.F.; Soman, S.; Unni, K.N. Natural rubber (Hevea Brasiliensis)-based quasi-solid electrolyte as a potential candidate for arresting recombination and improving performance in aqueous dye-sensitized solar cells. J. Mater. Sci. Mater. Electron. 2021, 32, 14207–14216. [Google Scholar] [CrossRef]
  106. Choi, H.; Jeong, B.-S.; Do, K.; Ju, M.J.; Song, K.; Ko, J. Aqueous electrolyte based dye-sensitized solar cells using organic sensitizers. New J. Chem. 2013, 37, 329–336. [Google Scholar] [CrossRef]
  107. Zhang, H.; Qiu, L.; Xu, D.; Zhang, W.; Yan, F. Performance enhancement for water based dye-sensitized solar cells via addition of ionic surfactants. J. Mater. Chem. A 2014, 2, 2221–2226. [Google Scholar] [CrossRef]
  108. Fayad, R.; Shoker, T.A.; Ghaddar, T.H. High photo-currents with a zwitterionic thiocyanate-free dye in aqueous-based dye sensitized solar cells. Dalton Trans. 2016, 45, 5622–5628. [Google Scholar] [CrossRef]
  109. Kato, R.; Kato, F.; Oyaizu, K.; Nishide, H. Redox-active hydroxy-TEMPO radical immobilized in Nafion layer for an aqueous electrolyte-based and dye-sensitized solar cell. Chem. Lett. 2014, 43, 480–482. [Google Scholar] [CrossRef]
  110. Suzuka, M.; Hara, S.; Sekiguchi, T.; Oyaizu, K.; Nishide, H. Polyviologen as the charge-storage electrode of an aqueous electrolyte-and organic-based dye-sensitized solar cell. Polymer 2015, 68, 353–357. [Google Scholar] [CrossRef]
  111. Leandri, V.; Ellis, H.; Gabrielsson, E.; Sun, L.; Boschloo, G.; Hagfeldt, A. An organic hydrophilic dye for water-based dye-sensitized solar cells. Phys. Chem. Chem. Phys. 2014, 16, 19964–19971. [Google Scholar] [CrossRef]
  112. Sonigara, K.K.; Vaghasiya, J.V.; Machhi, H.K.; Prasad, J.; Gibaud, A.; Soni, S.S. Anisotropic One-Dimensional Aqueous Polymer Gel Electrolyte for Photoelectrochemical Devices: Improvement in Hydrophobic TiO2–Dye/Electrolyte Interface. ACS Appl. Energy Mater. 2018, 1, 3665–3673. [Google Scholar] [CrossRef]
  113. Yogananda, K.; Ramasamy, E.; Kumar, S.; Kumar, S.V.; Rani, M.N.; Rangappa, D. Novel rice starch based aqueous gel electrolyte for dye sensitized solar cell application. Mater. Today Proc. 2017, 4, 12238–12244. [Google Scholar] [CrossRef]
  114. Saito, H.; Uegusa, S.; Murakami, T.N.; Kawashima, N.; Miyasaka, T. Fabrication and efficiency enhancement of water-based dye-sensitized solar cells by interfacial activation of TiO2 mesopores. Electrochemistry 2004, 72, 310–316. [Google Scholar] [CrossRef] [Green Version]
  115. Murakami, T.N.; Saito, H.; Uegusa, S.; Kawashima, N.; Miyasaka, T. Water-based dye-sensitized solar cells: Interfacial activation of TiO2 mesopores in contact with aqueous electrolyte for efficiency development. Chem. Lett. 2003, 32, 1154–1155. [Google Scholar] [CrossRef]
  116. Bella, F.; Galliano, S.; Piana, G.; Giacona, G.; Viscardi, G.; Grätzel, M.; Barolo, C.; Gerbaldi, C. Boosting the efficiency of aqueous solar cells: A photoelectrochemical estimation on the effectiveness of TiCl4 treatment. Electrochim. Acta 2019, 302, 31–37. [Google Scholar] [CrossRef]
  117. Pham, B.; Willinger, D.; McMillan, N.K.; Roye, J.; Burnett, W.; D’Achille, A.; Coffer, J.L.; Sherman, B.D. Tin (IV) oxide nanoparticulate films for aqueous dye-sensitized solar cells. Sol. Energy 2021, 224, 984–991. [Google Scholar] [CrossRef]
