Next Article in Journal
One-Pot Reactions of Triethyl Orthoformate with Amines
Previous Article in Journal
Review on the Synthesis, Recyclability, Degradability, Self-Healability and Potential Applications of Reversible Imine Bond Containing Biobased Epoxy Thermosets
Previous Article in Special Issue
A New Pd-Based Catalytic System for the Reductive Carbonylation of Nitrobenzene to Form N-(4-hydroxyphenyl)acetamide Selectively in One Pot
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Chlorophyll Sensitization of TiO2: A Mini-Review

by
Maria E. K. Fuziki
1,*,
Angelo M. Tusset
2,
Onélia A. A. dos Santos
1 and
Giane G. Lenzi
2
1
Department of Chemical Engineering, State University of Maringá, Colombo Av. 5790, Maringá 87020-900, Paraná, Brazil
2
Department of Production Engineering, Federal University of Technology-Paraná, Doutor Washington Subtil Chueire St. 330, Ponta Grossa 84017-220, Paraná, Brazil
*
Author to whom correspondence should be addressed.
Reactions 2023, 4(4), 766-778; https://doi.org/10.3390/reactions4040044
Submission received: 12 September 2023 / Revised: 4 November 2023 / Accepted: 28 November 2023 / Published: 1 December 2023

Abstract

:
Recent studies have shown that chlorophyll sensitization can improve the performance of semiconductors like TiO2 in photocatalytic reactions and light-harvesting technologies, such as solar cells. Faced with the search for renewable energy sources and sustainable technologies, the application of this natural pigment has been gaining prominence. The present work addresses some of the main possibilities of chlorophyll-TiO2 combination, presenting the most relevant aspects affecting chlorophyll extraction and TiO2 sensitization.

1. Introduction

Titanium dioxide (TiO2) is a semiconductor widely used in photocatalytic and photoelectrochemical processes, since the first works describing reactions involving TiO2 and UV radiation [1] and the Fujishima–Honda effect were reported in the 1960s [2]. In the early 1970s, Fujishima and Honda investigated the behavior of the rutile form of TiO2 under light irradiation, discussing the similarities between the observed phenomenon and the initial stages of photosynthesis, focusing on studying the mechanism of the latter [3]. Over the years, the number of studies about TiO2-based materials combined with light irradiation have multiplied, spreading to different areas of application, such as the oxidation of organic compounds [4], the degradation of pollutants such as pharmaceuticals and pesticides [5,6], the removal of heavy metals from water [7], bactericidal activity [8], hydrogen production [9,10,11], and CO2 reduction for solar fuel synthesis [12]. Notably, this last example has contributed to a strengthening of the aspiration for an artificial photosynthesis process, which, similarly to natural photosynthesis, consumes simple products, such as water and carbon dioxide, to produce energetic substances (e.g., H2, CO, CH3OH, CH4) using sunlight, which are also known as solar fuels [13,14].
The cited processes are based on TiO2’s ability to form electron (e)/hole (h+) pairs under UV irradiation. When the energy of the photons absorbed by TiO2 is more significant than its band gap, the electrons in the valence band (VB) are promoted to the conduction band (CB) of the semiconductor, generating holes in the VB, and electrons in the CB [15,16]. The photogenerated e/h+ pairs can recombine inside the photocatalyst particle or migrate to the surface and undergo recombination. If this does not occur, e/h+ pairs at the surface can promote reduction or oxidation reactions of adsorbed species [15,17]. This phenomenon is the basis of the various photocatalytic processes. As it can generate highly active radical species, e.g., hydroxyl radicals, photocatalysis can be classified as an advanced oxidation process (AOP) [16].
Despite its low toxicity, outstanding activity, and considerable chemical stability [18], titanium dioxide is only active under UV light, restricting TiO2 applications combined with solar radiation, which includes predominantly visible light and less than 5% of UV radiation [19]. As one of the possible alternatives, dye sensitization of wide band-gap semiconductors (>3.0 eV), such as TiO2, has been the subject of different studies for improved solar light harvesting technologies [20]. In dye sensitization, the dye molecule bonded to the semiconductor surface injects electrons into the conduction band of the semiconductor upon photoexcitation, as described in Equations (1) and (2) [21,22].
D y e   h v   D y e *
D y e *   T i O 2   D y e + + e C B ( T i O 2 )
Based on this principle, one of the third-generation solar cell technologies has gained ground, driven by growing energy demand and the search for renewable energy sources: dye-sensitized solar cells (DSSCs) [23,24]. This type of device aims to convert sunlight into electricity, consisting of a substrate made up of a glass conductor, a dye-sensitized semiconductor (metal oxide), and a catalyst counter electrode, separated from the other electrode using an electrolyte solution [24]. Although this technology still suffers from limitations in terms of efficiency and long-term stability, there is an excellent expectation regarding its development, especially concerning the use of new dyes that improve its performance [25]. In this context, natural dye extracts have gained attention due to the abundance, low cost, and environmentally friendly nature of the raw material [25,26,27,28].
Among natural dyes, chlorophyll has undoubtedly shown exciting results. As reported by [29], spinach extract produced the best efficiency results among the different natural extracts used in sensitizing TiO2 solar cells. Compared to the other natural dyes (black rice, dragon fruit, red cabbage, and blends), the spinach UV-Vis absorption spectrum presented the highest absorption peak at approximately 662 nm [29], which can be associated with chlorophyll-a. Haghighatzadeh, in a study about phenol photocatalytic degradation under visible light irradiation, observed that TiO2 nanoparticles sensitized with chlorophyll promoted higher percentages of degradation (85%) than those sensitized with curcumin (75%) [30]. Thus, TiO2 sensitized with chlorophyll has gained space in DSSC [31] and photocatalysis applications, including pollutant degradation [32], CO2 reduction [33], and even artificial photosynthesis processes involving light harvesting and oxygen production [34]. In this context, the present work aims to present some of the main practical aspects involved in the extraction of chlorophyll and its use in TiO2 sensitization for different applications described in the literature.

