Next Article in Journal
Hollow Nanospheres Organized by Ultra-Small CuFe2O4/C Subunits with Efficient Photo-Fenton-like Performance for Antibiotic Degradation and Cr(VI) Reduction
Previous Article in Journal
1,4-Butanediol Selective Dehydration to 3-Butene-1-ol over Ca–Zr–Sn Composite Oxide Catalysts
Previous Article in Special Issue
Visible Light Induced Nano-Photocatalysis Trimetallic Cu0.5Zn0.5-Fe: Synthesis, Characterization and Application as Alcohols Oxidation Catalyst
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Editorial

Heterogeneous Photocatalysis: A Solution for a Greener Earth

by
Julien G. Mahy
1,2,* and
Stéphanie D. Lambert
1
1
Department of Chemical Engineering—Nanomaterials, Catalysis, Electrochemistry, B6a, University of Liège, B-4000 Liège, Belgium
2
Institute of Condensed Matter and Nanosciences (IMCN), Université Catholique de Louvain, Place Louis Pasteur 1, B-1348 Louvain-la-Neuve, Belgium
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(7), 686; https://doi.org/10.3390/catal12070686
Submission received: 16 June 2022 / Accepted: 21 June 2022 / Published: 23 June 2022
(This article belongs to the Special Issue Heterogeneous Photocatalysis: A Solution for a Greener Earth)
Since the beginning of the industrial era, various human activities have steadily increased, leading to rapid technological developments and high population growth. In consequence, the expanding industry has heavily polluted the atmosphere, soil, and water, with negative consequences for humans and the environment.
To decrease pollution emissions, various chemical, physical, and biological treatment methods have been developed. The major technics for treating wastewater are based on wastewater treatment plants using dry cleaning, decantation, and biological treatments. Occasionally, pollutant molecules are not eliminated by these processes; therefore, other technics can be used as secondary treatments to remove these small residual fractions of pollution. Among these methods, photocatalysis is a well-developed technic in the past few years. Through a photocatalyst and using light’s energy, photocatalysis allows the production of highly reactive species that can react and decompose organic molecules, yielding, in the best case, the final decomposition products CO2 and H2O. The most commonly used photocatalysts are titanium dioxide (TiO2), zinc oxide (ZnO), and tin oxide (SnO2).
The papers contributing to this Special Issue address innovative photocatalytic processes for environmental applications. In what follows, we provide a synopsis of the obtained results in the 17 papers published in this Special Issue.
1.
TiO2-based photocatalysts
In this collection of articles, TiO2 was used as the photocatalyst for depollution applications, and different doping and shaping were used to produce photoefficient materials.
In a contribution by Almeida et al. [1], the efficiency of photodegradation, the volatilization profile of bidentates, and the role of reactive oxidizing species (ROS) were explored for nanocrystalline TiO2 modified with bidentate ligands (acetylacetone). In this study, TiO2-ACAC CTC calcined at 300 °C (TiO2-A300) was applied for the photocatalytic degradation of chlorophenol (4-CP) and tetracycline (TC) under low power visible light (26 W). Furthermore, the ROS scavengers isopropanol and benzoquinone were added for studying the photocatalytic role of OH and O2 radicals. The photocatalytic abatement of tetracycline (68.6%), performed via TiO2-A300, was two times higher than that observed for chlorophenol (31.3%) after 6 h, indicating the distinct participation of ROS in the degradation of these pollutants. The addition of the ROS scavenger revealed O2 radicals as primarily responsible for the high efficiency of TiO2-ACAC CTC under reduced visible light. On the other hand, the OH radicals were not efficiently generated in the CTC. Therefore, the development of heterostructures based on TiO2-ACAC CTC can increase the generation of ROS through coupling with semiconductors capable of generating OH under visible light.