  118. Kim, J.-H.; Park, S.-Y.; Lim, D.-H.; Lim, S.-Y.; Choi, J.; Koo, H.-J. Eco-Friendly Dye-Sensitized Solar Cells Based on Water-Electrolytes and Chlorophyll. Materials 2021, 14, 2150. [Google Scholar] [CrossRef]
  119. Kang, T.-S.; Chun, K.-H.; Hong, J.S.; Moon, S.-H.; Kim, K.-J. Enhanced Stability of Photocurrent-Voltage Curves in Ru(II)-Dye-Sensitized Nanocrystalline TiO2 Electrodes with Carboxylic Acids. J. Electrochem. Soc. 2000, 147, 3049. [Google Scholar] [CrossRef]
  120. Lu, H.-L.; Lee, Y.-H.; Huang, S.-T.; Su, C.; Yang, T.C.K. Influences of water in bis-benzimidazole-derivative electrolyte additives to the degradation of the dye-sensitized solar cells. Sol. Energy Mater. Sol. Cells 2011, 95, 158–162. [Google Scholar] [CrossRef]
  121. Macht, B.; Turrion, M.; Barkschat, A.; Salvador, P.; Ellmer, K.; Tributsch, H. Patterns of efficiency and degradation in dye sensitization solar cells measured with imaging techniques. Sol. Energy Mater. Sol. Cells 2002, 73, 163–173. [Google Scholar] [CrossRef]
  122. Jung, Y.-S.; Yoo, B.; Lim, M.K.; Lee, S.Y.; Kim, K.-J. Effect of Triton X-100 in water-added electrolytes on the performance of dye-sensitized solar cells. Electrochim. Acta 2009, 54, 6286–6291. [Google Scholar] [CrossRef]
  123. Weidmann, J.; Dittrich, T.; Konstantinova, E.; Lauermann, I.; Uhlendorf, I.; Koch, F. Influence of oxygen and water related surface defects on the dye sensitized TiO2 solar cell. Sol. Energy Mater. Sol. Cells 1999, 56, 153–165. [Google Scholar] [CrossRef]
  124. Liu, Y.; Hagfeldt, A.; Xiao, X.-R.; Lindquist, S.-E. Investigation of influence of redox species on the interfacial energetics of a dye-sensitized nanoporous TiO2 solar cell. Sol. Energy Mater. Sol. Cells 1998, 55, 267–281. [Google Scholar] [CrossRef]
  125. Zhang, Z.; Chen, P.; Murakami, T.N.; Zakeeruddin, S.M.; Grätzel, M. The 2,2,6,6-tetramethyl-1-piperidinyloxy radical: An efficient, iodine-free redox mediator for dye-sensitized solar cells. Adv. Funct. Mater. 2008, 18, 341–346. [Google Scholar] [CrossRef]
  126. Yang, W.; Vlachopoulos, N.; Hao, Y.; Hagfeldt, A.; Boschloo, G. Efficient dye regeneration at low driving force achieved in triphenylamine dye LEG4 and TEMPO redox mediator based dye-sensitized solar cells. Phys. Chem. Chem. Phys. 2015, 17, 15868–15875. [Google Scholar] [CrossRef]
  127. Choi, H.; Han, J.; Kang, M.-S.; Song, K.; Ko, J. Aqueous Electrolytes Based Dye-sensitized Solar Cells using I/I3 Redox Couple to Achieve≥ 4% Power Conversion Efficiency. Bull. Korean Chem. Soc. 2014, 35, 1433–1439. [Google Scholar] [CrossRef] [Green Version]
  128. Law, C.; Moudam, O.; Villarroya-Lidon, S.; O’Regan, B. Managing wetting behavior and collection efficiency in photoelectrochemical devices based on water electrolytes; improvement in efficiency of water/iodide dye sensitised cells to 4%. J. Mater. Chem. 2012, 22, 23387–23394. [Google Scholar] [CrossRef]
  129. Meen, T.-H.; Jhuo, Y.-T.; Chao, S.-M.; Lin, N.-Y.; Ji, L.-W.; Tsai, J.-K.; Wu, T.-C.; Chen, W.-R.; Water, W.; Huang, C.-J. Effect of TiO2 nanotubes with TiCl4 treatment on the photoelectrode of dye-sensitized solar cells. Nanoscale Res. Lett. 2012, 7, 579. [Google Scholar] [CrossRef] [Green Version]
  130. Lee, S.-W.; Ahn, K.-S.; Zhu, K.; Neale, N.R.; Frank, A.J. Effects of TiCl4 treatment of nanoporous TiO2 films on morphology, light harvesting, and charge-carrier dynamics in dye-sensitized solar cells. J. Phys. Chem. C 2012, 116, 21285–21290. [Google Scholar] [CrossRef]