2. Chlorophyll

Chlorophyll is a natural pigment of the porphyrin class, which has a Mg2+ ion coordinated to the four rings [35], as shown in Figure 1. The Mg2+ ion in the molecule plays a vital role in the light absorption phenomenon, being essential for the excited state of the molecule and affecting the efficiency of the excitation transfer between chlorophyll molecules in the chloroplast [35,36,37], giving chlorophyll a prominent position in photosynthesis and promoting solar energy conversion into chemical energy [38]. The green color of chlorophyll pigments is due to their high absorption in the red and blue regions of the light spectrum [39]. In the absorption spectrum, the ranges of 350 nm to 480 nm can be attributed to the charge transfer transitions of the porphyrin and the Mg ion [38].
While chlorophyll-a is a pigment common to all photosynthetic plants, chlorophyll-b is characteristic of algae and vascular plants [38]. As for the molecular structure, the two pigments are distinguished by the presence of a methyl group (chlorophyll-a, Figure 1a) or an aldehyde group (chlorophyll-b, Figure 1b) at position 3 [35]. Green plants usually contain both chlorophyll-a and chlorophyll-b, the former being the major pigment [35]. There are reports of an increase in the proportion of chlorophyll-b in shade plants, given that this pigment would be more effective in absorbing low-intensity light [35]. Despite the slight differences between the pigments, it is quite common to use extracts containing both to sensitize TiO2, whether in extracts obtained from plants [19] or other sources, such as Spirulina [40]. Even so, there are reports of the isolated use of chlorophyll-a [41] and chlorophyll-b [42] as TiO2 modifiers.
Chlorophyll-a’s two maximum absorption peaks are at approximately 432 and 670 nm [35]. However, the chlorophyll absorption spectrum may vary slightly depending on the solvent [35]. Shen et al., for example, reported a variation in the chlorophyll-a absorption maxima in different solvents, changing from 420 nm and 661 nm in chloroform to 421 nm and 667 nm in ethanol and 440 nm and 673 nm in phosphate-buffered saline (0.01 M, pH 7) [34]. This type of phenomenon is closely related to the polarity of the solvent, and it is common to observe a red shift as the polarity of the solvent increases [43,44].
Through a theoretical study using density functional theory (DFT) and time-dependent DFT (TD-DFT), Faiz et al. concluded that the solvent can also reduce the LUMO–HOMO band gap and affect the light-harvesting energy (LHE) [45]. Among the results obtained, the authors observed that water could improve chlorophyll’s performance in injecting electrons, even though chlorophyll is not soluble in water [45]. Similarly, Sabagh et al. reported that the solvent improves the LHE in a comparative study between water and the gas phase [46].

2.1. Chlorophyll-a and Chlorophyll-b Concentration Estimations

The concentration of chlorophyll-a and chlorophyll-b can be determined directly and simultaneously using spectrophotometry through calculations considering a system of two equations [35]. The equations may vary slightly since the solvent can affect the absorption spectrum [35]. Krishnan et al., for example, considered the chlorophyll optical densities ( O D ) in the extract—which were calculated by subtracting the absorbance at 750 nm from absorbances at 647 nm ( O D 647 ) and 664 nm ( O D 664 )—to estimate the concentrations of chlorophyll-a and chlorophyll-b in mg L−1 [19]. Pai et al., in turn, used the absorbance value to estimate the concentration of the two pigments [39]. Some examples are provided in Table 1.

2.2. Extraction of Chlorophyll

Natural chlorophyll extracts can be produced from different raw materials including plants [32,39] and microorganisms [22]. Among the plant species, we can mention leaves of fresh spinach [19], pandan [48], weeds such as Chromolaena odorata [39], and parsley [30]. Among microorganisms, the cyanobacteria stand out, among which we can mention Spirulina sp. [22].
The chlorophyll molecule is composed of a hydrophilic part and a hydrophobic part [39], and its extraction is commonly performed by using organic solvents such as methanol [22], ethanol [40], acetone [19], and petroleum ether [39], among others. Najihah et al. observed that polar organic solvents tend to promote better chlorophyll extraction (acetone > ethanol > methanol > acetic acid > acetonitrile) than non-polar solvents (hexane) [49]. This result justifies the widespread use of acetone in chlorophyll extraction [29]. Krishnan and Shriwastav, for example, extracted chlorophyll from ground fresh spinach leaves (after removing their midrib) using a 90% acetone aqueous solution, which was kept in contact with the leaves for 2 h in the dark at 4 °C. The extract was centrifuged for 20 min at 3000 rpm. The authors were able to produce an extract containing 0.39 ± 0.05 mg of chlorophyll per gram of spinach [19].
Other commonly used solvents are ethanol and methanol. Al-Alwani et al., for example, reported that ethanol showed the best performance in extracting chlorophyll from pandan leaves (P. amaryllifolius) compared to methanol, chloroform, ethyl ether, and acetonitrile [48]. Kathiravna et al. [22] extracted chlorophyll from a cyanobacteria Spirulina sp. using a 90% methanol solution and centrifugation as the separation method.
Regardless of the raw material or solvent used, the extraction process usually follows similar steps under mild conditions and relatively simple procedures to ensure the extraction of chlorophyll by the solvent. The initial steps involved preparing the raw material, generally with the reduction of the sample through crushing and grinding [19,22,30,49]. In some cases, it is also necessary to preliminarily remove the midrib of plant leaves [19] or carry out a drying step [48]. Once the sample is prepared, it proceeds to the next step, in which the contact between the raw material and the solvent is promoted for a defined time, which can vary from a few hours (1 h [22] or 2 h [19]) or days (from 24 h [29] up to 1 week [30,48,50]). The sample can be sonicated during this period to promote extraction [40]. Then, the extract is separated from solid waste using centrifugation [19,22] or filtration [29,30,40,48,49]. In some cases, it is also possible to increase the extract concentration in a rotary evaporator [29,48]. Some extraction conditions are summarized in Table 2.
Even though these are the most common procedures, there are alternative processes that can differ significantly from the extraction methods described, such as the procedure followed by Phongamwong et al., who obtained the chlorophyll extract after a short incubation period (2 min) of Spirulina at 70 °C, at the end of which the mixture was centrifuged and the supernatant collected [51].
Chlorophyll tends to be unstable, suffering from the action of temperature, light, oxygen, or other chemical reactions [35]. After the extraction, it is essential to take care of the extract’s storage conditions to avoid its degradation, whether that be by keeping it at low temperatures (4 °C [19,22,30]) or by protecting it from exposure to atmospheric air [48] or light [22,29,39,40,48] to prevent autoxidation [40]. Furthermore, it is necessary to consider that the extract obtained from plants and microorganisms may contain other organic compounds, such as sugars and amino acids, which may lead to chlorophyll degradation during storage or even the detachment of the dye from the TiO2 surface [49].
Phongamwong et al. did not perform the chlorophyll extraction. Still, they incorporated Spirulina directly into the TiO2 using the incipient wetness impregnation method, using deionized water to disperse the ground Spirulina and then adding it to the N-TiO2, which was kept under constant stirring at 40 °C until the complete evaporation of water [33]. In a later work, the authors compared the incorporation of Spirulina (Sp) and chlorophyll (Chl) to P25 using the incipient wetness impregnation method. They observed that, although both contributed to a significant improvement in P25 performance, the incorporation of extracted chlorophyll led to superior results. While P25 presented a rate constant of k = 8.05 ± 0.23 (10−3 min−1) in the degradation of Rhodamine B, the modified catalysts 0.5Sp/P25 and 0.5Chl/P25 presented rate constants equal to k = 23.53 ± 0.91 (10−3 min−1) and k = 60.80 ± 2.21 (10−3 min−1), respectively [51].