In the study of Pascariu et al. [2], Ag–TiO2 nanostructures were prepared via electrospinning, followed by calcination at 400 °C. Morphological characterization revealed the presence of one-dimensional uniform Ag–TiO2 nanostructured nanofibers, with a diameter from 65 to 100 nm, depending on the Ag loading, composed of small crystals interconnected with each other. Structural characterization indicated that Ag was successfully integrated as small nanocrystals without affecting much of the TiO2 crystal lattice. Moreover, the presence of nano-Ag was found to contribute to reducing the band gap energy, which enables the activation by the absorption of visible light while, at the same time, delaying the electron–hole recombination. Tests of their photocatalytic activity in methylene blue, amaranth, Congo red, and orange II degradation revealed an increase by more than 20% in color removal efficiency at an almost double rate for the case of 0.1% Ag–TiO2 nanofibers compared with pure TiO2.
In another study [3], Mancuso et al. successfully prepared different tri-doped TiO2 photocatalysts (Fe-N-P/TiO2, Fe-N-S/TiO2, Fe-Pr-N/TiO2, Pr-N-S/TiO2, and P-N-S/TiO2) and tested them in the photocatalytic removal of thiacloprid (THI) under UV-A, visible, and direct solar light irradiation. The physical–chemical properties of the prepared catalysts revealed that dopants were effectively incorporated into the anatase TiO2 lattice, resulting in a decrease in the energy band gap. The reduction in photoluminescence intensity indicated a lower combination rate and longer lifespan of photogenerated carriers of all doped samples in comparison with the undoped TiO2. The doped photocatalysts promoted photodegradation under UV-A light irradiation and also extended the optical response of TiO2 to the visible light region. Fe–N–P tri-doped TiO2 sample exhibited the highest THI photodegradation degree (64% under UV-A light, 29% under visible light, and 73% under solar light).
In the contribution of Benkhennouche-Bouchene et al. [4], TiO2 that was prepared using a green aqueous sol–gel peptization process was co-doped with nitrogen and zirconium to improve and extend its photoactivity to the visible region. For all doped and co-doped samples, TiO2 nanoparticles with sizes ranging from 4 to 8 nm were formed of anatase–brookite phases. X-ray photoelectron (XPS) measurements showed that nitrogen was incorporated into the TiO2 materials through Ti–O–N bonds, allowing light absorption in the visible region. The XPS spectra of the Zr-(co)doped powders showed the presence of TiO2–ZrO2 mixed oxide materials. Under visible light, the best co-doped sample yielded a degradation of p-nitrophenol (PNP) equal to 70%, instead of 25% with pure TiO2 and 10% with P25 under the same conditions. Similarly, the photocatalytic activity improved under UV–Vis, reaching 95% with the best sample, compared with 50% with pure TiO2.
In the study of Eun et al. [5], sol–gel-synthesized N-doped and carbon–nitrogen–sulfur (CNS)-doped TiO2 solutions were deposited on upconversion phosphor using a dip-coating method. Scanning electron microscopy (SEM) imaging showed that there was a change in the morphology of TiO2 coated on NaYF4:Yb,Er from spherical to nanorods caused by additional urea and thiourea doping reagents. Fourier transform infrared (FTIR) spectroscopy further verified the existence of nitrate–hyponitrite, carboxylate, and SO42− because of the doping effect. NaYF4:Yb,Er composites coated with N- and CNS-doped TiO2 exhibited a slight shift in UV–Vis spectra toward the visible light region. The photocatalytic reactivity with CNS-doped TiO2/NaYF4:Yb,Er surpassed that of the undoped TiO2/NaYF4:Yb,Er for the MB solution and toluene. The photocatalytic activity was increased by CNS doping of TiO2, which improved light sensitization as a result of band gap narrowing due to impurity sites.