  131. Sommeling, P.; O’Regan, B.C.; Haswell, R.; Smit, H.; Bakker, N.; Smits, J.; Kroon, J.M.; Van Roosmalen, J. Influence of a TiCl4 post-treatment on nanocrystalline TiO2 films in dye-sensitized solar cells. J. Phys. Chem. B 2006, 110, 19191–19197. [Google Scholar] [CrossRef]
  132. O’Regan, B.C.; Durrant, J.R.; Sommeling, P.M.; Bakker, N.J. Influence of the TiCl4 treatment on nanocrystalline TiO2 films in dye-sensitized solar cells. 2. Charge density, band edge shifts, and quantification of recombination losses at short circuit. J. Phys. Chem. C 2007, 111, 14001–14010. [Google Scholar] [CrossRef]
  133. Salgado-Castro, K.; Lijanova, I.V.; Jaramillo-Vigueras, D.; Castillo-Cervantes, J.N. Effect of the TiO2 Anchoring of a Hydrophobic Ionic Liquid in a Fully Aqueous DSSC. IEEE J. Photovolt. 2019, 9, 1708–1715. [Google Scholar] [CrossRef]
  134. Gu, P.; Yang, D.; Zhu, X.; Sun, H.; Wangyang, P.; Li, J.; Tian, H. Influence of electrolyte proportion on the performance of dye-sensitized solar cells. AIP Adv. 2017, 7, 105219. [Google Scholar] [CrossRef]
  135. Lai, W.H.; Su, Y.H.; Teoh, L.G.; Hon, M.H. Commercial and natural dyes as photosensitizers for a water-based dye-sensitized solar cell loaded with gold nanoparticles. J. Photochem. Photobiol. A 2008, 195, 307–313. [Google Scholar] [CrossRef]
  136. Kokal, R.K.; Bhattacharya, S.; Cardoso, L.S.; Miranda, P.B.; Soma, V.R.; Chetti, P.; Melepurath, D.; Raavi, S.S.K. Low cost ‘green’dye sensitized solar cells based on New Fuchsin dye with aqueous electrolyte and platinum-free counter electrodes. Sol. Energy 2019, 188, 913–923. [Google Scholar] [CrossRef]
Figure 1. Chemical structures of (a) anthocyanins, (b) chlorophylls, (c) carotenoids, and (d) betalain, reprinted with permission from [50]. Copyright 2019 Springer Nature.
Figure 1. Chemical structures of (a) anthocyanins, (b) chlorophylls, (c) carotenoids, and (d) betalain, reprinted with permission from [50]. Copyright 2019 Springer Nature.
Energies 15 00219 g001
Figure 2. (a) Absorbance of the anthocyanin extracted from Areca catechu depending on the types of extracting solvents. Wavelength region of adsorption is from 400 to 800 nm. Reprinted with permission from [71]. Copyright 2020 Elsevier. (b) HOMO and LUMO levels of the chlorophyll extracted from PLs depending on the pH of the extracting solvent. Reprinted with permission from [58]. Copyright 2015 Elsevier.
Figure 2. (a) Absorbance of the anthocyanin extracted from Areca catechu depending on the types of extracting solvents. Wavelength region of adsorption is from 400 to 800 nm. Reprinted with permission from [71]. Copyright 2020 Elsevier. (b) HOMO and LUMO levels of the chlorophyll extracted from PLs depending on the pH of the extracting solvent. Reprinted with permission from [58]. Copyright 2015 Elsevier.
Energies 15 00219 g002
Figure 3. (a) UV-vis absorption spectra of the natural dyes from bermudagrass (anthocyanin) and cactus (chlorophyll), and their mixture. Reprinted with permission from [59]. Copyright 2016 Elsevier. (b) Energy band schematic of a DSSC containing mixed natural dyes. (c) J-V curve of DSSCs sensitized by chlorophyll, anthocyanin, and their mixture. Reprinted with permission from [59]. Copyright 2016 Elsevier. (d) Fabrication process of a layered co-sensitized solar cell. Reprinted with permission from [42]. Copyright 2013 Elsevier.