3. Chlorophyll Fixation on TiO2

Dye sensitization can occur through electrostatic, hydrophobic, or chemical interactions between the dye and the semiconductor surface [21]. An adequate interaction between the pigment and the semiconductor film is fundamental for improving the energy conversion and charge transfer in DSSCs [29]. The simplest and most common form of sensitization described in the literature involves promoting direct contact between the TiO2 and the chlorophyll extract for a certain period. In the case of Chl-TiO2 electrodes, it is common to dip the electrode in the dye extract [39]. An essential factor to consider is the contact time, which can take a few hours, 24 h [48,49], or longer [41]. Mahadik et al. prepared Chl-a-sensitized TiO2 nanorods by dipping the synthesized thin films horizontally in a Chl-a ethanol solution for 10–40 h [41]. As for photocatalytic processes, several works reported applying the incipient wetness impregnation method to sensitize catalysts [19]. Phongamwong et al. applied the incipient wetness impregnation method, promoting contact between the chlorophyll extract and the catalyst overnight under agitation at 40 °C until the solvent evaporated [51].
According to the molecular dynamic simulations performed by Christwardana et al. [40], chlorophyll binds to TiO2 through interactions between some of its functional groups and the O of titanium dioxide, with a predominance of polar bonds. TiO2 FTIR spectra typically show peaks at approximately 3400 cm−1 (O–H stretching), 1630 cm-1 (O–H bending mode), and 800–400 cm−1 (O–Ti–O stretching) [19,51]. Chlorophyll’s presence in TiO2-sensitized samples is normally indicated by the peaks associated with the symmetric C–H stretching vibration of methine (CH), methyl (CH3), or methylene (CH2) groups, which has already been reported by different authors to occur at 2975 cm−1 [52], 2932 cm−1 [53], or 2920 cm−1 [51]. Several authors also highlighted the peak referring to C=O stretching in the region between 1632 and 1641cm−1, which is associated with ketone or aldehyde groups, highlighting the importance of C=O for chlorophyll anchoring on TiO2. [51,52,53]. The FTIR spectra of chlorophyll extracted from different sources also tend to show a broad peak in the region at approximately 3500 cm−1 associated with the OH group, as already described by [49] (3700–3000 cm−1), [38] (3600–3400 cm−1), [52] (3296 cm−1), and [48] (3350 cm−1).
The presence of alkyl groups instead of hydroxyl or carboxyl groups in the chlorophyll molecule can impair the Chl-TiO2 interaction due to steric effects, hindering the electron transfer from the dye to the electrode [48]. The dye sensitization of TiO2 can be directly influenced by the number and distribution of carboxyl groups in the dye molecule [54]. The three possible coordination modes of carboxylate groups to metal are represented in Figure 2 [20]:
Since there are no carboxyl, carbonyl, or hydroxyl groups in the chlorophyll molecule that favor its anchorage to the surface [39,55], its fixation on TiO2 can be improved with the help of some fatty acids with a long carbon chain such as myristic acid, stearic acid, and cholic acid [55]. The fatty acids bind to the TiO2 surface via carboxylate groups, while their hydrophobic domains allow the anchoring of chlorophyll molecules [55]. The long hydrophobic part of the chlorophyll molecule facilitates its interaction with lipids and its consequent insertion into lipid bilayers [34].
Chlorophyll derivatives containing carboxy groups [56] can be a solution for binding and efficiently injecting electrons into TiO2 [57]. Another strategy to overcome this limitation is the incorporation of dopants into TiO2 [39]. Phongamwong et al., for example, added Spirulina containing chlorophyll to N-doped TiO2 catalysts, observing a synergistic effect between doping and chlorophyll loading, achieving a 21.3-fold increase in yield compared to pure TiO2 [33]. Phongamwong et al. observed that the modification of TiO2 with Spirulina led to a decrease in the catalyst’s surface area and pore volume, most likely due to the large size and low porosity of the Spirulina particles [33].

3.1. Chlorophyll-Modified TiO2 as Photocatalyst

Chlorophyll-modified TiO2 has gained space as a photocatalyst for applications in the environmental area, whether in the degradation of organic pollutants [19,32,51], bactericidal activity [58], or for energy purposes, such as in the synthesis of solar fuels from CO2 reduction [12,33]. Figure 3 presents a schematic representation of a typical photocatalytic process based on chlorophyll-sensitized TiO2. Among the pollutants usually degraded, we can cite dyes such as methylene blue [19] and rhodamine B [51,59]. However, given that the molecules of these dyes are also able to absorb visible photons and donate electrons to TiO2, it is necessary to carefully consider the contribution of the dye molecules being degraded to accurately assess the improvement produced by the addition of chlorophyll to titanium dioxide [19,51].
Furthermore, the association between chlorophyll and TiO2 has also found space in artificial photosynthesis applications. Shen et al. combined chlorophyll-a and TiO2 into a fascinating structure mimicking the thylakoid membrane [34]. By applying the synthesized chlorophyll-TiO2-liposomes, the authors achieved O2 productions that were approximately three times greater than free chloroplasts.
The improved visible light response is one of the main advantages pointed out in studies on incorporating chlorophyll into TiO2. TiO2 thin films sensitized with a cyanobacterial biomass showed approximately 25% methylene blue degradation in 140 min of visible light irradiation compared to less than 10% removal from bare TiO2 films in the same time interval, as described by Patiño-Camelo et al. [60]. Phongamwong et al. described an increase in the kinetic degradation rate of rhodamine B under visible light after P25 sensitization using chlorophyll, rising from 8.05 × 10−3 min−1 using pure P25 to 60.80 × 10−3 min−1 after dye incorporation [51]. Phongamwong et al. also reported a decrease in the band gap of Chl/P25 samples from 3.02 eV of pure P25 to 2.82–2.87 eV after chlorophyll incorporation [51]. This result was in agreement with that stated by Patiño-Camelo [60], who observed a band-gap reduction from 3.2 eV (TiO2) to 2.55 eV (sensitized TiO2).
This improvement in performance under visible light is directly related to the sensitizing role of chlorophyll, which acts as an electron donor to the conduction band of the photocatalyst the pigment is incorporated into [51]. When chlorophyll in the ground state ( C h l ) absorbs photons in the visible light region of the spectrum (around 457 nm), the electrons present in its highest occupied molecular orbital (HOMO) acquire energy to be promoted to last unoccupied molecular orbital (LUMO), resulting in a singlet excited state ( C h l *   1 , Equation (3)) [19].
C h l   h v   C h l *   1
Some of the chlorophyll electrons in the singlet state can undergo intersystem crossing ( I S C ), leading to the triplet state ( C h l *   3 , Equation (4)), or even a return to the ground state by releasing energy [19,51].
C h l *   1   I S C   C h l *   3
Chlorophyll in an excited state can donate electrons to TiO2, transferring them to the conduction band of the photocatalyst, resulting in the formation of the cationic form of chlorophyll ( C h l     + , Equation (5)) [19]. The electron transfer mechanism is represented in Figure 4.
C h l *   1       C h l     + + T i O 2   ( e C B )
The photogenerated electrons can then participate in reactions (Equations (6)–(11)), forming superoxide ( O 2 ) and hydroxyl radicals ( O H ), which can promote the degradation of various compounds [19,61].
T i O 2 e C B + O 2 T i O 2 + O 2
C h l 1 * + O 2 C h l + O 2 1
T i O 2 e C B + O 2 1 T i O 2 + O 2
O 2 + H 2 O O H + H 2 O
2 H 2 O + H 2 O H 2 O 2 + O H
H 2 O 2 h v 2 O H
However, the excessive loading of chlorophyll in the catalyst can compromise its photocatalytic activity by forming recombination centers in the catalyst. Phongamwong et al. reported a decrease in Rhodamine B degradation from 64% to 61% in 1 h of testing when the chlorophyll loading was increased from 0.5 to 1.0 wt.% [51]. There was a slight decrease in photocatalyst activity with the excessive addition of Spirulina, probably because it started acting as a recombination center for photogenerated electron–hole pairs [33].
Phongamwong et al. reported an enhanced photocatalytic stability of N-doped TiO2 catalysts after Spirulina incorporation, in contrast to non-sensitized materials, which showed a decrease in H2 production after 4 h of irradiation [33]. In the same sense, 0.5 Chl-0.1 Mg/P25 catalysts prepared by Phongamwong et al. were considerably stable even after seven cycles of photocatalysis, making it possible to maintain a degradation efficiency greater than 75% of rhodamine B. FTIR analyses carried out with used catalysts compared to fresh ones indicated that, at the end of the tests, only 88% of the chlorophyll was still present in the catalyst, and it was possible to observe this because the intensity of the peak was located at 2920 cm−1. The authors attributed the excellent stability of chlorophyll to the incorporation of Mg into the catalyst, which would have contributed to the inhibition of chlorophyll degradation [51].
Phongamwong et al. described the increased selectivity in CO2 reduction towards C2+ products. According to the authors, the greater availability of electrons from chlorophyll could probably favor chain growth during hydrocarbon formation [33].