In the presented study by Mahy et al. [6], the objective was to improve the efficiency of TiO2 photocatalysts via activation treatments and through modification with palladium nanoparticles and doping with SiO2. X-ray diffraction provided evidence that the crystallographic structure of TiO2 was anatase and that Pd was present, either in its oxidized form after calcination or in its reduced form after reduction. The results on methylene blue degradation showed that the photocatalytic activity of the catalysts was inversely proportional to the content of silica present in the matrix. A small amount of silica improved the photocatalytic activity, compared with the pure TiO2 sample. By contrast, a high amount of silica delayed the crystallization of TiO2 in its anatase form. The introduction of Pd species increased the photocatalytic activity of the samples because it allowed for a decrease in the rate of electron–hole recombination in TiO2. The reduction treatment improved the activity of photocatalysts, regardless of the palladium content, owing to the reduction of Ti4+ to Ti3+, and the formation of defects in the crystallographic structure of anatase.
A review of the green synthesis of TiO2 is also presented in this issue [7]. Indeed, in this study, the authors reported an eco-friendly process for producing TiO2 via colloidal aqueous sol–gel synthesis, resulting in crystalline materials without a calcination step. Three types of colloidal aqueous TiO2 were reviewed: the as-synthesized type obtained directly after synthesis, without any specific treatment; the calcined, obtained after a subsequent calcination step; and the hydrothermal, obtained after a specific autoclave treatment. This eco-friendly process is based on the hydrolysis of a Ti precursor in excess of water, followed by the peptization of the precipitated TiO2. Depending on the synthesis parameters, the three crystalline phases of TiO2 (anatase, brookite, and rutile) can be obtained. The morphology of the nanoparticles can also be tailored by the synthesis parameters. The most important parameter is the peptizing agent. Indeed, depending on its acidic or basic character and also on its amount, it can modulate the crystallinity and morphology of TiO2. Colloidal aqueous TiO2 photocatalysts are mainly being used in various photocatalytic reactions for organic pollutant degradation. The as-synthesized materials seem to have equivalent photocatalytic efficiency to the photocatalysts post-treated with thermal treatments and the commercial Evonik Aeroxide P25, which is produced via a high-temperature process.
TiO2 in form of film is also presented in the contribution of De Ceglie et al. [8]. In this study, the photocatalytic efficiency of an innovative UV-light catalyst consisting of a mesoporous TiO2 coating on glass fibers was investigated. Photocatalytic activity of the synthesized material was tested, for the first time, on a secondary wastewater effluent spiked with nine pharmaceuticals (PhACs), and the results were compared with the photolysis used as a benchmark treatment. Interestingly, the novel photocatalyst led to an increase in the degradation of carbamazepine and trimethoprim (about 2.2 times faster than the photolysis). Several transformation products (TPs) resulting from both the spiked PhACs and the compounds naturally occurring in the secondary wastewater effluent were identified through UPLC–QTOF/MS/MS. Some of them, produced mainly from carbamazepine and trimethoprim, were still present at the end of the photolytic treatment, while they were completely or partially removed by the photocatalytic treatment.
2.
Composite materials as photocatalysts
In the papers under this theme, different composite materials were produced to be used as photocatalysts.
In one study [9], Ghazzy et al. reported a visible-light-induced, trimetallic catalyst (Cu0.5Zn0.5Fe2O4) prepared through green synthesis using Tilia plant extract. The spinel crystalline material was ~34 nm. In benign reaction conditions, the prepared photocatalyst oxidized various benzylic alcohols with excellent yield and selectivity toward aldehyde with 99% and 98%, respectively. Aromatic and aliphatic alcohols (such as furfuryl alcohol and 1-octanol) were photocatalytically oxidized using Cu0.5Zn0.5Fe2O4, LED light, and H2O2 as oxidant, 2 h reaction time, and ambient temperature. The advantages of the catalyst were found in terms of reduced catalyst loading, activating catalyst using visible light in mild conditions, high conversion of the starting material, and recyclability of up to 5 times without loss of selectivity.