Figure 3. (a) UV-vis absorption spectra of the natural dyes from bermudagrass (anthocyanin) and cactus (chlorophyll), and their mixture. Reprinted with permission from [59]. Copyright 2016 Elsevier. (b) Energy band schematic of a DSSC containing mixed natural dyes. (c) J-V curve of DSSCs sensitized by chlorophyll, anthocyanin, and their mixture. Reprinted with permission from [59]. Copyright 2016 Elsevier. (d) Fabrication process of a layered co-sensitized solar cell. Reprinted with permission from [42]. Copyright 2013 Elsevier.
Energies 15 00219 g003
Figure 4. (a) J-V curves of the aqueous electrolyte-based DSSCs according to the content of water in the electrolyte based on the water–MPN mixture. Reprinted with permission from [88]. Copyright 2010 Wiley. (b) Stability in photovoltaic performance of DSSCs based on aqueous electrolytes (black squares) and hydrogel electrolytes (red circles). Reprinted with permission from [97]. Copyright 2020 MDPI.
Figure 4. (a) J-V curves of the aqueous electrolyte-based DSSCs according to the content of water in the electrolyte based on the water–MPN mixture. Reprinted with permission from [88]. Copyright 2010 Wiley. (b) Stability in photovoltaic performance of DSSCs based on aqueous electrolytes (black squares) and hydrogel electrolytes (red circles). Reprinted with permission from [97]. Copyright 2020 MDPI.
Energies 15 00219 g004
Figure 5. (a) Redox reaction of the TEMPOL/TEMPOL+. (b) J-V curves of the DSSCs based on aqueous electrolytes containing 5 M NaI and 0.01 M I2, with or without TiCl4 treatment on photoanodes. Reprinted with permission from [116]. Copyright 2019 Elsevier.
Figure 5. (a) Redox reaction of the TEMPOL/TEMPOL+. (b) J-V curves of the DSSCs based on aqueous electrolytes containing 5 M NaI and 0.01 M I2, with or without TiCl4 treatment on photoanodes. Reprinted with permission from [116]. Copyright 2019 Elsevier.
Energies 15 00219 g005
Figure 6. (a) Scheme of general DSSCs containing a synthesized dye and organic electrolyte and eco-friendly DSSCs containing a natural dye and aqueous electrolyte. (b) DSSC assembled with an aqueous gel electrolyte mimicking a leaf vein. Reprinted with permission from [99]. Copyright 2013 Elsevier.
Figure 6. (a) Scheme of general DSSCs containing a synthesized dye and organic electrolyte and eco-friendly DSSCs containing a natural dye and aqueous electrolyte. (b) DSSC assembled with an aqueous gel electrolyte mimicking a leaf vein. Reprinted with permission from [99]. Copyright 2013 Elsevier.
Energies 15 00219 g006
Table 1. Melting point, boiling point, vapor pressure, and viscosity of organic solvents popularly used. All parameter values at 25 °C unless otherwise indicated.
Table 1. Melting point, boiling point, vapor pressure, and viscosity of organic solvents popularly used. All parameter values at 25 °C unless otherwise indicated.
Organic SolventMelting Point (°C)Boiling Point (°C)Vapor Pressure
(Torr, at 25 °C)
Viscosity
(cP, at 25 °C)
Acetonitrile (ACN)−4581.688.80.334
3-methoxypropionitrile (MPN)−62.91641.72 (30 °C)2.5
Valerontrile−961392.7940.78 (19 °C)
3-methyl-2-oxazolidinone (NMO)15880.008772.5
Ethylene carbonate (EC)362380.009890
Propylene carbonate (PC)−492410.0582.5
γ-butyrolactone (GBL)−442040.45 (20 °C)1.7
N-methyl-2-pyrrolidone (NMP)−242030.3421.65
Table 2. Types and sources of natural dyes, photoanodes, electrolytes, and cathodes used for the natural dye based-DSSCs and their corresponding photovoltaic performance.
Table 2. Types and sources of natural dyes, photoanodes, electrolytes, and cathodes used for the natural dye based-DSSCs and their corresponding photovoltaic performance.
Natural DyeDye SourcesPhotoanodeElectrolyteCathodePCE (%)Ref.