3.2. Chlorophyll Sensitization in TiO2 Solar Cells

DSSCs can be manufactured by sensitizing TiO2 with extracts from different sources, as shown in Table 3. Figure 5 represents a typical DDSC based on chlorophyll-sensitized TiO2. The spinach extract is a typical example. Ahliha et al. prepared DSSCs with spinach dye extracted, achieving a voltage of 0.639 mV, a current of 0.33 mA, an FF equal to 0.337, and an efficiency of 0.0713% [29]. Yang et al. used spinach leaf extract to sensitize TiO2 solar cells with light-harvesting complex II (LHCII) pigment [62].
One of the critical aspects in preparing chlorophyll-based DSSCs is the promotion of the interaction between the pigment and TiO2. Amao and Kato [55] improved the chlorophyll-a anchoring to the nanocrystalline TiO2 electrodes by using three fatty acids as modifiers: myristic acid (Myr), stearic acid (Ste), and cholic acid (Cho). The photocurrent responses obtained by the authors were equal to 0.27, 0.19, and 0.19 mA cm−2 for Chl-a/Ste-TiO2, Chl-a/Myr-TiO2, and Chl-a/Cho-TiO2 electrodes, respectively. The incident photon to current efficiency (IPCE) at 660 nm was 5.0, 4.1, and 4.1%, respectively [55]. These results highlight the relevance of the adequate chlorophyll anchorage to TiO2 for an improved performance of DSSCs. The adsorption of chlorophyll-a on TiO2 films is complex in many cases, impairing the performance of solar cells despite the relatively high short-circuit current density [39].
The proper contact of the dye with the TiO2 electrode is fundamental for the efficient transport of electrons between them, justifying the need to promote a good incorporation of chlorophyll in the sensitization step, as well as preventing the detachment of the dye from the TiO2 surface, which may lead to a DSSC performance decrease over time [49]. As for the light-harvesting device’s stability, Najihah et al. observed the decrease in the JSC and η at 96 h to 84% and 63% of the as-fabricated values, respectively, and attributed the reduction in the DSSC performance to the detachment of chlorophyll from the TiO2 surface, probably caused by the presence of impurities in the dye extract [49]. TiO2 nanotubes sensitized by Tsui et al. with bacteriochlorophyll-C showed photocurrent stability for 14 min under simulated sunlight [50].
In the work of Yang et al. [62], the measured short-circuit current density (JSC) was 1.386 µA cm−2 for the TiO2 electrode modified with large LHCII aggregates with 100% of adequate surface coverage (ESC), in comparison to 0.876 µA cm−2 for the bare APTES-TiO2 electrodes. This corresponded to a 58.2% increase in the JSC value after incorporating LHCII aggregates in the electrode [62]. Interestingly, the authors observed the excellent stability of the sensitized solar cell and the improvement over time of properties such as JSC, fill factor, and power conversion efficiency during the 30-day interval.

4. Conclusions

Using natural pigments has grown in TiO2 sensitization applications in photocatalytic processes and DSSCs. Chlorophyll is a natural pigment in photosynthetic organisms, such as plants and cyanobacteria, that plays a fundamental role in transforming solar energy into chemical energy. It can be extracted from different biomasses through relatively simple processes involving organic solvents. Once extracted, it can be applied to the sensitization of semiconductors, such as TiO2, bringing as one of the main benefits its increased response under the action of visible light. This response under visible light directly impacts its performance in photocatalytic processes, affecting its band gap and selectivity in some processes. Furthermore, it is an excellent option for DSSCs, and the main point to be considered is the proper anchoring of chlorophyll to the TiO2 surface, which is fundamental for the adequate performance of the solar cell. Promoting a good interaction between chlorophyll and TiO2 has been one of the main points affecting the process, which still requires improvements from the conversion efficiency and stability point of view.

Author Contributions

Conceptualization, M.E.K.F.; writing, review, and editing, M.E.K.F., A.M.T., O.A.A.d.S. and G.G.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data is contained within the article.

Acknowledgments

The authors thank the Capes, Fundação Araucaria, and CNPq agencies.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schneider, J.; Matsuoka, M.; Takeuchi, M.; Zhang, J.; Horiuchi, Y.; Anpo, M.; Bahnemann, D.W. Understanding TiO2 photocatalysis: Mechanisms and Materials. Chem. Rev. 2014, 114, 9919–9986. [Google Scholar] [CrossRef] [PubMed]
  2. Hashimoto, K.; Irie, H.; Fujishima, A. TiO2 Photocatalysis: A Historical Overview and Future Prospects. Jpn. J. Appl. Phys. 2005, 44, 8269. [Google Scholar] [CrossRef]
  3. Fujishima, A.; Honda, K. Electrochemical Evidence for the Mechanism of the Primary Stage of Photosynthesis. Bull. Chem. Soc. Jpn. 1971, 44, 1148–1150. [Google Scholar] [CrossRef]
  4. Augugliaro, V.; Bellardita, M.; Loddo, V.; Palmisano, G.; Palmisano, L.; Yurdakal, S. Overview on Oxidation Mechanisms of Organic Compounds by TiO2 in Heterogeneous Photocatalysis. J. Photochem. Photobiol. C Photochem. Rev. 2012, 13, 224–245. [Google Scholar] [CrossRef]
  5. Varma, K.S.; Tayade, R.J.; Shah, K.J.; Joshi, P.A.; Shukla, A.D.; Gandhi, V.G. Photocatalytic Degradation of Pharmaceutical and Pesticide Compounds (PPCs) Using Doped TiO2 Nanomaterials: A Review. Water-Energy Nexus 2020, 3, 46–61. [Google Scholar] [CrossRef]
  6. Murgolo, S.; De Ceglie, C.; Di Iaconi, C.; Mascolo, G. Novel TiO2-Based Catalysts Employed in Photocatalysis and Photoelectrocatalysis for Effective Degradation of Pharmaceuticals (PhACs) in Water: A Short Review. Curr. Opin. Green. Sustain. Chem. 2021, 30, 100473. [Google Scholar] [CrossRef]
  7. Shoneye, A.; Sen Chang, J.; Chong, M.N.; Tang, J. Recent Progress in Photocatalytic Degradation of Chlorinated Phenols and Reduction of Heavy Metal Ions in Water by TiO2-Based Catalysts. Int. Mater. Rev. 2022, 67, 47–64. [Google Scholar] [CrossRef]
  8. Bono, N.; Ponti, F.; Punta, C.; Candiani, G. Effect of UV Irradiation and TiO2-Photocatalysis on Airborne Bacteria and Viruses: An Overview. Materials 2021, 14, 1075. [Google Scholar] [CrossRef] [PubMed]
  9. Liao, C.H.; Huang, C.W.; Wu, J.C.S. Hydrogen Production from Semiconductor-Based Photocatalysis via Water Splitting. Catalysts 2012, 2, 490–516. [Google Scholar] [CrossRef]
  10. Chiarello, G.L.; Dozzi, M.V.; Selli, E. TiO2-Based Materials for Photocatalytic Hydrogen Production. J. Energy Chem. 2017, 26, 250–258. [Google Scholar] [CrossRef]
  11. Kumaravel, V.; Mathew, S.; Bartlett, J.; Pillai, S.C. Photocatalytic Hydrogen Production Using Metal Doped TiO2: A Review of Recent Advances. Appl. Catal. B 2019, 244, 1021–1064. [Google Scholar] [CrossRef]
  12. Wang, Z.; Zhou, W.; Wang, X.; Zhang, X.; Chen, H.; Hu, H.; Liu, L.; Ye, J.; Wang, D. Enhanced Photocatalytic CO2 Reduction over TiO2 Using Metalloporphyrin as the Cocatalyst. Catalysts 2020, 10, 654. [Google Scholar] [CrossRef]
  13. Roy, N.; Suzuki, N.; Terashima, C.; Fujishima, A. Recent Improvements in the Production of Solar Fuels: From CO2 Reduction to Water Splitting and Artificial Photosynthesis. Bull. Chem. Soc. Jpn. 2019, 92, 178–192. [Google Scholar] [CrossRef]
  14. Gust, D.; Moore, T.A.; Moore, A.L. Solar Fuels via Artificial Photosynthesis. Acc. Chem. Res. 2009, 42, 1890–1898. [Google Scholar] [CrossRef] [PubMed]
  15. Guo, Q.; Zhou, C.; Ma, Z.; Yang, X. Fundamentals of TiO2 Photocatalysis: Concepts, Mechanisms, and Challenges. Adv. Mater. 2019, 31, 1901997. [Google Scholar] [CrossRef] [PubMed]
  16. Bora, L.V.; Mewada, R.K. Visible/Solar Light Active Photocatalysts for Organic Effluent Treatment: Fundamentals, Mechanisms and Parametric Review. Renew. Sustain. Energy Rev. 2017, 76, 1393–1421. [Google Scholar] [CrossRef]
  17. Linsebigler, A.L.; Lu, G.; Yates, J.T. Photocatalysis on TiO2 Surfaces: Principles, Mechanisms, and Selected Results. Chem. Rev. 1995, 95, 735–758. [Google Scholar] [CrossRef]
  18. Chen, D.; Cheng, Y.; Zhou, N.; Chen, P.; Wang, Y.; Li, K.; Huo, S.; Cheng, P.; Peng, P.; Zhang, R.; et al. Photocatalytic Degradation of Organic Pollutants Using TiO2-Based Photocatalysts: A Review. J. Clean. Prod. 2020, 268, 121725. [Google Scholar] [CrossRef]
  19. Krishnan, S.; Shriwastav, A. Application of TiO2 Nanoparticles Sensitized with Natural Chlorophyll Pigments as Catalyst for Visible Light Photocatalytic Degradation of Methylene Blue. J. Environ. Chem. Eng. 2021, 9, 104699. [Google Scholar] [CrossRef]
  20. Finnie, K.S.; Bartlett, J.R.; Woolfrey, J.L. Vibrational Spectroscopic Study of the Coordination of (2,2‘-Bipyridyl-4,4‘-Dicarboxylic Acid)Ruthenium(II) Complexes to the Surface of Nanocrystalline Titania. Langmuir 1998, 14, 2744–2749. [Google Scholar] [CrossRef]
  21. Zhang, H.; Zhou, Y.; Zhang, M.; Shen, T.; Li, Y.; Zhu, D. Photoinduced Interaction between Fluorescein Ester Derivatives and CdS Colloid. J. Colloid. Interface Sci. 2003, 264, 290–295. [Google Scholar] [CrossRef] [PubMed]
  22. Kathiravan, A.; Chandramohan, M.; Renganathan, R.; Sekar, S. Cyanobacterial Chlorophyll as a Sensitizer for Colloidal TiO2. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2009, 71, 1783–1787. [Google Scholar] [CrossRef] [PubMed]
  23. Mujtahid, F.; Gareso, P.L.; Armynah, B.; Tahir, D. Review Effect of Various Types of Dyes and Structures in Supporting Performance of Dye-Sensitized Solar Cell TiO2-Based Nanocomposites. Int. J. Energy Res. 2022, 46, 726–742. [Google Scholar] [CrossRef]
  24. Karim, N.A.; Mehmood, U.; Zahid, H.F.; Asif, T. Nanostructured Photoanode and Counter Electrode Materials for Efficient Dye-Sensitized Solar Cells (DSSCs). Solar Energy 2019, 185, 165–188. [Google Scholar] [CrossRef]
  25. Hosseinnezhad, M.; Gharanjig, K.; Yazdi, M.K.; Zarrintaj, P.; Moradian, S.; Saeb, M.R.; Stadler, F.J. Dye-Sensitized Solar Cells Based on Natural Photosensitizers: A Green View from Iran. J. Alloys Compd. 2020, 828, 154329. [Google Scholar] [CrossRef]
  26. Richhariya, G.; Kumar, A.; Tekasakul, P.; Gupta, B. Natural Dyes for Dye Sensitized Solar Cell: A Review. Renew. Sustain. Energy Rev. 2017, 69, 705–718. [Google Scholar] [CrossRef]
  27. Omar, A.; Ali, M.S.; Abd Rahim, N. Electron Transport Properties Analysis of Titanium Dioxide Dye-Sensitized Solar Cells (TiO2-DSSCs) Based Natural Dyes Using Electrochemical Impedance Spectroscopy Concept: A Review. Sol. Energy 2020, 207, 1088–1121. [Google Scholar] [CrossRef]
  28. Mukhokosi, E.P.; Maaza, M.; Tibenkana, M.; Botha, N.L.; Namanya, L.; Madiba, I.G.; Okullo, M. Optical Absorption and Photoluminescence Properties of Cucurbita Maxima Dye Adsorption on TiO2nanoparticles. Mater. Res. Express 2023, 10, 046203. [Google Scholar] [CrossRef]
  29. Ahliha, A.H.; Nurosyid, F.; Supriyanto, A.; Kusumaningsih, T. The Chemical Bonds Effect of Anthocyanin and Chlorophyll Dyes on TiO2 for Dye-Sensitized Solar Cell (DSSC). In Journal of Physics: Conference Series, Proceedings of the International Conference on Science and Applied Science 2017, Solo, Indonesia, 29 July 2017; Institute of Physics Publishing: Bristol, UK, 2017; Volume 909. [Google Scholar]
  30. Haghighatzadeh, A. Comparative Analysis on Optical and Photocatalytic Properties of Chlorophyll/Curcumin-Sensitized TiO2 Nanoparticles for Phenol Degradation. Bull. Mater. Sci. 2020, 43, 52. [Google Scholar] [CrossRef]
  31. Arof, A.K.; Ping, T.L. Chlorophyll as Photosensitizer in Dye-Sensitized Solar Cells. In Chlorophyll; InTech: London, UK, 2017. [Google Scholar]
  32. Krishnan, S.; Shriwastav, A. Chlorophyll Sensitized and Salicylic Acid Functionalized TiO2 Nanoparticles as a Stable and Efficient Catalyst for the Photocatalytic Degradation of Ciprofloxacin with Visible Light. Environ. Res. 2023, 216, 114568. [Google Scholar] [CrossRef]
  33. Phongamwong, T.; Chareonpanich, M.; Limtrakul, J. Role of Chlorophyll in Spirulina on Photocatalytic Activity of CO2 Reduction under Visible Light over Modified N-Doped TiO2 Photocatalysts. Appl. Catal. B 2015, 168–169, 114–124. [Google Scholar] [CrossRef]
  34. Shen, S.; Wang, Y.; Dong, J.; Zhang, R.; Parikh, A.; Chen, J.G.; Hu, D. Mimicking Thylakoid Membrane with Chlorophyll/TiO2/Lipid Co-Assembly for Light-Harvesting and Oxygen Releasing. ACS Appl. Mater. Interfaces 2021, 13, 11461–11469. [Google Scholar] [CrossRef] [PubMed]
  35. Gross, J. Pigments in Vegetables; Springer: Boston, MA, USA, 1991; ISBN 978-1-4613-5842-8. [Google Scholar]
  36. Murata, N. Control of Excitation Transfer in Photosynthesis II. Magnesium Ion-Dependent Distribution of Excitation Energy between Two Pigment Systems in Spinach Chloroplasts. Biochim. Biophys. Acta BBA 1969, 189, 171–181. [Google Scholar] [CrossRef]
  37. Jamali Jaghdani, S.; Jahns, P.; Tränkner, M. The Impact of Magnesium Deficiency on Photosynthesis and Photoprotection in Spinacia Oleracea. Plant Stress 2021, 2, 100040. [Google Scholar] [CrossRef]
  38. Mary Rosana, N.T.; Joshua Amarnath, D.; Senthil Kumar, P.; Vincent Joseph, K.L. Potential of Plant-Based Photosensitizers in Dye-Sensitized Solar Cell Applications. Environ. Prog. Sustain. Energy 2020, 39, e13351. [Google Scholar] [CrossRef]
  39. Pai, A.R.; Nair, B. Synthesis and Characterization of a Binary Oxide ZrO2-TiO2 and Its Application in Chlorophyll Dye-Sensitized Solar Cell with Reduced Graphene Oxide as Counter Electrodes. Bull. Mater. Sci. 2015, 38, 1129–1133. [Google Scholar] [CrossRef]
  40. Christwardana, M.; Septevani, A.A.; Yoshi, L.A. Outstanding Photo-Bioelectrochemical Cell by Integrating TiO2 and Chlorophyll as Photo-Bioanode for Sustainable Energy Generation. Int. J. Renew. Energy Dev. 2022, 11, 385–391. [Google Scholar] [CrossRef]
  41. Mahadik, S.A.; Yadav, H.M.; Mahadik, S.S. Surface Properties of Chlorophyll-a Sensitized TiO2 Nanorods for Dye-Sensitized Solar Cells Applications. Colloids Interface Sci. Commun. 2022, 46, 100558. [Google Scholar] [CrossRef]
  42. Heidari, A.; Safa, K.D.; Teimuri-Mofrad, R. Chlorophyll B-Modified TiO2 Nanoparticles for Visible-Light-Induced Photocatalytic Synthesis of New Tetrahydroquinoline Derivatives. Mol. Catal. 2023, 547, 113338. [Google Scholar] [CrossRef]
  43. Chen, C.; Gong, N.; Qu, F.; Gao, Y.; Fang, W.; Sun, C.; Men, Z. Effects of Carotenoids on the Absorption and Fluorescence Spectral Properties and Fluorescence Quenching of Chlorophyll a. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2018, 204, 440–445. [Google Scholar] [CrossRef]
  44. Ungureanu, E.M.; Tatu, M.L.; Georgescu, E.; Boscornea, C.; Popa, M.M.; Stanciu, G. Influence of the Chemical Structure and Solvent Polarity on the Fluorescence of 3-Aryl-7-Benzoyl-Pyrrolo [1,2-c]Pyrimidines. Dyes Pigments 2020, 174, 108023. [Google Scholar] [CrossRef]
  45. Faiz, M.R.; Widhiyanuriyawan, D.; Siswanto, E.; Wardana, I.N.G. Theoretical Study on the Effect of Solvents in Chlorophyll Solution for Dye-Sensitized Solar Cell. In IOP Conference Series: Materials Science and Engineering, Proceedings of the International Conference on Chemistry and Material Science (IC2MS) 2017, Malang, Indonesia, 4–5 November 2017; Institute of Physics Publishing: Bristol, UK, 2018; Volume 299, p. 299. [Google Scholar]
  46. Sabagh, S.; Izadyar, M.; Arkan, F. Insight into Incident Photon to Current Conversion Efficiency in Chlorophylls. Int. J. Quantum Chem. 2021, 121, e26483. [Google Scholar] [CrossRef]
  47. Qiao, L.; Tang, W.; Gao, D.; Zhao, R.; An, L.; Li, M.; Sun, H.; Song, D. UAV-Based Chlorophyll Content Estimation by Evaluating Vegetation Index Responses under Different Crop Coverages. Comput. Electron. Agric. 2022, 196, 106775. [Google Scholar] [CrossRef]
  48. Al-Alwani, M.A.M.; Mohamad, A.B.; Kadhum, A.A.H.; Ludin, N.A.; Safie, N.E.; Razali, M.Z.; Ismail, M.; Sopian, K. Natural Dye Extracted from Pandannus Amaryllifolius Leaves as Sensitizer in Fabrication of Dye-Sensitized Solar Cells. Int. J. Electrochem. Sci. 2017, 12, 747–761. [Google Scholar] [CrossRef]
  49. Najihah, M.Z.; Noor, I.M.; Winie, T. Long-Run Performance of Dye-Sensitized Solar Cell Using Natural Dye Extracted from Costus Woodsonii Leaves. Opt. Mater. 2022, 123, 111915. [Google Scholar] [CrossRef]
  50. Tsui, L.K.; Huang, J.; Sabat, M.; Zangari, G. Visible Light Sensitization of TiO2 Nanotubes by Bacteriochlorophyll-C Dyes for Photoelectrochemical Solar Cells. ACS Sustain. Chem. Eng. 2014, 2, 2097–2101. [Google Scholar] [CrossRef]
  51. Phongamwong, T.; Donphai, W.; Prasitchoke, P.; Rameshan, C.; Barrabés, N.; Klysubun, W.; Rupprechter, G.; Chareonpanich, M. Novel Visible-Light-Sensitized Chl-Mg/P25 Catalysts for Photocatalytic Degradation of Rhodamine B. Appl. Catal. B 2017, 207, 326–334. [Google Scholar] [CrossRef]
  52. Zakaria, P.N.M.; Noor, I.S.M.; Winie, T. Caulerpa lentillifera Seaweed Extract as Natural Sensitizer for Dye-Sensitized Solar Cell. Macromol. Symp. 2023, 407, 2100359. [Google Scholar] [CrossRef]
  53. Giang, N.T.H.; Thinh, N.T.; Hai, N.D.; Loc, P.T.; Thu, T.N.A.; Loan, N.H.P.; Quang, D.M.; Anh, L.D.; Truong An, V.N.T.; Phong, M.T.; et al. Application of TiO2 Nanoparticles with Natural Chlorophyll as the Catalyst for Visible Light Photocatalytic Degradation of Methyl Orange and Antibacterial. Inorg. Chem. Commun. 2023, 150, 110513. [Google Scholar] [CrossRef]
  54. Ma, T.; Inoue, K.; Yao, K.; Noma, H.; Shuji, T.; Abe, E.; Yu, J.; Wang, X.; Zhang, B. Photoelectrochemical Properties of TiO2 Electrodes Sensitized by Porphyrin Derivatives with Different Numbers of Carboxyl Groups. J. Electroanal. Chem. 2002, 537, 31–38. [Google Scholar] [CrossRef]
  55. Amao, Y.; Kato, K. Chlorophyll Assembled Electrode for Photovoltaic Conversion Device. Electrochim. Acta 2007, 53, 42–45. [Google Scholar] [CrossRef]
  56. Hashimoto, Y.; Suzuki, H.; Kondo, T.; Abe, R.; Tamiaki, H. Visible-Light-Induced Hydrogen Evolution from Water on Hybrid Photocatalysts Consisting of Synthetic Chlorophyll-a Derivatives with a Carboxy Group in the 20-Substituent Adsorbed on Semiconductors. J. Photochem. Photobiol. A Chem. 2022, 426, 113750. [Google Scholar] [CrossRef]
  57. Zhao, W.; Wang, L.; Pan, L.; Duan, S.; Tamai, N.; Sasaki, S.I.; Tamiaki, H.; Sanehira, Y.; Wei, Y.; Chen, G.; et al. Charge Transfer Dynamics in Chlorophyll-Based Biosolar Cells. Phys. Chem. Chem. Phys. 2019, 21, 22563–22568. [Google Scholar] [CrossRef] [PubMed]
  58. Pan, X.; Dong, W.; Zhang, J.; Xie, Z.; Li, W.; Zhang, H.; Zhang, X.; Chen, P.; Zhou, W.; Lei, B. TiO2/Chlorophyll S-Scheme Composite Photocatalyst with Improved Photocatalytic Bactericidal Performance. ACS Appl. Mater. Interfaces 2021, 13, 39446–39457. [Google Scholar] [CrossRef] [PubMed]
  59. Valadez-Renteria, E.; Oliva, J.; Rodriguez-Gonzalez, V. A Sustainable and Green Chlorophyll/TiO2:W Composite Supported on Recycled Plastic Bottle Caps for the Complete Removal of Rhodamine B Contaminant from Drinking Water. J. Environ. Manage 2022, 315, 115204. [Google Scholar] [CrossRef] [PubMed]
  60. Patiño-Camelo, K.; Diaz-Uribe, C.; Gallego-Cartagena, E.; Vallejo, W.; Martinez, V.; Quiñones, C.; Hurtado, M.; Schott, E. Cyanobacterial Biomass Pigments as Natural Sensitizer for TiO2 Thin Films. Int. J. Photoenergy 2019, 2019, 7184327. [Google Scholar] [CrossRef]
  61. Cabir, B.; Yurderi, M.; Caner, N.; Agirtas, M.S.; Zahmakiran, M.; Kaya, M. Methylene Blue Photocatalytic Degradation under Visible Light Irradiation on Copper Phthalocyanine-Sensitized TiO2 Nanopowders. Mater. Sci. Eng. B Solid. State Mater. Adv. Technol. 2017, 224, 9–17. [Google Scholar] [CrossRef]
  62. Yang, Y.; Jankowiak, R.; Lin, C.; Pawlak, K.; Reus, M.; Holzwarth, A.R.; Li, J. Effect of the LHCII Pigment-Protein Complex Aggregation on Photovoltaic Properties of Sensitized TiO2 Solar Cells. Phys. Chem. Chem. Phys. 2014, 16, 20856–20865. [Google Scholar] [CrossRef]
  63. Li, Y.; Yoo, K.; Lee, D.K.; Kim, J.Y.; Son, H.J.; Kim, J.H.; Lee, C.H.; Míguez, H.; Ko, M.J. Synergistic Strategies for the Preparation of Highly Efficient Dye-Sensitized Solar Cells on Plastic Substrates: Combination of Chemical and Physical Sintering. RSC Adv. 2015, 5, 76795–76803. [Google Scholar] [CrossRef]
  64. Amao, Y.; Yamada, Y.; Aoki, K. Preparation and Properties of Dye-Sensitized Solar Cell Using Chlorophyll Derivative Immobilized TiO2 Film Electrode. J. Photochem. Photobiol. A Chem. 2004, 164, 47–51. [Google Scholar] [CrossRef]
Figure 1. Chlorophyll-a (a) and Chlorophyll-b (b).
Figure 1. Chlorophyll-a (a) and Chlorophyll-b (b).
Reactions 04 00044 g001
Figure 2. Representation of (a) unidentate, (b) bidentate chelating, and (c) bridging coordination modes.
Figure 2. Representation of (a) unidentate, (b) bidentate chelating, and (c) bridging coordination modes.
Reactions 04 00044 g002
Figure 3. Schematic representation of a typical photocatalytic process based on chlorophyll-sensitized TiO2.
Figure 3. Schematic representation of a typical photocatalytic process based on chlorophyll-sensitized TiO2.
Reactions 04 00044 g003
Figure 4. Schematic representation of chlorophyll pigment excitation (left) and excited electrons transfer from chlorophyl to TiO2 conduction band (right).
Figure 4. Schematic representation of chlorophyll pigment excitation (left) and excited electrons transfer from chlorophyl to TiO2 conduction band (right).
Reactions 04 00044 g004
Figure 5. Schematic representation of a typical DDSC based on chlorophyll-sensitized TiO2.
Figure 5. Schematic representation of a typical DDSC based on chlorophyll-sensitized TiO2.
Reactions 04 00044 g005
Table 1. Different equations described in the literature used to determine the concentration of chlorophyll-a and chlorophyll-b for different solvents.
Table 1. Different equations described in the literature used to determine the concentration of chlorophyll-a and chlorophyll-b for different solvents.
EquationsSolventRef.
C h l a m g   L 1 = 11.93   · ( O D 664 ) 1.93   · O D 647 90% Aqueous acetone[19]
C h l b m g   L 1 = 20.36   · O D 647 5.50   · ( O D 664 )
C h l a μ g   m L 1 = 12.21   · A 663 2.81   · ( A 646 ) Acetone/
petroleum ether
[39]
C h l b μ g   m L 1 = 20.13   · A 646 5.03   · ( A 663 )
C h l a μ g   m L 1 = 13.95   · A 665 6.88   · ( A 649 ) 95% Ethanol[47]
C h l b μ g   m L 1 = 24.96   · A 649 7.32   · ( A 665 )
Table 2. Some extraction conditions described in the literature.
Table 2. Some extraction conditions described in the literature.
SourceSolventContact TimeSeparationRef.
Spinach leaves90% aqueous acetone2 hCentrifugation[19]
Spinach leavesEthanol20 minFiltration[41]
Spirulina sp.
(cyanobacteria)
90% methanol1 hCentrifugation[22]
SpirulinaMethanol and water
(incubated for 2 min in a water bath at 70 °C)
2 min (70 °C, incubated)Centrifugation[51]
SpirulinaEthanol1 h (sonicated)Filtration[40]
Pandan leaves
(P. amaryllifolius)
96% Ethanol1 weekFiltration[48]
Parsley leaves96 Ethanol7 daysFiltration[30]
Pumpkin (Cucurbita maxima) leavesEthanol36 hFiltration[28]
Table 3. Types of Extracts for TiO2 sensitizing.
Table 3. Types of Extracts for TiO2 sensitizing.
Dye/PigmentDSSC ConfigurationJSC
(mA cm−2)
VOC
(V)
FFη
(%)
Ref.
Spinach extractSpinach dye-sensitized TiO2/FTO electrode, Pt/FTO counter electrode0.330.6390.3370.0713[29]
Spinach extractChlorophyll-a-sensitized TiO2 nanorods and DSSC devices as described in [63]0.25660.53270.5621 [41]
C. woodsonii extractC. woodsonii-sensitized TiO2/FTO electrode and Pt/FTO counter electrode2.250.570.510.65[49]
P. amaryllifolius extractChl-TiO2/conductive glass electrode, I/I3− electrolyte solution in acetonitrile.0.40.550.60510.1[48]
Chl-e6 in ethanolChl-e6–TiO2 electrode, Pt-coated OTE electrode1.470.4250.570.4[64]
Plectranthus amboinicus extractDye-sensitized TiO2/FTO electrode, Pt counter electrode, I3/I redox couple.0.990.6370.630.46[38]
Coriandrum sativum extract0.620.6260.6150.28
Murraya koenigii extract0.630.6210.5950.27
H2Chl and ZnChlChl-based biosolar cell
(ZnChl)n/TiO2–H2Chl/TiO2/FTO electrode
4.490.670.441.33[57]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Fuziki, M.E.K.; Tusset, A.M.; dos Santos, O.A.A.; Lenzi, G.G. Chlorophyll Sensitization of TiO2: A Mini-Review. Reactions 2023, 4, 766-778. https://doi.org/10.3390/reactions4040044

AMA Style

Fuziki MEK, Tusset AM, dos Santos OAA, Lenzi GG. Chlorophyll Sensitization of TiO2: A Mini-Review. Reactions. 2023; 4(4):766-778. https://doi.org/10.3390/reactions4040044

Chicago/Turabian Style

Fuziki, Maria E. K., Angelo M. Tusset, Onélia A. A. dos Santos, and Giane G. Lenzi. 2023. "Chlorophyll Sensitization of TiO2: A Mini-Review" Reactions 4, no. 4: 766-778. https://doi.org/10.3390/reactions4040044

Article Metrics

Back to TopTop