In the contribution of Mahy et al. [10], raw clays, from Cameroon, were modified with semiconductors (TiO2 and ZnO) to improve their depollution properties with the addition of photocatalytic properties. Cu2+ ions were also added to the clay via an ionic exchange to increase the specific surface area. The presence of TiO2 and ZnO was confirmed by the detection of anatase and wurtzite, respectively. The composite clays showed increased specific surface areas. The adsorption property of the raw clays was evaluated on two pollutants—namely, fluorescein (FL) and p-nitrophenol (PNP). The experiments showed that the raw clays can adsorb FL but were not efficient for PNP. To demonstrate the photocatalytic property resulting from the added semiconductors, photocatalytic experiments were performed under UV-A light on PNP. These experiments showed degradation up to 90% after 8 h of exposure with the best ZnO-modified clay. The proposed treatment of raw clays seems promising to treat pollutants, especially in developing countries.
In another contribution [11], Dao et al. prepared co-doped NiTiO3/g-C3N4 composite photocatalysts using a modified Pechini method to improve their photocatalytic activity toward methylene blue photodegradation under visible light irradiation. The combination of Co-doped NiTiO3 and g-C3N4 and Co-doping into the NiTiO3 lattice synergistically enhanced the photocatalytic performance of the composite photocatalysts. X-ray photoelectron spectroscopy results for the Co-doped NiTiO3/g-C3N4 composite confirmed Ti–N linkages between the Co-doped NiTiO3 and g-C3N4. In addition, characteristic X-ray diffraction peaks for the NiTiO3 lattice structure clearly indicated the substitution of Co into the N–TiO3 lattice structure. The composite structure and Co-doping of the C–x composite photocatalysts (x wt % Co-doped NiTiO3/g-C3N4) decreased the emission intensity of the photoluminescence spectra but also the semicircle radius of the Nyquist plot in electrochemical impedance spectroscopy, giving the highest Kapp value (7.15 × 10−3 min−1) for the Co-1 composite photocatalyst.
In the study of Bazta et al. [12], pure and Ce-modified ZnO nanosheet-like polycrystalline samples were successfully synthesized via a microwave-based process. The XRD results showed that the obtained photocatalysts were composed of hexagonal, wurtzite-type crystallites in the 34–44 nm size range. The microscopy showed nanosheet-shaped crystallites, with a composition close to Ce0.68Zn0.32Ox. Importantly, the STEM–XEDS characterization of the photocatalyst samples revealed that Ce did not incorporate into the ZnO crystal lattice as a dopant but that a heterojunction formed between the ZnO nanosheets and the Ce–Zn mixed oxide phase nanoparticles. The optical analysis revealed that in the ZnO:Ce samples optical band gap was found to decrease to 3.17 eV in the samples with the highest Ce content. It was also found that the ZnO:Ce (2 at.%) sample exhibited the highest photocatalytic activity for the degradation of methylene blue (MB) when compared with both the pure ZnO and commercial TiO2–P25 under simulated sunlight irradiation.
3.
Pilot and comparative process studies, and other photocatalytic processes
In this last collection of papers, pilot reactors were tested for larger-scale photocatalytic processes, and other advanced oxidative processes were explored.
In one study [13], Faccani et al. offered easily scalable solutions for adapting TiO2-based photocatalysts, which were deposited on different kinds of fabrics and implemented in a 6 L semipilot plant, using the photodegradation of rhodamine B (RhB) as a model of water pollution. They took advantage of a multivariable optimization approach to identify the best design options in terms of photodegradation efficiency and turnover frequency (TOF). Surprisingly, in the condition of use, the irradiation with a light-emitting diode (LED) visible lamp appeared as a valid alternative to the use of UV LED. The identification of the best design options in the semipilot plant allowed scaling up the technology in a 100 L pilot plant suitable for the treatment of industrial wastewater.
In the presented paper by Gonzalez-Burciaga et al. [14], several advanced oxidative processes were used to degrade 6-mercaptopurine (6-MP), a commonly used cytostatic agent. To degrade 6-MP, three processes were applied: photolysis (UV-C), photocatalysis (UV-C/TiO2), and their combination with H2O2. Each process was performed with variable initial pH (3.5, 7.0, and 9.5). Pilot-scale reactors were used, using UV-C lamps as a radiation source. Kinetic calculations for the first 20 min of reaction proved the significance of the addition of H2O2: In UV-C experiments, the highest k was reached under pH 3.5, k = 0.0094 min−1, while under UV-C/H2O2, k = 0.1071 min−1 was reached under the same initial pH; similar behavior was observed for photocatalysis since k values of 0.0335 and 0.1387 min−1 were calculated for UV-C/TiO2 and UV-C/TiO2/H2O2 processes, respectively, also under acidic conditions. Degradation percentages here reported for UV-C/H2O2 and UV-C/TiO2/H2O2 processes were above 90% for all tested pH values.
In another contribution [15], Gonçalves et al. investigated the impact of different oxidation processes on the maprotiline degradation pathways via liquid chromatography–high-resolution mass spectrometry (LC/HRMS) experiments. Semiconductors photocatalysts—namely, Fe–ZnO, Ce–ZnO, and TiO2—proved to be more efficient than heterogeneous photo-Fenton processes in the presence of hydrogen peroxide and persulfate. No significant differences were observed in the degradation pathways in the presence of photocatalysis, while the SO4-mediated process promoted the formation of different transformation products (TPs). Species resulting from ring openings were observed with higher persistence in the presence of SO4. In silico tests on mutagenicity, developmental/reproductive toxicity, Fathead minnow LC50, D. Magna LC50, and fish acute LC50 were carried out to estimate the toxicity of the identified transformation products. Low toxicant properties were estimated for TPs resulting from hydroxylation onto the bridge rather than aromatic rings, as well as those resulting from ring openings.
In the contribution of Tejera et al. [16], an integral treatment process for landfill leachate reverse osmosis concentrate (LLROC) was designed and assessed, aiming to reduce organic matter content and conductivity, as well as to increase its biodegradability. The process consisted of three steps. First, a coagulation/flocculation treatment (removal of the 76% chemical oxygen demand (COD), 57% specific ultraviolet absorption (SUVA), and 92% color) was utilized. In the second step, a photo-Fenton process resulted in enhanced biodegradability (i.e., the ratio between the biochemical oxygen demand (BOD5) and the COD increased from 0.06 to 0.4), and an extra 43% of the COD was removed at the best-trialed reaction conditions. A UV-A light-emitting diode lamp was tested and compared with conventional high-pressure mercury vapor lamps, achieving a 16% power consumption reduction. Finally, an optimized 30 g L−1 lime treatment was implemented, which reduced conductivity by 43%, and the contents of sulfate, total nitrogen, chloride, and metals by 90%. Overall, the integral treatment of LLROC achieved the removal of 99.9% color, 90% COD, 90% sulfate, 90% nitrogen, 86% Al, 77% Zn, 84% Mn, 99% Mg, and 98% Si, in addition to significantly increasing biodegradability up to BOD5/COD = 0.4.
In the presented paper by Selvanathan et al. [17], the green synthesis of nickel oxide nanoparticles using phytochemicals from three different sources was employed to synthesize nickel oxide nanoparticles (NiOx NPs) as an efficient nanomaterial-based electrocatalyst for water splitting. Nickel (II) acetate tetrahydrate was reacted in presence of aloe vera leaves extract, papaya peel extract, and dragon fruit peel extract, respectively, and the physicochemical properties of the biosynthesized NPs were compared with sodium hydroxide (NaOH)-mediated NiOx. Based on the average particle size calculation from Scherrer’s equation, using X-ray diffractograms and field-emission scanning electron microscope analysis revealed that all three biosynthesized NiOx NPs had smaller particle sizes than that synthesized using the base. Aloe-vera-mediated NiOx NPs exhibited the best electrocatalytic performance, with an overpotential of 413 mV at 10 mA cm−2 and a Tafel slope of 95 mV dec−1. Electrochemical surface area (ECSA) measurement and electrochemical impedance spectroscopic analysis verified that the high surface area, efficient charge-transfer kinetics, and higher conductivity of aloe-vera-mediated NiOx NPs contribute to its low overpotential values.
In conclusion, this Special Issue entitled “Heterogeneous Photocatalysis: A Solution for a Greener Earth” gives an overview of the latest advances in the development of innovative photocatalytic processes. These studies pave the path for the development of innovative processes for a greener Earth.
Finally, we are grateful to all authors for their valuable contributions and to the editorial team of Catalysts for their kind support, especially to Pamela Li for her constant help and availability during this Special Issue submission and publication process.

Funding

This research received no external funding.

Acknowledgments

J.G.M. and S.D.L. are grateful to F.R.S.-F.N.R.S. for, respectively, his postdoctoral position and her Senior Research Associate position.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Almeida, L.A.; Dosen, A.; Viol, J.; Marinkovic, B.A. TiO2-Acetylacetone as an Efficient Source of Superoxide Radicals under Reduced Power Visible Light: Photocatalytic Degradation of Chlorophenol and Tetracycline. Catalysts 2022, 12, 116. [Google Scholar] [CrossRef]
  2. Pascariu, P.; Cojocaru, C.; Airinei, A.; Olaru, N.; Rosca, I.; Koudoumas, E.; Suchea, M.P. Innovative Ag–TiO2 Nanofibers with Excellent Photocatalytic and Antibacterial Actions. Catalysts 2021, 11, 1234. [Google Scholar] [CrossRef]
  3. Mancuso, A.; Navarra, W.; Sacco, O.; Pragliola, S.; Vaiano, V.; Venditto, V. Photocatalytic Degradation of Thiacloprid Using Tri-Doped TiO2 Photocatalysts: A Preliminary Comparative Study. Catalysts 2021, 11, 927. [Google Scholar] [CrossRef]
  4. Benkhennouche-Bouchene, H.; Mahy, J.G.; Wolfs, C.; Vertruyen, B.; Poelman, D.; Eloy, P.; Hermans, S.; Bouhali, M.; Souici, A.; Bourouina-Bacha, S.; et al. Green Synthesis of N/Zr Co-Doped TiO2 for Photocatalytic Degradation of p-Nitrophenol in Wastewater. Catalysts 2021, 11, 235. [Google Scholar] [CrossRef]
  5. Eun, S.R.; Mavengere, S.; Cho, B.; Kim, J.S. Photocatalytic Reactivity of Carbon–Nitrogen– Sulfur-Doped TiO2 Upconversion Phosphor Composites. Catalysts 2020, 10, 1188. [Google Scholar] [CrossRef]
  6. Mahy, J.G.; Sotrez, V.; Tasseroul, L.; Hermans, S.; Lambert, S.D. Activation Treatments and SiO2/Pd Modification of Sol-Gel TiO2 Photocatalysts for Enhanced Photoactivity under UV Radiation. Catalysts 2020, 10, 1184. [Google Scholar] [CrossRef]
  7. Mahy, J.G.; Lejeune, L.; Haynes, T.; Lambert, S.D.; Marcilli, R.H.M.; Fustin, C.A.; Hermans, S. Eco-Friendly Colloidal Aqueous Sol-Gel Process for TiO2 Synthesis: The Peptization Method to Obtain Crystalline and Photoactive Materials at Low Temperature. Catalysts 2021, 11, 768. [Google Scholar] [CrossRef]
  8. de Ceglie, C.; Pal, S.; Murgolo, S.; Licciulli, A.; Mascolo, G. Investigation of Photocatalysis by Mesoporous Titanium Dioxide Supported on Glass Fibers as an Integrated Technology for Water Remediation. Catalysts 2022, 12, 41. [Google Scholar] [CrossRef]
  9. Ghazzy, A.; Yousef, L.; Al-Hunaiti, A. Visible Light Induced Nano-Photocatalysis Trimetallic Cu0.5Zn0.5-Fe: Synthesis, Characterization and Application as Alcohols Oxidation Catalyst. Catalysts 2022, 12, 611. [Google Scholar] [CrossRef]
  10. Mahy, J.G.; Mbognou, M.H.T.; Léonard, C.; Fagel, N.; Woumfo, E.D.; Lambert, S.D. Natural Clay Modified with ZnO/TiO2 to Enhance Pollutant Removal from Water. Catalysts 2022, 12, 148. [Google Scholar] [CrossRef]
  11. Dao, D.Q.; Nguyen, T.K.A.; Pham, T.T.; Shin, E.W. Synergistic Effect on Photocatalytic Activity of Co-Doped Nitio3/g-C3n4 Composites under Visible Light Irradiation. Catalysts 2020, 10, 1332. [Google Scholar] [CrossRef]
  12. Bazta, O.; Urbieta, A.; Trasobares, S.; Piqueras, J.; Fernández, P.; Addou, M.; Calvino, J.J.; Hungría, A.B. In-Depth Structural and Optical Analysis of Ce-Modified Zno Nanopowders with Enhanced Photocatalytic Activity Prepared by Microwave-Assisted Hydrothermal Method. Catalysts 2020, 10, 551. [Google Scholar] [CrossRef]
  13. Faccani, L.; Ortelli, S.; Blosi, M.; Costa, A.L. Ceramized Fabrics and Their Integration in a Semi-Pilot Plant for the Photodegradation of Water Pollutants. Catalysts 2021, 11, 1418. [Google Scholar] [CrossRef]
  14. González-Burciaga, L.A.; García-Prieto, J.C.; García-Roig, M.; Lares-Asef, I.; Núñez-Núñez, C.M.; Proal-Nájera, J.B. Cytostatic Drug 6-Mercaptopurine Degradation on Pilot Scale Reactors by Advanced Oxidation Processes: Uv-c/H2O2 and Uv-c/TiO2/H2O2 Kinetics. Catalysts 2021, 11, 567. [Google Scholar] [CrossRef]
  15. Gonçalves, N.P.F.; Varga, Z.; Nicol, E.; Calza, P.; Bouchonnet, S. Comparison of Advanced Oxidation Processes for the Degradation of Maprotiline in Water—Kinetics, Degradation Products and Potential Ecotoxicity. Catalysts 2021, 11, 240. [Google Scholar] [CrossRef]
  16. Tejera, J.; Hermosilla, D.; Miranda, R.; Gascó, A.; Alonso, V.; Negro, C.; Blanco, Á. Assessing an Integral Treatment for Landfill Leachate Reverse Osmosis Concentrate. Catalysts 2020, 10, 1389. [Google Scholar] [CrossRef]
  17. Selvanathan, V.; Shahinuzzaman, M.; Selvanathan, S.; Sarkar, D.K.; Algethami, N.; Alkhammash, H.I.; Anuar, F.H.; Zainuddin, Z.; Aminuzzaman, M.; Abdullah, H.; et al. Phytochemical-Assisted Green Synthesis of Nickel Oxide Nanoparticles for Application as Electrocatalysts in Oxygen Evolution Reaction. Catalysts 2021, 11, 1523. [Google Scholar] [CrossRef]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mahy, J.G.; Lambert, S.D. Heterogeneous Photocatalysis: A Solution for a Greener Earth. Catalysts 2022, 12, 686. https://doi.org/10.3390/catal12070686

AMA Style

Mahy JG, Lambert SD. Heterogeneous Photocatalysis: A Solution for a Greener Earth. Catalysts. 2022; 12(7):686. https://doi.org/10.3390/catal12070686

Chicago/Turabian Style

Mahy, Julien G., and Stéphanie D. Lambert. 2022. "Heterogeneous Photocatalysis: A Solution for a Greener Earth" Catalysts 12, no. 7: 686. https://doi.org/10.3390/catal12070686

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