ChlorophyllSpinachTiO2 treated with TiCl4I/I3 in EGCarbon0.56[51]
SpinachTiO2I/I3 in EGGraphite0.398[52]
SpinachZnOI/I3Carbon0.1312[49]
SpinachTiO2I/I3Graphite0.49[53]
SpinachTiO2I/I3 in ACNPt0.29[54]
SpinachTiO2I/I3Pt0.1712[55]
NeemZnOI/I3 in EG/ACNStainless foil0.13[56]
WormwoodTiO2I/I3 in ACNPt0.538[29]
IpomoeaTiO2I/I3Pt0.278[57]
Lemon leavesTiO2I/I3 in ACNPt0.04[41]
Papaya leavesTiO2I/I3 in ACNPt0.07[58]
Bermuda grassTiO2I/I3 in t-BuOH/ACNPt0.113[59]
Papaya peelsZnOI/I3FTO0.017[60]
Pandan leavesTiO2NaI in PVDF-HFPPt0.51[61]
Pterocarpus Indicus WilldTiO2I/I3 in EGCarbon0.0232[62]
AnthocyaninMelinjo skinTiO2I/I3Pt0.036[63]
Purple cabbageZnOI/I3Carbon0.102[64]
Purple cabbageZnOI/I3Carbon0.1015[49]
Siahkooti peelTiO2I/I3 in ACNPt0.32[45]
RaspberriesTiO2I/I3 in ACNPt1.5[41]
Mangosteen peelTiO2 treated with TiCl4T2/T in ACNMangosteen peel carbon (MPC)2.63[65]
Dragon fruitTiO2I/I3Pt0.22[66]
CuminiTiO2I/I3 in PEO:PEGPt0.07[67]
PomegranateTiO2I/I3:PEG in ACNPt0.028[68]
Red cabbageTiO2I/I3 in PEGCarbon0.024[40]
Fistula flowerTiO2I/I3 in ACNPt0.21[69]
Rhododendron flower (red)TiO2I/I3Pt0.33[70]
Canarium odontophyllumTiO2I/I3 in EG/ACNPt0.96[42]
Areca catechuTiO2I/I3Pt0.38[71]
PomegranateTiO2-WO3I/I3 in chitosanPt1.8[72]
Mangosteen peelTiO2I/I3Pt0.199[73]
Black riceTiO2NaI in PVDF-HFPPt0.56[61]
RosellaTiO2I/I3 in EGPt0.37[43]
Onion peelTiO2I/I3Pt0.0647[55]
Acanthus sennii chiovenda flowerTiO2I/I3 in gel electrolytePEDOT0.15[74]
Consolida jacisTiO2I/I3 in ACN/VNSteel mesh0.6[75]
Petals of lxora coccineaTiO2I/I3 in ACN/ECPt0.76[76]
BlueberryTiO2I/I3Pt0.69[77]
BetalainCactusTiO2I/I3 in t-BuOH/ACNPt0.674[59]
Yellow sweet potatoTiO2I/I3Pt0.057[63]
BeetrootTiO2 treated with TiCl4I/I3 in EGCarbon0.49[51]
BeetrootZnOI/I3Carbon0.179[64]
BeetrootTiO2I/I3 in ACN/ECGraphite1.3[78]
Turmeric stemZnOI/I3Carbon0.3045[79]
TurmericTiO2I/I3 in EGCarbon0.33[80]
Bougainvillea spectabilisTiO2I/I3 in ACN/ECPt0.21[76]
LawsoneLawsonia inermisTiO2I/I3Pt1.47[81]
Henna leavesTiO2I/I3 in ACNGraphite1.08[78]
CarotenoidOrange peelTiO2I/I3:PEG in ACNPt0.005[68]
CurcuminTurmeric rootTiO2I/I3 in EGCarbon0.11[82]
IndigoIndigofera tinctoriaTiO2 treated with TiCl4I/I3 in MPNPt0.114[14]
Table 3. Combination and ratio of natural dyes and the photovoltaic parameters.
Table 3. Combination and ratio of natural dyes and the photovoltaic parameters.
Combination of Natural DyeRatioVOC (V)JSC (mA/cm2)FFPCE (%)Ref.
Anthocyanin + Chlorophyll1:10.5321.450.670.5175[49]
Anthocyanin + Chlorophyll2:10.6752.550.671.15[29]
Anthocyanin + Chlorophyll1:10.663.160.621.29
Anthocyanin + Chlorophyll1:10.472.630.580.72[61]
Anthocyanin + Betalain1:10.561.120.60.3824[64]
Anthocyanin + Anthocyanin1:10.386.260.471.13[42]
Betalain + Chlorophyll1:10.4954.970.461.139[59]
Betalain + Chlorophyll4:10.3864.740.540.99[51]
Anthocyanin + Betalain + Chlorophyll1:1:10.531.650.680.602[79]
Table 4. Photovoltaic performance of DSSCs with aqueous electrolytes.
Table 4. Photovoltaic performance of DSSCs with aqueous electrolytes.
