Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 1931 2023-12-10 19:50:32 |
2 layout + 8 word(s) 1939 2023-12-11 04:25:53 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Fuziki, M.E.K.; Tusset, A.M.; Dos Santos, O.A.A.; Lenzi, G.G. Chlorophyll and TiO2. Encyclopedia. Available online: https://encyclopedia.pub/entry/52553 (accessed on 05 May 2024).
Fuziki MEK, Tusset AM, Dos Santos OAA, Lenzi GG. Chlorophyll and TiO2. Encyclopedia. Available at: https://encyclopedia.pub/entry/52553. Accessed May 05, 2024.
Fuziki, Maria E. K., Angelo M. Tusset, Onélia A. A. Dos Santos, Giane G. Lenzi. "Chlorophyll and TiO2" Encyclopedia, https://encyclopedia.pub/entry/52553 (accessed May 05, 2024).
Fuziki, M.E.K., Tusset, A.M., Dos Santos, O.A.A., & Lenzi, G.G. (2023, December 10). Chlorophyll and TiO2. In Encyclopedia. https://encyclopedia.pub/entry/52553
Fuziki, Maria E. K., et al. "Chlorophyll and TiO2." Encyclopedia. Web. 10 December, 2023.
Chlorophyll and TiO2
Edit

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.

Chlorophyll TiO2 dye sensitization photocatalysis

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].

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].
Figure 1. Chlorophyll-a (a) and Chlorophyll-b (b).
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 (𝑂𝐷) in the extract—which were calculated by subtracting the absorbance at 750 nm from absorbances at 647 nm (𝑂𝐷647) and 664 nm (𝑂𝐷664)—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]

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, researchers can mention leaves of fresh spinach [19], pandan [47], weeds such as Chromolaena odorata [39], and parsley [30]. Among microorganisms, the cyanobacteria stand out, among which researchers 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) [48]. 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 [47]. 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][48]. In some cases, it is also necessary to preliminarily remove the midrib of plant leaves [19] or carry out a drying step [47]. 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][47][49]). 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][47][48]. In some cases, it is also possible to increase the extract concentration in a rotary evaporator [29][47].
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 [50].
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 [47] or light [22][29][39][40][47] 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 [48].
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 [50].

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.
  2. Hashimoto, K.; Irie, H.; Fujishima, A. TiO2 Photocatalysis: A Historical Overview and Future Prospects. Jpn. J. Appl. Phys. 2005, 44, 8269.
  3. Fujishima, A.; Honda, K. Electrochemical Evidence for the Mechanism of the Primary Stage of Photosynthesis. Bull. Chem. Soc. Jpn. 1971, 44, 1148–1150.
  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.
  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.
  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.
  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.
  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.
  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.
  10. Chiarello, G.L.; Dozzi, M.V.; Selli, E. TiO2-Based Materials for Photocatalytic Hydrogen Production. J. Energy Chem. 2017, 26, 250–258.
  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.
  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.
  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.
  14. Gust, D.; Moore, T.A.; Moore, A.L. Solar Fuels via Artificial Photosynthesis. Acc. Chem. Res. 2009, 42, 1890–1898.
  15. Guo, Q.; Zhou, C.; Ma, Z.; Yang, X. Fundamentals of TiO2 Photocatalysis: Concepts, Mechanisms, and Challenges. Adv. Mater. 2019, 31, 1901997.
  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.
  17. Linsebigler, A.L.; Lu, G.; Yates, J.T. Photocatalysis on TiO2 Surfaces: Principles, Mechanisms, and Selected Results. Chem. Rev. 1995, 95, 735–758.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  31. Arof, A.K.; Ping, T.L. Chlorophyll as Photosensitizer in Dye-Sensitized Solar Cells. In Chlorophyll; InTech: London, UK, 2017.
  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.
  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.
  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.
  35. Gross, J. Pigments in Vegetables; Springer: Boston, MA, USA, 1991; ISBN 978-1-4613-5842-8.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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 Pyrimidines. Dyes Pigments 2020, 174, 108023.
  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.
  46. Sabagh, S.; Izadyar, M.; Arkan, F. Insight into Incident Photon to Current Conversion Efficiency in Chlorophylls. Int. J. Quantum Chem. 2021, 121, e26483.
  47. 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.
  48. 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.
  49. 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.
  50. 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.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , ,
View Times: 103
Revisions: 2 times (View History)
Update Date: 11 Dec 2023
1000/1000