DyeElectrolyteContent of Water (%)PhotoanodeCathodePCE (%)Ref.
LEG40.15 M TEMPO, 0.05 M TEMPOBF4, LiClO4, 0.2 M NMBI in H2O100TiO2Pt4.14[103]
D1315.5 M KI, 0.05 M I2 in H2O100TiO2Pt0.73[90]
BH22 M NaI, 0.02 M I2, 0.5 M GuSCN in an aqueous solution saturated CDCA100TiO2NiO0.056[91]
CdS QD0.5 M Na2S, 2 M S, 0.2 M KCl in MeOH:H2O30TiO2Pt1.15[104]
N719/Z9072 M NaI, 0.02 M I2, 0.5 M GuSCN in H2O100TiO2Pt0.68[105]
2 M NaI, 0.02 M I2, 0.5 M GuSCN, and 1 g natural rubber in H2O0.46
JK-2592 M PMMI, 0.05 M I2, 0.1 m GuSCN, 0.5 M TBP, 1% Triton X-100 in H2O100TiO2Pt1.16[106]
JK-2622.1
N7192 M NaI, 0.2 M I2, 0.1 M GuSCN in H2O100TiO2Pt2.51[107]
2 M NaI, 0.2 M I2, 0.1 M GuSCN, and 0.2 wt.% FC-134 in H2O3.69
T169T-/DS in the presence of H2O2100TiO2PEDOT4.5[108]
D2051 M TEMPOL in an aqueous 1 M NaBF4 solution100TiO2Nafion2.1[109]
D1310.5 M TEMPOL in an aqueous 0.5 M NaCl solution in the presence of 0.1 M H2O2100TiO2Pt1.3[110]
TG62 M PMMI, 0.05 M I2, 0.1 M GuSCN, 0.5 M TBP in MPN:H2O0TiO2Pt5.5[88]
604.5
1002.4
D1310.5 M NaI, 25 mM I2 in H2O100TiO2Pt0.2[94]
D2050.1
D1490.14
V352 M KI, 0.01 M I2 in an aqueous solution saturated CDCA100TiO2PEDOT3.01[111]
SK32 M LiI, 0.02 M I2, 1 M GuSCN in H2O100TiO2Pt1.27[112]
N7191 M LiI, 0.02 M I2 in H2O100TiO2Pt0.1[113]
1 M LiI, 0.02 M I2, Rice starch in H2O0.35
D1315 M NaI, 0.03 M I2 in an aqueous solution saturated CDCA100TiO2Pt2.44[97]
5 M NaI, 0.03 M I2, 5 wt.% of XG in an aqueous solution saturated CDCA2.23
MK20.21 M Co(bpy)3Cl2, 0.07 M Co(bpy)3Cl3, 1.5 wt.% of XG in H2O100TiO2Pt4.47[98]
D1315.5 M KI, 0.05 M I2, 5.5 wt.% CMC in H2O100TiO2Pt0.72[96]
4.5 M NaI, 0.05 M I2, 5.5 wt.% CMC in H2O0.61
N30.5 M KI, 0.025 M I2 in H2O100TiO2Pt0.6[114,115]
N7190.5
N30.5 M KI, 0.025 M I2 in 35% aqueous ethanol solution651.3
N7191.1
D1315 M NaI, 0.01 M I2 in an aqueous solution saturated CDCA100TiO2Pt2.37[116]
N30.5 M LiI, 0.025 M I2 in H2O100SnO2/TiO2Pt0.66[117]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kim, J.-H.; Kim, D.-H.; So, J.-H.; Koo, H.-J. Toward Eco-Friendly Dye-Sensitized Solar Cells (DSSCs): Natural Dyes and Aqueous Electrolytes. Energies 2022, 15, 219. https://doi.org/10.3390/en15010219

AMA Style

Kim J-H, Kim D-H, So J-H, Koo H-J. Toward Eco-Friendly Dye-Sensitized Solar Cells (DSSCs): Natural Dyes and Aqueous Electrolytes. Energies. 2022; 15(1):219. https://doi.org/10.3390/en15010219

Chicago/Turabian Style

Kim, Ji-Hye, Dong-Hyuk Kim, Ju-Hee So, and Hyung-Jun Koo. 2022. "Toward Eco-Friendly Dye-Sensitized Solar Cells (DSSCs): Natural Dyes and Aqueous Electrolytes" Energies 15, no. 1: 219. https://doi.org/10.3390/en15010219

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop