Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Recent advances in removal of pharmaceutical pollutants in wastewater using metal oxides and carbonaceous materials as photocatalysts: a review

Suneel Kumar Srivastava *
Department of Chemistry, Indian Institute of Technology, Kharagpur-721302, India. E-mail: suneel@chem.iitkgp.ac.in; suneelchemkgp@gmail.com

Received 22nd August 2023 , Accepted 14th December 2023

First published on 31st January 2024


Abstract

The pharmaceuticals industry has played an important role in developing medicines for improving health and quality of life in treating humans and animals around the world. But it is also considered to be one of the sources of pollutants entering deliberately or accidentally into global water bodies causing toxicity that eventually threatens human health, aquatic organisms and environments even at low concentrations. These contaminants are non-biodegradable and cannot be completely removed from various water matrices following conventional treatment methods. In this regard, photodegradation techniques involving modified/unmodified semiconducting materials have attracted a lot of attention as a promising solution in achieving complete antibiotic degradation with the generation of non-toxic by-products. In view of this, the present review article summarizes current research progress in the removal of several emerging contaminants, such as acetaminophen, amoxicillin, sulfamethoxazole, norfloxacin, ibuprofen, ciprofloxacin, tetracycline, diclofenac and atenolol in water. Considerable emphasis has been placed on metal oxides and carbon-based photocatalysts following their modification through doping with metals and non-metals, metal loading, the formation of composites, immobilization and heterostructure/heterojunction approaches. Finally, the review ends with future prospects for nanomaterial-based heterogeneous photocatalysts in the removal of pharmaceutical contaminants from water.


image file: d3lf00142c-p1.tif

Suneel Kumar Srivastava

Suneel Kumar Srivastava received his Ph.D degree from the Indian Institute of Technology, Kharagpur in 1984. He is a former Professor in the Department of Chemistry of the same Institute, serving from 1986 to 2021. Dr Srivastava carried out his post-doctoral work as a DAAD Fellow in the Technical University, Karlsruhe (1988–89, 2002, 2006), University of Siegen (1994, 1999), Technical University, Munchen (2009), Leibniz Institute of Polymer Research, Dresden (2013) Germany, and University of Nantes, France (2003, 2007). His research interests are in the field of nondimensional nanomaterials for their application in the fields of energy, environments and polymer nanocomposites. Dr Srivastava has guided 23 Ph.Ds, published about 200 research papers in referred journals, contributed to 16 chapters in books and edited 2 books.


1 Introduction

Water plays an essential role in sustaining a cherished healthy life for living organisms as well as ecosystems. Therefore, the purity of water remains of utmost concern for the survival of human beings, plants, animals and several other living species in the world. A report presented by UNESCO at the UN 2023 Water Conference revealed the non-availability of safe drinking water for 26% of the global population.1 This problem is also compounded by the presence of several pollutants in water bodies. This contributes to the depletion of fresh water, resulting in an overall water crisis worldwide.2 This adversely affects human health, several other living organisms and sustainable social development. According to an estimate, about 80% of wastewater is discharged globally into the environment without any prior treatment, jeopardizing human health, the ecosystem, and the environment.3 In this regard, dye effluents, heavy metals and pesticides discharged as wastewater from different industries contribute significantly to water pollution.4–12

In addition, the wide application of pharmaceuticals in daily life for the treatment of complex diseases is also the major contributor of emerging contaminants, with potential adverse effects on humans and the aquatic environment.13–22 The presence of these pharmaceutical pollutants could lead to cancers, severe bleeding, organ damage, birth defects, reproductive disorders, endocrine disorders, and mild to severe toxic effects in human beings in the global population.14 The toxic effects are also threats to mammals, other organisms, and the ecosystem. Fig. 1 shows the effect of pharmaceuticals in reducing the quality of water.14 The presence of these pharmaceutical pollutants in water through improper disposal, irrigation of crops, and consumption by agriculture, humans, and animals seriously affects the ecosystem.


image file: d3lf00142c-f1.tif
Fig. 1 Routes of pharmaceutical contaminants (PCs). Reproduced from ref. 14 with permission from Elsevier (2022).

Further, the accumulation of antibiotic drugs in water can result in the development of antibiotic-resistant bacteria and the dissemination of antibiotic-resistant genes in humans and other living organisms.15,16 According to a recent report, urban wastewater treatment plants are recognized sources for the dissemination of antibiotic resistance in the environment.17 In view of the rising effects of this antibiotic resistance on the global population, the removal of these bioactive molecules from the environment is important to slow down the growth of resistant microorganisms. In addition, antibiotic residues absorbed by plants could interfere with physiological processes, leading to potential ecotoxicological effects.18 These contaminants cannot be completely removed from various water matrices by conventional chemical, physical, flocculation, reverse osmosis or a few other processes, due to the formation of secondary pollutants, high cost, and operational time.19 Therefore, the development of cost-effective, eco-friendly, economical, and effective technologies is urgently needed to remove these emerging contaminants, due to the rising effects of antibiotic resistance in aquatic environments.

Design of the surface and interface plays a promising role in the performance of photocatalysts through maximizing the efficacy of catalysts. Therefore, heterogeneous photocatalysis has been receiving considerable attention as one of the most attractive, low-cost, efficient and outstanding approaches in the degradation of pharmaceutical pollutants.19–55 In this regard, a considerable amount of research interest has focused mostly on TiO2 and to some extent on other semiconducting materials and transition metal oxides as photocatalysts in the degradation of pharmaceutical pollutants in water.23–39 The choice of semiconducting metal oxides as photocatalysts is motivated by the availability of a renewable energy source (solar energy) and the generation of non-toxic degradation products (chemicals and gases). They can be commonly prepared by sol–gel, hydrothermal, solvo-thermal, microwave heating, wet chemical, physical vapour deposition and chemical vapour deposition methods.30 However, the potential of TiO2 and other semiconducting metal oxides could not be harnessed due to the higher rate of recombination of electron–hole pairs and its limited photocatalytic activity under visible light exposure.

Recently, carbonaceous materials have also been reported as promising materials for use in the photocatalytic degradation of antibiotics in water.40–50 This is facilitated by combining these carbon-based materials with other semiconductors, which is considered to be an outstanding approach to enhancing photocatalytic performance. In order to facilitate this, carbonaceous materials with different structures and properties are used as additives in semiconductor materials. This invariably results in enhanced charge separation and visible light activity and is considered the best solution. In addition, semiconducting metal oxides and carbonaceous materials are subjected to doping with metals, non-metals, metal oxides, coupling with noble metal nanoparticles and the formation of composites.36,39,49 Other approaches involving immobilization and the formation of a heterojunction are reported as imperative alternative strategies for achieving enhanced photocatalytic efficiency for these photocatalysts in water treatment.51

According to the available literature, several reviews have been published focusing on metal oxides,23–30 TiO2,31–33 ZnO-based photocatalysts,34 semiconductors,35 doped TiO2,36 hybrids,37 TiO2–carbon dot nanocomposites,38 plasmonic metal–TiO2 composites,39 carbonaceous/carbon-based materials,40,41 g-C3N4,42 MWCNT,43 carbon dots,38,44 activated carbon,45 graphene-based composites,46–48 graphene–TiO2 and doped graphene–TiO2 nanocomposites,49 graphene-based materials,50 and nanomaterial-based heterogeneous photocatalysts51 as photocatalysts for the treatment of wastewater containing pharmaceuticals. Alternatively, several review articles have reported on the photodegradation of antibiotic contaminants in water, such as amoxicillin,21 ibuprofen,22 tetracycline,52,54 ciprofloxacin,53,54 and norfloxacin54 antibiotics in wastewater and several others, which are referred to in section 3. However, there is still a need for an extensive review article in this field, covering in a single window a larger number of pharmaceutical pollutant photocatalysts for their photocatalytic performance.

The present review is focused primarily on the photocatalytic degradation of acetaminophen, amoxicillin, sulfamethoxazole, ibuprofen, norfloxacin, ciprofloxacin, tetracycline, diclofenac, etc. The structure and uses as well as the solubility of these antibiotics in water are provided in Table 1 (ref. 55) and ESI, respectively. In view of this, the article describes the fundamental properties of semiconducting materials as photocatalysts as well as role of metal oxides, carbon-based materials, and heterojunctions and the immobilization approaches employed and the mechanisms involved in the removal of these pharmaceutical pollutants. Subsequently, the article deals with the removal of the above-mentioned drugs from contaminated water using semiconducting TiO2, ZnO, and many other oxides, their combination with graphitic-carbon nitride (g-C3N4), carbon nanotubes (CNTs), activated carbon (AC), graphene oxide, graphene and graphene quantum dots, doping with metals and nonmetals, the formation of composites, semiconducting materials deposited on certain supports as photocatalysts and a heterojunction approach. It is anticipated that, in the light of this, the current review could be of immense help in identifying cost-effective and efficient photocatalytic methods for the remediation of these pharmaceutical pollutants. In addition, various research gaps, their possible solutions and several future prospects are also provided at the end of this article for the possible enhancement of environmental conservation.

Table 1 Structure and uses of different pharmaceutical pollutants. Adopted from PubChem55
Pollutant (formula) Structure Uses
Acetaminophen (C8H9NO2) image file: d3lf00142c-u1.tif Nonprescription analgesic and antipyretic medication for mild-to-moderate pain and fever
Amoxicillin (C16H19N3O5S) image file: d3lf00142c-u2.tif Bacterial infections, and dental abscesses
Sulfamethoxazole (C10H11N3O3S) image file: d3lf00142c-u3.tif Used in treatment of a variety of bacterial infections, including those of the urinary, respiratory, and gastrointestinal tracts
Ibuprofen (C13H18O2) image file: d3lf00142c-u4.tif Anti-inflammatory; analgesic; antipyretic
Norfloxacin (C16H18FN3O3) image file: d3lf00142c-u5.tif In treatment of urinary tract infections and prostatitis
Ciprofloxacin (C17H18FN3O3) image file: d3lf00142c-u6.tif Therapy of mild-to-moderate urinary and respiratory tract infections caused by susceptible organisms
Tetracycline (C22H24N2O8) image file: d3lf00142c-u7.tif Role as an antimicrobial agent, an antibacterial drug, an antiprotozoal drug, a protein synthesis inhibitor and an Escherichia coli metabolite
Diclofenac (C14H11Cl2NO2) image file: d3lf00142c-u8.tif Therapy of chronic forms of arthritis and mild-to-moderate acute pain
Atenolol (C14H22N2O3) image file: d3lf00142c-u9.tif As a cardioselective beta-blocker that is widely used in the treatment of hypertension and angina pectoris


2 Important photocatalysts and their role in the removal of pharmaceutical pollutants

The primary mechanism for the degradation of organic pollutants by a semiconducting material involves irradiating it with light energy in the form of photons (hv) sufficiently greater than the band gap energy of the photocatalyst (Fig. 2 (ref. 37)). Holes (hVB+) and electrons (eCB) are generated in this manner in the valence band (VB) and the conduction band (CB), respectively. The separated holes reacts with hydroxyl ions (OH) or water molecules (H2O) to produce hydroxyl radicals (·OH). In addition, the separated electrons reacts with dissolved O2 in water to produce superoxide radicals (·O2), which upon further reaction, produce ·OH.37,51 Subsequently, the active species generated in this manner react with pharmaceutical pollutants on the surface of the semiconductor catalyst to give H2O, CO2 and other by-products.
Semiconductor + hv → hVB+ + eCB

hVB+ + H2O → H+ + ·OH

eCB + O2 → ·O2

·O2 + H+ → HO2·

HO2· + HO2· → H2O2 + O2

H2O2 + ·O2 → ·OH + OH + O2

H2O + hVB+ → ·OH + H+

hVB+ + OH → ·OH
It should be mentioned that the efficiency of a photocatalytic reaction depends on the capability of the photocatalyst to generate longer-lived e and h+ that lead to the formation of reactive free radicals. In addition, photodegradation efficiency also depends on catalyst loading, contaminant concentration, pH, the presence of ions in the water, hydrogen peroxide, ultrasound irradiation, bubbling of O2 and N2 into the solution and irradiation time.13,26,34

image file: d3lf00142c-f2.tif
Fig. 2 Photocatalytic processes over a heterogeneous photocatalyst. Reproduced from ref. 37 with permission from MDPI (2021).

2.1 Metal oxides

Several semiconductor metal oxides have been used as photocatalysts in the abatement of aqueous pollution due to organic pollutants. From this point of view, TiO2 has received a considerable amount of attention and its choice is mainly guided by its superior photocatalytic degradation efficiency, low processing cost, high environmental stability, nontoxicity, chemical stability, and high oxidizing ability.31–33 However, its wide band gap (∼3–3.2 eV),32 and the fast e–h+ recombination rate of photogenerated electron–hole pairs in TiO2 limit its applications. Semiconducting ZnO (band gap: 3.37 eV) has been used as another photocatalyst in water treatment as an alternative to TiO2.56 Several other metal oxides (ZrO2, Fe2O3, γ-Fe3O4, SnO2, Mn2O3, WO3, CeO2, CuO, and NiO) have also been investigated as alternatives to TiO2 and ZnO.26 Nano-engineered metal-oxide-based photocatalysts have also attracted a lot of attention in wastewater treatment.57 However, metal oxide catalysts experience similar drawbacks to TiO2. As a consequence, significant developments have taken place in recent years in tailoring these metal oxide photocatalysts. This is achieved by reducing their band gap by the addition of dopants that include both metals and non-metals, such as iron, copper, carbon, nitrogen, platinum and sulfur. In addition, metal sulfides,58 metal ferrites,59 and oxychlorides60 have also been explored as emerging photocatalysts for the removal of pharmaceutical pollutants.

Photocatalytic studies have been reported on the performance of semiconductor–metal composites in the removal of several pollutants from water. In this regard, plasmonic composites in combination with various semiconducting photocatalysts have been widely studied for enhancing overall photocatalytic performance.61,62 The improved photocatalytic efficiency is attributed to the surface plasmon resonance effect. In addition, metal nanoparticles can decrease the recombination rate of the photo-induced e–h+ pairs of the semiconductor material by effective electron trapping in the conduction band. Metal oxide nanocomposites derived from a mixture of two or more oxides or between these oxides and other functional semiconductor materials have also been found to be efficient, economical, and environmentally friendly photocatalysts in water pollutant remediation.63,64

2.2 Carbonaceous materials

The photocatalytic performance of various carbonaceous materials has been receiving more attention for antibiotic removal owing to their intriguing properties and good stability.40,41 The choice of these carbonaceous materials in removing antibiotics is mainly guided by simple and cost-effective synthesis methods, the easy availability of raw materials and their unique physiochemical properties, such as the presence of micropores, mesopores, and macropores, the large number of oxygen-functional groups, high porosity, and high surface area, coupled with good visible-light adsorption ability, chemical stability, excellent electrical conductivity and high intrinsic electron mobility.40 The carbonaceous materials explored for this purpose include carbon dots,38 g-C3N4,42,65 activated carbon45,66 and carbon nanotubes (CNTs).67 Graphene is another carbon-based material composed of a one-atom-thick layer of carbon atoms arranged in a hexagonal lattice.68 It is a semimetal with a small degree of overlap between the valency band and the conduction band.69 This makes graphene a promising candidate for application in photocatalysis. However, the photocatalytic performances and practical applications of carbon-based materials have not been encouraging, due to poor solar-light absorption and the rapid recombination of photogenerated electron–hole pairs.41 Interestingly, combinations of these carbon-based materials with other semiconductor metal oxides have been utilized as promising photocatalysts owing to their notable properties like stability, conductivity, durability and high absorptivity. In addition, carbon-based materials–metal oxide nanocomposites have also enhanced the degradation efficiency of pharmaceuticals by improving the generation of radical species, through improved surface area and light absorption, and reducing the recombination of generated charge carriers.48,69

2.3 Heterojunction nanocomposites as photocatalysts

A heterojunction is defined as the interface between two layers or regions of different semiconductors with unequal band structures that can result in band alignments. Based on this concept, semiconductor–semiconductor-based heterojunction composites showed excellent improvements in photocatalytic efficiency. This is ascribed to minimized charge carrier recombination, the interface of the heterojunction, superior charge transfer, prolonged charge carrier lifetime, separate active sites, and extended light absorbance characteristics.51 These semiconductor heterojunction photocatalysts are classified into several types: i.e., conventional heterojunctions (type-I, type-II, and type-III), p–n heterojunctions, direct Z-scheme heterojunctions, and S-scheme heterojunctions.70–73 The schematic separation of charges via electron migration from one semiconductor to another in various heterojunction mechanisms is represented in Fig. 3.51 Among these, in a type-I heterojunction, the VB and CB of semiconductor-1 are respectively lower and higher than those of semiconductor-2 (Fig. 3(a)). The photogenerated holes migrate from the VB of semiconductor-1 to the VB of semiconductor-2 accompanied by the transfer of photoelectrons from the CB of semiconductor-1 to the CB of semiconductor-2.52 However, this type-I heterojunction cannot spatially separate e–h+ pairs and this leads to the accumulation of charge carriers and their accelerated recombination rate. A type-II heterojunction (Fig. 3(b)) involves the transfer of photogenerated holes generated in semiconductor-2 to semiconductor-1, considering the VB of semiconductor-1 to be lower than that of semiconductor-2 on irradiating with light.52 In contrast, photogenerated electrons in the CB of semiconductor-1 can migrate to that of semiconductor-2, if the level of the CB in semiconductor-1 is higher than that of semiconductor-2. It should be noted that the spatial separation of electron–hole pairs can occur in a type-II heterojunction. Furthermore, the structure of a type-III heterojunction is similar to that of a type-II heterojunction; however, charge-carrier separation cannot occur in a type-III heterojunction because the band gaps of both semiconductors do not overlap, since the levels of the VB and CB of both semiconductors are very far apart (Fig. 3(c)). When p-type and n-type semiconductors are combined, a p–n heterojunction can be formed. A space-charge region could be formed at the interface before light irradiation due to diffusion of the majority of charge carriers, leading to a built-in electric field, as shown in Fig. 3(d). In the Z-scheme heterojunction system, the band structure is quite analogous to that of a type-II heterojunction, but the direction of charge transfer is the opposite. The photogenerated electrons from the second semiconductor migrate aggressively to the VB of the first semiconductor and occupy the available holes, while the strongly oxidative holes in the VB of the second semiconductor and strongly reductive electrons in the CB of the first semiconductor take part in the redox reaction (Fig. 3(e)). In a step-scheme (S-scheme) heterojunction, two n-type semiconductors are combined with a staggered band structure similar to a type-II heterojunction (Fig. 3(f)).
image file: d3lf00142c-f3.tif
Fig. 3 Schematic illustration of various types of heterojunction: (a) straddling bandgap (type I), (b) staggered bandgap (type II), (c) broken bandgap (type III), (d) p–n type, (e) direct Z-scheme, and (f) S-scheme. Reproduced from ref. 51 with permission from Amer Sci Publ (2023).

2.4 Immobilized photocatalysts

The immobilization of photocatalysts on supports (Fig. 4)51 can maximize the activity of semiconductors by offering a greater number of active sites. The high photocatalytic activity of such immobilized semiconductor photocatalysts is guided by the properties of their semiconductor-active species and the kind of support employed.51 The high catalytic performance of these immobilized photocatalysts originates from impeding the rate of electron–hole pair recombination. The recovery, reusability, and stability issues of a photocatalyst remain challenging after several reaction runs. In this regard, the immobilization of a catalyst on a support facilitates the rapid separation and efficient recycling of the catalyst. This reduces production costs as well as minimizing waste generation, especially in industrial applications compared to conventional pure photocatalysts.74
image file: d3lf00142c-f4.tif
Fig. 4 Supporting materials used for the immobilization of photocatalysts. Reproduced from ref. 51 with permission from Amer Sci Publ (2023).

3 Removal of pharmaceutical components using different Photocatalysts

In this review article, we present the use of photocatalysts based on bare metal oxides (TiO2, ZnO and other oxides) and carbon-based materials (graphitic carbon nitride, g-C3N4, carbon nanotubes CNTs, activated carbon AC, and graphene) in the removal of pharmaceutical pollutants from water. In addition, several modification approaches are also highlighted and those involving metal loading, doping with metals and nonmetals, the formation of composites, immobilization and the formation of heterojunctions for this purpose are described below for pharmaceutical pollutants.

3.1 Acetaminophen

Acetaminophen (ACT), also known as paracetamol is commonly used all over the world as a painkilling, anti-inflammatory, analgesic, and antipyretic drug.75–78 It is available both as a single-entity formulation and in combination with other medications. The presence of acetaminophen in wastewater, surface water and groundwater can have an adverse effect on living organisms and environmental ecology owing to its oxidative transformation to toxic N-acetyl-p-benzoquinone imine. The stable chemical structure of acetaminophen remains one of the major constraints to its removal through conventional wastewater treatment. Therefore, attention has focused on its removal from aqueous media following a photocatalysis approach, as described below.79–147
3.1.1 Metal oxides. Two titania photocatalysts prepared by a sol–gel method showed higher photocatalytic activity than commercial TiO2–P25 when tested for the photodegradation of paracetamol in aqueous solution.79 Marizcal-Barba et al.80 studied the photocatalytic degradation of acetaminophen in the presence of TiO2 synthesized by a sol–gel method and observed its 99% degradation of acetaminophen corresponding to a pH of 10, acetaminophen concentration of 35 mg L−1 and a catalyst dose of 0.15 g of TiO2. Hollow mesoporous TiO2 microspheres have also been investigated as a photocatalyst to study the degradation of acetaminophen in water owing to its large surface area and the possibility of efficient light harvesting capability.81 These findings showed an increase in the conversion fraction of the drug to 94% in 60 min following a 25% increase in the initial reaction rate and good photodegradation activity even after 10 repeated runs.

Zhang et al.82 reported about 95% photocatalytic degradation of acetaminophen in an aqueous solution of TiO2 (1.0 g L−1) after 100 min of irradiation under a 250 W metal halide lamp. This is attributed to direct hole (h+) oxidation and ipso-substitution comprising the main initial steps in the degradation. The photodegradation of paracetamol (20 mg L−1) has been investigated in the presence of nanostructured TiO2 catalysts with a nanotube-type morphology using ultraviolet radiation (λ: 254 nm) and the removal efficiency was found to be 99% after 100 min.83 The photocatalytic degradation of acetaminophen in water has also been reported using ZnO,84 faceted-TiO285 and molecularly imprinted ZnO nanonuts.86

3.1.2 Metal-incorporated metal oxides. The introduction of metal species into TiO2 and other metal oxides could modify their structural, electronic, optical and morphological properties. In view of this, several studies have been reported on the photodegradation of pharmaceutical pollutants in metal-loaded metal oxides. Jiménez-Salcedo et al.87 applied an organometallic approach for the preparation of Au–TiO2 nanohybrids and studied the degradation of paracetamol (0.3 mg L−1) under UVA light. These studies revealed 100% degradation of paracetamol in 30 min for Au–TiO2 photocatalysts compared to TiO2 (40 min). The kinetic studies also supported these findings as being inevitable from the higher rate constant of Au–TiO2 photocatalysts (0.14 min−1) compared to TiO2 photocatalysts (0.12 min−1) in the degradation of paracetamol. In addition, Ag-, Au- and Pt-loaded TiO2 (Ag/TiO2, Au/TiO2 and Pt/TiO2) have shown significant enhancement in the photocatalytic degradation (>90%) of acetaminophen in water over a wide pH range (4.2–8.0) under solar light.88

Pd-decorated CuO nanostructured thin film showed enhanced visible-light degradation of acetaminophen.89 The influence of radical trappers revealed no role for ·OH, ·O2 (or 1O2) radicals on the photocatalytic degradation of acetaminophen. The photocatalyst possessed good stability, as indicated by the observed insignificant change in photodegradation even after 5 cycles. According to the available literature, ZnFe2O4 (bandgap: 1.9 eV) is non-toxic and exhibits good photostability.90 Its photocatalytic behaviour is guided by several factors, such as its preparative method, morphology, and the presence of impurities. In view of this, Huerta-Aguilar et al.91 reported the efficient degradation of paracetamol during water treatment using Au nanoparticles grown on ZnFe2O4 as a visible light (200 W halogen lamp, C-type R7s, λ > 400 nm) assisted photocatalyst. TiO2/BN/Pd nanofibers showed significantly enhanced degradation of ACT (>90%), compared to pure TiO2 (20%) after 4 h under visible-light irradiation.92 This was explained on the basis of the good dispersion of Pd nanoparticles on TiO2–BN nanofibers to facilitate the transfer of photoexcited hole carriers and a decrease in photogenerated electron–charge recombination. Reusability studies and recycling tests on the TiO2/BN/Pd photocatalyst indicated its good stability over 5 cycles under UV and visible light.

3.1.3 Doped metal oxides. C,N-co-doped TiO2 (20 mg) degraded 69.31% paracetamol (4 mg L−1) under UV light and 70.39% under solar light in 120 min.93 According to Shaban and Fallata,94 carbon-doped TiO2 nanoparticles (2.0 g L−1) successfully photocatalytically degraded acetaminophen (2 ppm) in aqueous solution, seawater, and real polluted seawater on irradiation with UV and natural sunlight. This enhancement could be attributed to the lowering of its bandgap as a result of carbon doping in TiO2. In addition, Mg-doped TiO2 has also been reported in the photodegradation of paracetamol.95 Accordingly, 25 wt% Mg-doped TiO2 produced 60% and 48.3% degradation of paracetamol under UV and visible light, respectively. In all likelihood, the Mg dopant in TiO2 acts as a photosensitizer for photocatalysts and hinders the recombination of electron–hole pairs. In another study, TiO2 and Ta-doped TiO2 nanomaterials showed 70–80% degradation of paracetamol in 2 h in UV-irradiated aqueous suspensions, which was attributed to surface acidity as a key parameter.96 Mn-doped TiO2 exhibited 53% degradation of an aqueous solution of acetaminophen in 3 h under ultrasound and UV irradiation owing to the reduced band gap (1.6 eV) and the high surface area (158 m2 g−1).97 Fe-doped TiO2,98 KAl(SO4)2 and NaAlO2-doped TiO2,99 N-doped halloysite (HNT)/TiO2,100 carbon-self-doped TiO2,101 Bi3+-doped TiO2102 and Ba0.95Bi0.05Fe0.95Cu0.05O3[thin space (1/6-em)]103 have also been prepared and examined for the photocatalytic degradation of acetaminophen and paracetamol.

The degradation of acetaminophen and its reaction mechanism have been investigated in presence of Ag–ZnO104 and La-doped ZnO105 photocatalysts under visible-light irradiation. Abri et al.106 studied the photocatalytic degradation of nizatidine, acetaminophen and levofloxacin over ZnO (1[thin space (1/6-em)]:[thin space (1/6-em)]6) nanostructured photocatalysts under UVB light for 240 min and the findings are displayed in Fig. 5(a). Similar studies on using 1% Ce-doped ZnO produced almost no change in the degradation of acetaminophen and levofloxacin compared to that observed for nizatidine (∼95%), as evidenced from Fig. 5(b). Such different photocatalytic degradation of these pharmaceuticals in the presence of ZnO and 1% Ce–ZnO photocatalysts could be attributed to their chemical structures.


image file: d3lf00142c-f5.tif
Fig. 5 (a) Photocatalytic degradation of pharmaceuticals over (a) ZnO (1[thin space (1/6-em)]:[thin space (1/6-em)]6) and (b) 1% Ce–ZnO nanostructured photocatalysts [experimental conditions: catalyst dosage: 1 mg mL−1; concentration of pharmaceutical: 5 mg L−1]. Reproduced from ref. 106 with permission from Elsevier (2019).

Kumar et al.107 investigated the photocatalytic degradation of acetophenone by irradiating nitrogen-implanted ZnO nanorod arrays (NRAs) with visible light. It should be noted that an N ion (1 × 1016 ions per cm2) doped ZnO NRA sample (referred to as N–ZnO4) showed maximum degradation efficiency (98.46%) of acetaminophen (20 ppm) in the presence of sunlight under 120 minute duration. The linear variation in ln(C0/C) versus irradiation time followed pseudo-first-order degradation kinetics for acetaminophen. Furthermore, the superior photocatalytic activity of the N–ZnO4 catalyst was inevitable from the high value of its rate constant (0.038 min−1) compared to pristine ZnO NRAs (0.0045 min−1). In addition, further investigations also revealed a more or less unaltered degradation efficiency (98.46% to 97.63%) of N–ZnO4 after five repeated cycles. The findings of the effect of scavengers on the photocatalytic degradation of acetaminophen in the presence of N–ZnO4 showed a decrease in degradation efficiency for acetaminophen (98.4%) in the presence of benzoquinone (BQ 28.52%), EDTA (65.6%) and methanol (98.4%) due to the major role played by O2. The mechanism of acetaminophen degradation on subjecting N-ion-implanted ZnO NRAs to visible light suggested a shifting of the band gap to the visible region.

3.1.4 Metal oxide composites. Nanosized Fe2O3–TiO2 nanocomposites exhibited higher degradation (95.85%) of acetaminophen compared to bare TiO2 under stimulated solar radiation (optimal conditions: initial concentration of ACT: 30 mg L−1; catalyst loading: 1.25 g L−1; initial pH: 11).108 Khasawneh et al.109 synthesized a hematite (α-Fe2O3)-doped TiO2 nanocomposite via a sol–gel method and investigated the role of UV light on the degradation of paracetamol. The photocatalytic degradation of acetaminophen has also been investigated using montmorillonite nanosheets modified with TiO2 under UV radiation.110 These findings revealed 100% removal efficiency for acetaminophen in aqueous solution corresponding to pH 7, catalyst dose of 0.75 g L−1, acetaminophen concentration of 2 mg L−1 and contact time within 120 min.

Magnetic TiO2/Fe3O4 (1.16 g L−1) and TiO2/SiO2/Fe3O4 (1.34 g L−1) nanoparticles degraded acetaminophen, antipyrine, caffeine, and metoprolol pharmaceuticals on illuminating its aqueous solution (pH: 7, ACT concentration: 30 mg L−1).111 TiO2/SiO2/Fe3O4 nanoparticles also showed good reusability, as evidenced within four repeated experiments. Czech and Tyszczuk-Rotko112 explored the visible-light (centered at 500–550 nm) driven photocatalytic removal of acetaminophen from water using MWCNT (1.72 wt%)–TiO2–SiO2 nanocomposites and observed ∼82% efficiency due to the key role played by photogenerated holes. In another study, Fernandes et al.113 selected combinations of Fe2O3 and Fe3O4 nanoparticles due to their easy availability and used them in the photodegradation of acetaminophen under UV-vis irradiation. The total acetaminophen (and caffeine) degradation (20 ppm/150 mL) took place by means of 0.13 g catalyst L−1 solution in 45 min (and 60 min) and it remained almost unaltered over five cycles. A ternary heterogeneous anatase-TiO2 (B) biphasic nanowires/Bi4O5I2 composite exhibited 95% degradation of acetaminophen in 6 min under visible-light irradiation.114 This is ascribed to the multiphase structure, including the synergistic effect of anatase TiO2 and Bi4O5I2. A schematic of the possible charge separation and photocatalytic mechanism of the TiO2–Bi4O5I2 composite under visible-light irradiation is displayed in Fig. 6(a).


image file: d3lf00142c-f6.tif
Fig. 6 (a) Schematic of the possible charge separation and photocatalytic mechanism of TiO2–Bi4O5I2 composite under visible-light irradiation. Reproduced from ref. 114 with permission from Elsevier (2020). (b) Schematic diagram of charge transfer in the photoexcited TiO2/Fe2O3 core–shell photocatalyst. Reproduced from ref. 117 with permission from Elsevier (2017).

Chau et al.115 synthesized a Cu2O/WO3/TiO2 ternary composite in view of the narrow band gaps of Cu2O (2.20 eV) and 2.70 eV (WO3) guided by their low cost, nontoxicity, chemical stability and strong absorption ability towards visible light. The composite fabricated in this manner produced 92.50% photodegradation of ACT (1 mg L−1) compared to pure TiO2 under 60 min of solar irradiation. This is attributed to the effective separation and low recombination rate of the charge carriers. The produced composite exhibited high reusability for photodegradation with 83% at the fifth cycle of ACT photodegradation. Nanostructured titania supported on activated carbon (AC) has been used to study the effects of photocatalyst dosage, initial solution pH and irradiation (UV) time on the photocatalytic degradation of aqueous acetaminophen.116 Abdel-Wahab et al.117 prepared flower-like core–shell TiO2/Fe2O3 photocatalysts instead of TiO2/Fe3O4 due to the photostability of Fe2O3 compared to Fe3O4 and investigated its activity in the degradation of paracetamol in aqueous solution using a medium-pressure mercury lamp (450 W). These findings indicated increases in the photocatalytic degradation of paracetamol (52.5%) to 87.8% for 50% content of TiO2. This is ascribed to the separation of the photogenerated electron–hole pairs accomplished by coupling the narrow band gap with the wide band gaps of Fe2O3 and TiO2, respectively. A schematic diagram of charge transfer in the photoexcited TiO2/Fe2O3 core–shell photocatalyst is displayed in Fig. 6(b). Jallouli et al.118 used TiO2 nanoparticles and TiO2/cellulosic fiber to carry out the photocatalytic degradation of paracetamol under UV and sunlight irradiation. WO3/TiO2/SiO2119 and TiO2/ZSM-5 (ref. 120) also exhibited enhanced photocatalytic degradation of acetaminophen in contaminated wastewater.

TiO2 immobilized on glass spheres (sunlight)121 and ZnO–polystyrene (UV-LED)122 photocatalysts effectively removed acetaminophen and paracetamol, respectively. The photodegradation of acetaminophen is also reported with zeolite-supported TiO2 and ZnO under UV and sunlight,123 bi-modified titanate nanomaterials (visible light),124 BaTiO3/TiO2 composite (UV-vis),125 and Ag/AgCl@ZIF-8 (visible light).126

3.1.5 C3N4 and C-dot-based composites. The rapid photocatalytic degradation of acetaminophen (and levofloxacin) targeted by modifying g-C3N4 bulk material to g-C3N4 nanosheets under solar-light irradiation reached 99% in 60 min compared to bulk g-C3N4 (38% in 240 min).127 Such performance of g-C3N4 nanosheets could be assigned to multiple contributions, such as smaller particle size, rich carbon surface and lower band gap. Contemporary studies on exfoliated g-C3N4 have also been reported for the degradation of paracetamol (and ibuprofen) in an aqueous environment under visible light.128 A ZnO/Ph–g-C3N4 nanocomposite acted as an efficient visible-light-active catalyst for the photodegradation of paracetamol in aqueous suspension.129 The findings revealed hydroxyl and superoxide radical anions to be responsible for the degradation process.

Heterostructures comprising α-Fe2O3/g-C3N4130 have been examined for the photocatalytic degradation of acetaminophen. The photocatalytic activity of g-C3N4 combined with UiO-66-NH2 in different proportions (25%-g-C3N4/UiO-66-NH2, 50%-g-C3N4/UiO-66-NH2, 75%-g-C3N4/UiO-66-NH2) was tested for the removal of acetaminophen from an aqueous solution under given experimental conditions ([ACT]: 5 mg L−1, [Cat]: 0.5 g L−1, V: 350 mL).131 The corresponding findings on the temporal evolution of acetaminophen with the different samples and their pseudo-first-order rate constants (kobs) are displayed in Fig. 7(a) and (b). These findings depict complete removal of acetaminophens by the 75%-g-C3N4/UiO-66-NH2 heterostructure in 120 min with a pseudo-first-order rate constant of 2 h−1. It is suggested that incorporation of UiO-66-NH2 in g-C3N4 enhanced the separation of the photogenerated charges. Silica–carbon quantum dots (1 wt%) decorated TiO2 as a sunlight-driven photocatalyst completely removed acetaminophen 33.3% faster than pure TiO2.75 Gupta et al.132 studied the augmented photocatalytic degradation of acetaminophen using hydrothermally treated g-C3N4 and persulfate under LED irradiation.


image file: d3lf00142c-f7.tif
Fig. 7 (a) Photocatalytic degradation of acetaminophen with different g-C3N4/UiO-66-NH2 samples. (b) Pseudo-first-order rate constant (kobs) of different g-C3N4/UiO-66-NH2 samples. Experimental conditions: V = 350 mL; T = 20 °C, CACE = 5 mg L−1; CCAT = 0.5 g L−1. Reproduced from ref. 131 with permission from MDPI (2022). (c) Schematic illustration of the TiNiW NPs decorating the surface of the graphene composites and (d) TiNiW nanoparticle showing the possible chemical reactions for the formation of reactive oxygen species that degrade the ACT contaminant. Reproduced from ref. 138 with permission from Elsevier (2021).
3.1.6 Graphene and its composites. Khavar et al.133 observed the complete degradation of acetaminophen (pH 5.4) for 3 wt% rGO@TiO2 under visible UVA-LED irradiation within 50 min. A graphene-oxide-supported bioinspired CuO photocatalyst (50 wt%) showed 96.2% acetaminophen degradation.134 A calcined ZnFe-layered double hydroxide (CLDH)/rGO (for initial wt. of GO: 30 mg) exhibited the highest degradation of about 95% of paracetamol in 420 min, owing to the synergistic effect between Zn–Fe calcined LDH and rGO.135 Tao et al.136 synthesized nanocomposites comprising 5% graphene/TiO2 nanotubes by a hydrothermal method and observed a 96% degradation rate for acetaminophen (5 mg L−1) under UV-light irradiation for 3 h. Further investigations indicated holes to be the main oxidation species in the photocatalytic process. According to Umejuru et al.,137 coal fly ash (CFA) decorated with graphene oxide nanorods with Pb2+-ion-loaded spent adsorbent exhibited 93% degradation of acetaminophen on subjection to photocatalysis. Ni@TiO2:W nanoparticles (TiNiW) and TiNiW immobilized on the surface of a flexible graphene (FG) composite on subjection to natural solar irradiation (3 h) achieved acetaminophen degradation efficiencies of 100% and 86%, respectively.138 Subsequent findings suggested that acetaminophen degradation was mainly caused by reactive oxygen species, such as ·OH radicals and h+. Reusability experiments confirmed the stability of TiNiW and FG/TiNiW composite for the degradation of acetaminophen. Fig. 7(c) schematically represents the TiNiW nanoparticles decorated on the flexible graphene support and a proposed use in the mechanism of acetaminophen degradation. It is suggested that on subjecting it to solar excitation, photogenerated electrons could be rapidly trapped by the graphene layers, as evident through the scheme displayed in Fig. 7(d). Core/shell rGO/BiOBr139 and vitamin-C-assisted synthesis of rGO–Ag/PANI140 have also been reported to successfully achieve the improved photocatalytic degradation of acetaminophen.
3.1.7 Heterojunctions and Z-scheme-based photocatalysts. Recently, Parida et al.20 fabricated a Bi2O3/MnO2 Z-scheme heterojunction and achieved 94.3% photocatalytic degradation efficiency (0.0202 min−1) for acetaminophen in 120 min. This was found to be about 3.5 and 3.8 times higher than MnO2 and Bi2O3, respectively, in deionized water. Their studies on real water systems further revealed relatively inferior degradation efficiency in tapwater (88.7%), municipal (75.5%), hospital (63.6%) and pharmaceutical industry (55.4%) wastewater compared to that in deionized water (94.3%). The assembly of Sr@TiO2 with UiO-66-NH2 in different ratios was used to construct Sr@TiO2/UiO-66-NH2 heterostructures and achieved more than 90% conversion of acetaminophen under solar light.141 A visible-light-driven 15 wt% CeO2/I,K-co-doped C3N4 heterojunction photocatalyst removed about 98% acetaminophen from aqueous solution after 120 min of irradiation compared to pure g-C3N4 (47%) and doped IK-C3N4 (75%).142 In another study, a g-C3N4/TiO2 (weight ratio: 5%)–persulfate (PS) photocatalytic system showed almost complete photodegradation ability and stability for acetaminophen under visible-light irradiation.143 Visible-light-mediated CdO–ZnO demonstrated efficient photocatalytic performance as a heterogeneous photocatalyst in the decomposition of paracetamol in an aqueous solution.144 Radical scavenger tests established the dominance of ·OH and h+ for this photocatalytic process.

A heterojunction magnetic ternary g-C3N4/TiO2–MnFe2O4 halloysite photocatalyst showed about 79.1% removal of acetaminophen (10 ppm) within 90 min under visible light.145 The ternary photocatalyst could be easily recovered by applying an external magnetic field and reused several times without any significant reduction in its catalytic activity. The removal efficiency for acetaminophen under optimum conditions in the presence of a magnetic carbon heterojunction coupled with UV light and peroxymonosulfate was insignificantly reduced from 97.4% even after five consecutive cycles.146 Moradi et al.147 used 0.6 g L−1 of TiO2/graphene/g-C3N4 (60[thin space (1/6-em)]:[thin space (1/6-em)]10[thin space (1/6-em)]:[thin space (1/6-em)]30) Z-type photocatalyst and observed complete degradation of acetaminophen (50 mg L−1) at a pH of 9.0 in 120 min due to a synergistic effect. Their investigations also showed HO· and O2· radicals to be the dominant species in the degradation of acetaminophen.

Table 2 records the performance data of different photocatalysts on the removal of acetaminophen from wastewater.

Table 2 Performance data on removal of acetaminophen in water using variety of photocatalysts
Photocatalyst Preparative method ACT Catalyst dose pH Light source Degradation and time Rate constant
TiO2-rutile76 Precipitation 20 ppm 0.1 g (50 mL) 9 Tungsten halogen lamp (400 W), 0.0146 W cm−2 68% (60 min)
TiO2-anatase76 Thermal precipitation method 20 ppm 0.1 g (50 mL) 9 Tungsten halogen lamp (400 W), 0.0146 W cm−2 60% (60 min)
ZnO76 Thermal precipitation method 20 ppm 0.1 g (50 mL) 9 Tungsten halogen lamp (400 W), (0.0146 W cm−2) ∼100% (60 min) in 1 h
TiO2: 80% anatase + 20% rutile (Degussa P25)77 Commercial 40 mg L−1 (250 mL) 2 g L−1 UV lamp (15 W) 97% (300 min)
TiO2/Ag (5%)78 Photodeposition method 20 μg L−1 (O2: 100 cm3 min−1) 1 g L−1 UV radiation (365 nm) 94.50% (240 min)
TiO279 Sol–gel method 50 ppm (750 mL) 1.33 g L−1 TQ159-ZO lamp (150 W) ∼50% (180 min) 0.0056 min−1
TiO280 Sol–gel method 35 mg L−1 0.15 g 10 UV lamp with a wavelength of 256 nm, 1 mW cm−2 99% (180 min)
Solid TiO2 spheres81 Template-free solvothermal route 50 mg L−1 0.1 g L−1 Mercury lamp (500 W) 90% (60 min) 0.075 min−1
Mesoporous TiO2 microspheres81 Template-free solvothermal route 50 mg L−1 0.1 g L−1 Mercury lamp (500 W) 94% (60 min) 0.043 min−1
TiO2 (High Techn. Nano co. Ltd)82 Commercial 50 μM 1.0 g L−1 9 Metal halide lamp (250 W), λ ≥ 365 nm ∼95% (100 min)
ZnO powders (Fluka)84 Commercial (thermally calcined at 100 °C) 50 mg L−1 0.25 g (0.25 L) UV-lamp (315–400 nm), P.D: 0.66 mW cm−2 ∼97% (240 min) 0.0136 min−1
ZnO nanonuts86 Chemical method 5 × 10−5 M ∼1.0 mg 7.2 UV lamp: 4 mW cm−2, 368 nm ∼92% (180 min) 1.32 × 10−2 min−1
TiO2 (Degussa P25)87 Commercial 0.3 mg L−1 40.5 mg (70 mL) Neutral LED lamp – UVA light (15 W), 365 nm 100% (40 min) 0.12 min−1
Au–TiO287 Mixing tempered colloidal solution of au and TiO2 in water 0.3 mg L−1 40.5 mg (70 mL) Neutral LED lamp – UVA light (15 W), 365 nm 100% (32 min) 0.14 min−1
Au–g-C3N487 Reflex method 0.3 mg L−1 40.5 mg (70 mL) 5.9 Visible light 100% (25 min) 0.17 min−1
Ag(1 wt%)/TiO288 Sonicating mixture of TiO2 and aqueous AgNO3, stirring and irradiating with 450-W ACE lamp for 1 h 20 mg L−1 0.4 g L−1 6.3 Simulated solar light xenon lamp (1000 W), 50.0 mW cm−2 ∼98% (180 min) 0.019 min−1
Au(1 wt%)/TiO288 Sonicating mixture of TiO2 and aqueous H2AuCl6, stirring and irradiating with 450 W ACE lamp for 1 h 20 mg L−1 0.4 g L−1 6.3 Simulated solar light xenon lamp (1000 W), 50.0 mW cm−2 ∼93% (180 min) 0.016 min−1
Pt(1 wt%)//TiO288 Sonicating mixture of TiO2 and aqueous H2AuCl6, stirring and irradiating with 450 W ACE lamp for 1 h 20 mg L−1 0.4 g L−1 4.2 Simulated solar light xenon lamp (1000 W), 50.0 mW cm−2 ∼100% (180 min) 0.020 min−1
Pd/CuO89 Deposition and sputtering 10 mg L−1 (20 mL) 15 (l) × 15 (w) × 1 (t) mm film Xenon arc lamp: 150 W, λ > 420 nm ∼90% (240 min) 0.796 h−1
TiO2/BN/Pd92 Electrospinning and atomic layer deposition 1 mg L−1 (250 mL) 0.5 g L−1 6.8 Medium-pressure metal halide UV lamp (400 W) 100% (10 min) 0.019 min−1
TiO2/BN100/Pd10092 Electrospinning and atomic layer deposition 1 mg L−1 (250 mL) 0.5 g L−1 6.8 400 W halogen linear lamp (visible irradiation) 98% (180 min) 0.28 min−1
C,N-co-doped TiO293 Peroxo–gel method 4 mg L−1 20 mg UV-light (10 W), λ: 365 nm 69.31% (120 min)
C-doped TiO294 Sol–gel method 2.0 ppm 2.0 g L−1 7 Low UV lamp pressure (15 W), 365 nm, 65 W m−2 100% (90 min) 0.0817 min−1
Supported titania-based catalysts (25 wt% mg doping)95 Industrial petrochemical (source) 20 mg L−1 0.7 g L−1 (25 mL) 4.3 UV lamp: 365 nm, 30 W m−2 60% (60 min)
Mercury vapour lamp (125 W), (202 W m−2) 48.3% (60 min)
TiO296 Hydrolysis of Ti isopropoxide (sol–gel method) 35 mg L−1 0.5 g L−1 5.5 UV irradiation: HG500 lamp (30 mW cm−2) ∼84% (120 min) 12.4 ± 0.2 × 10−3 min−1
Ta-doped TiO2 (Ti/Ta molar ratio: 1%)96 Hydrolysis of Ti isopropoxide (sol–gel method) followed by Ta doping through impregnation method 35 mg L−1 0.5 g L−1 5.5 UV irradiation: HG500 lamp (30 mW cm−2) ∼70% (120 min) 9.4 ± 0.1 × 10−3 min−1
TiO296 Hydrolysis of Ti isopropoxide in presence of CH3COOH 35 mg L−1 0.5 g L−1 5.5 UV irradiation: HG500 lamp (30 mW cm−2) ∼70% (120 min) 9.3 ± 0.1 × 10−3 min−1
Ta-doped TiO2 (Ti/Ta molar ratio: 1%)96 Hydrolysis of Ti isopropoxide in presence of CH3COOH followed by ta doping through impregnation method 35 mg L−1 0.5 g L−1 5.5 UV irradiation: HG500 lamp (30 mW cm−2) ∼73% (60 min) 10.4 ± 0.1 × 103 min−1
Mesoporous MnOx–TiO297 Sol–gel method 25 ppm (150 mL) 0.1 g L−1 Continuous sonication (20 W) and UVA radiation (160 W m−2) 26% (180 min)
IL-Fe-doped TiO2 with Fe to Ti molar ratios (%): 298 Sol–gel method 10 mg L−1 (200 mL) 0.65 g L−1 7 UV lamps 90.35% (90 min) 0.25 min−1
Synthetic TiO2 doped with (KAl(SO4)2)99 Sol–gel method 0.10 mM 1.0 g L−1 6.9 Visible light: source (light emitting diodes) with λ > 440 nm 95% (540 min) 5.20 × 10−3 min−1
Carbon-self-doped TiO2101 Sol–gel method (product calcined at 300 °C) 0.1 mM (500 mL) 1.0 g L−1 6.9 LEDs (λ > 440 nm) ∼96% (540 min) 5.0 × 10−3 min−1
Bi3+(10%)-doped anatase TiO2102 Hydrolysis method 104 M (100 mL) 0.1 g L−1 5 Source: UV-vis, (4 W cm−2) ∼100% (240 min) 0.97 h−1
Ba1−xBiFe1−xCuxO3 (x = 0.05)103 Pechini method 50 mg L−1 0.75 g L−1 9 Metal halide efficacy lamp 98.1% (120 min)
Ag/ZnO104 Chemical method 5 mg L−1 (500 mL) 1 g L−1 8.5 Tungsten halogen lamp (300 W) 90.8% (120 min) 0.020 min−1
1.0 wt% La-doped ZnO105 Precipitation method 100 mg L−1 (500 mL) 0.1 g Compact fluorescent lamps: 20 W 99% (3 h)
1% Ce-doped ZnO106 Hydrothermal method 5 mg L−1 1 mg mL−1 6.8 UV-B mercury lamp (8 W) 68% (240 min) 0.0058 min−1
N-Implanted ZnO nanorod array (NRA)107 ZnO NRAs by two-step process followed by N implantation by low energy ion beam 20 ppm (5 mL) 10 × 10 mm aligned ZnO NRA Visible-light irradiation 98.46% (120 min) 0.038 min−1
TiO2/SiO2/Fe3O4111 Ultrasonic-assisted sol–gel method 30 mg L−1 (400 mL) 1.34 g L−1 7 Low-pressure mercury lamp: λ: 254 nm, 3.8 × 10−6 Ein L−1 s−1 ∼97% (300 min) 1.7 × 109 M−1 s−1
MWCNT (1.72 wt%) TiO2–SiO2112 Sol–gel method 10 mg L−1 Nearly neutral High-pressure mercury lamp, 500–550 nm, 7.31–7.53 mW m−2 81.6% (60 min) 0.0113 min−1
Magnetite–hematite113 Hydrothermal 20 mg 0.13 g L−1 Medium-pressure hg vapour lamp (400 W) ∼100% (45 min)
TiO2 (438 mg)–Bi4O5I2114 In situ calcination method 3 ppm 25 mg Xenon lamp with a light filter of 400 nm ∼95% (6 min) 0.425 min−1
Cu2O/WO3/TiO2115 Hydrothermal 1 mg L−1 (80 mL) 20 mg 9 Solar-light irradiation (source) 92.5% (60 mL) 4.42 × 10−2 min−1
Flower-like 50% TiO2/Fe2O3117 Modified ultrasonic assisted sol–gel method 50 mg L−1 (50 mL) 0.1 g L−1 Medium-pressure Hg lamp (450 W) 87.8% (90 min) 0.0219 min−1
3% WO3/TiO2/SiO2119 Solution method 10 mg L−1 1.0 g L−1 9 Xenon lamp (500 W) without cut-off filter 800 nm cut-off filter (800 nm > λ > 200 nm) 88% (240 min) 0.70 h−1
TiO2 (40 wt%) /ZSM-5120 Sol–gel method 15 mg L−1 (500 mL) 1.0 g L−1 6.8 UV lamp (14 W), 254 nm, 0.97 mW cm−2 96.6% (180 min)
1.1% ZnO/polystyrene122 Solvent casting method 12.5 mg L−1 25 g (50 mL) 6.5 UV light (13 W m−2) 77% (240 min)
Bi modified titanate124 Hydrothermal method 0.7 mg L−1 1.0 g L−1 7 Metal halogen lamp with UV and IR cut-off filters 88% (180 min) 12.61 × 10−3 min−1
BaTiO3/TiO2 ratio of 3[thin space (1/6-em)]:[thin space (1/6-em)]1 (w/w)125 Grounding followed by drying and calcination 5 mg L−1 1 g L−1 7 Xenon lamp: 500 W (200 nm < λ < 800 nm) 95% (240 min) 0.5529 h−1
Ag/AgCl@ZiF-8126 Stirring method 1 mg L−1 0.5 g L−1 5 Metal halogen lamp (500 W) combined with UV and IR cut-off wave length 99% (90 min) 0.0579 min−1
g-C3N4127 Thermal oxidation etching process 5 mg L−1 0.1 g (250 mL) Solar irradiation (source) 99% (60 min)
Exfoliated g-C3N4128 Thermal synthesis 25 g dm−3 0.9 g UVA lamp: 368 nm, 0.96 mW cm−2 41% (120 min) 4.5 × 10−3 Mol dm−3 min−1
Exfoliated g-C3N4128 Thermal synthesis 25 g dm−3 0.9 g Visible light lamp (446 nm), an intensity of 8.5 mW cm−2 54% (120 min)
0.05% ZnO/Ph–g-C3N4129 Single-step calcination and combustion process 20 mg L−1 1 g L−1 Halogen lamp (500 W) 90.8% (120 min)
α-Fe2O3/g-C3N4130 Dispersion under sonication followed by heating in air 2.0 mg L−1 (H2O2: 5.0 mM) 0.1 g L−1 5.0 Xenon lamp: 35.0 W (λ > 420 nm) 100% (25 min) 0.134 min−1
g-C3N4(75%)/UiO-66-NH2131 Hydrothermal method 5 mg L−1 (350 mL) 0.5 g L−1 4–5 9 W lamps, 365 nm 100% (120 min) 2.0 h−1
Bi2O3/MnO220 Room temperature solution phase synthesis 5 mg L−1 1 g L−1 6.8 200 W LED strip (λ > 420 nm) 94.3% (120 min) 0.0202 min−1
TiO2@rGO prepared by using 3 wt% GO133 Sol–gel method 50 mg L−1 (25 mL) 2.0 g L−1 5.4 LED lamps (18 no.) and each of l3 W, λ: 365 nm, 95 μW cm−2 100% (50 min) 0.061 min−1
Calcined ZnFe-LDH/rGO (using 30 mg of GO)135 Hydrothermal calcined method (using 30 mg GO) 5 mg L−1 (50 mL) 25 mg Xenon lamp (500 W), 300 nm cut-off filter 95% (420 min) 0.00737 min−1
5% graphene/TiO2 nanotubes136 Hydrothermal 5 mg L−1 (500 mL) 0.1 g L−1 7 UV lamp (14 W), 254 nm 96% (180 min) 00197 min−1
Coal fly ash (CFA)/GO/WO3 NRs137 Hydrothermal 5 mg L−1 100 mg 250 HW lamp 86% (180 min) −0.0116 min−1
Ni@TiO2:W138 Hydrothermal treatment immobilizing 25 mg L−1 30 mg (100 mL) 7 Solar natural irradiation (754 ± 13 W m−2) 100% (180 min) 10.7 × 10−3 min−1
Flexible graphene/Ni@TiO2:W138 TiNiW grown on the surface of graphene 25 mg L−1 30 mg (100 mL) 7 Solar natural irradiation (754 ± 13 W m−2) 86% (180 min) 8.8 × 10−3 min−1
1% rGO/BiOBr core/shell139 Hydrothermal 5 mg L−1 (30 mL) 5.5–9.5 Hg/xenon lamp (visible light irradiated with 400 nm cut-off filter), 20 mW cm−2 93% (105 min) 0.006 min−1
rGO–Ag/PANI140 Mixing reduced GO with polyaniline AgNO3 by vitamin C 25 mg L−1 50 mg 5 Visible light 99.6% (100 min)
Sr@TiO2 with UiO-66-NH2141 By carrying out growth of UiO-66-NH2 on SrTiO3 5 mg L−1 250 mg L−1 (150 mL) Xenon lamp: 600 W m−2 (λ cut-off filter: 320 nm) ∼94% (240 min) 0.67 h−1
15 wt%CeO2/IK–g-C3N4142 Mixing method 10 mg L−1 (20 mL) 2.0 g L−1 9 Visible light lamps (8 W), 465 ± 40 nm 98% (90 min) 0.0386 min−1
5% g-C3N4/TiO2/persulfate143 Ultrasonic mixing 5 mg L−1 (100 mL) and PS: 2 mM 0.331 g L−1 7 Xenon lamp (300 W) with 400 nm cut-off filter 99.3% (30 min) 0.181 min−1
CdO–ZnO (0.1[thin space (1/6-em)]:[thin space (1/6-em)]0.2 mole ratio)144 Homogeneous co-precipitation 12 ppm 1 g L−1 6.15 Halogen lamp (500 W) 96% (160 min) 0.05 min−1
Magnetic mesoporous carbon146 In situ chemical co-precipitation method 20 mg L−1, PMS: 0.6 mM 0.12 g L−1 6 UVC lamp – Philips (6 W) with 254 nm cut-off filter 97.4% (40 min)
TiO2/graphene/g-C3N4 (60[thin space (1/6-em)]:[thin space (1/6-em)]10[thin space (1/6-em)]:[thin space (1/6-em)]30)147 Hydrothermal method 50 mg L−1 0.6 g L−1 9 Xenon lamp (SSL irradiation): 300 W, λ cut-off filter: 420 nm 100% (120 min) 2.7 × 10−2 min−1


3.2 Amoxicillin

Amoxicillin (AMX) is a widely used semi-synthetic β-lactam and broad-spectrum antibiotic in the treatment of different types of infection for treating both human and animal diseases.148 Therefore, it is possible to find traces of this drug or its degradation products in various aquatic environments in the treated discharge from wastewater treatment plants. Its presence in aquatic animals and humans contributes to toxic effects though the aquatic system due to its structure, high polarity, and water solubility. However, amoxicillin in water is not easy to remove by conventional wastewater treatment processes due to its resistance to biodegradation. Hence, it is necessary to conduct a large amount of research on the treatment and removal of amoxicillin from wastewater using a variety of photocatalysts before discharging it into the natural aquatic environment.149–216
3.2.1 Metal oxides.
3.2.1.1 TiO2. Radosavljević et al.149 applied TiO2 in a nanocrystalline form and compared it with commercial TiO2 to study the photocatalytic degradation of amoxicillin using an Osram Ultra-Vitalux® lamp as the light source. Their findings indicated almost complete degradation of AMX after 210 min for catalyst and AMX concentrations of 2 g dm−3 and 100 mg dm−3, respectively. The UV-mediated photocatalytic degradation of amoxicillin was found to be low (27.6%) in the presence of TiO2 (10–25 nm) compared to cephalexin (63.5%) and tetracycline (100%) under optimal conditions.150 Pereira et al.151 used photoreactors and studied the degradation of amoxicillin in aqueous solution (pH: 7.5) by subjecting it to a solar-driven TiO2 (0.5 g L−1) assisted photocatalytic process. According to this, TiO2/solar UV radiation was able to reduce the antibiotic concentration from 40 to 3.1 mg L−1 after 4.6 kJUV of UV accumulated energy per liter of solution.

The degradation of amoxicillin (10 mg L−1) was also examined under UV and visible irradiation (15 min) and found to be nearly 100% for TiO2 and ZnO (both 0.01 g), respectively.152 Amoxicillin (104 mg L−1) in aqueous solution (pH ∼ 5) was completely degraded under TiO2/UVA (365 nm) in 30 min in the presence of H2O2 (100 mg L−1).153 TiO2-catalyzed photodegradation of amoxicillin (10 mg L−1) was found to be ∼100% under UV irradiation of 30 min duration.154 According to Klauson et al.,155 Degussa P25 TiO2 showed about 83% degradation of AMX (pH: 6.0) after 2 h under solar radiation. Moosavi and Tavakoli156 studied amoxicillin degradation in contaminated water using TiO2 in solar photocatalysis, considering variations in pH, catalyst dose and initial concentration of amoxicillin. These studies showed 84.12% degradation of amoxicillin after 240 min under optimum conditions of pH 9.5, catalyst dose of 1.5 g L−1 and initial concentration of amoxicillin of 17 mg L−1 under 240 min of solar irradiation due to a synergistic effect. In addition, several other studies have also been reported using TiO2,157–159 and supported TiO2160 on the photocatalytic remediation of amoxicillin.


3.2.1.2 ZnO and other metal oxides. The effect of operating variables has been studied on the degradation of amoxicillin (104 mg L−1) in aqueous solution driven by a UV/ZnO photocatalyst prepared by a microwave-assisted gel combustion method, which achieved complete degradation corresponding to a zinc oxide concentration of 0.5 g L−1, irradiation time of 180 min and pH 11.161 The photocatalytic reactions followed pseudo-first-order kinetics with a rate constant of 0.018 min−1. In another study, the photocatalytic removal of amoxicillin (and sulfamethoxazole) was achieved in 6 h from aqueous solutions using ZnO nanoparticles irradiated with UVC irradiation.162 Al-zobai et al.163 reported the recovery of 72.3%, 85.3%, and 100% of amoxicillin under optimum conditions using UV/TiO2, UV/ZnO/TiO2 and UV/ZnO.163 Bi2O3/Fe (3 wt%), successfully synthesized by a microwave-assisted precipitation method, exhibited a degradation efficiency of 76.34% and a degradation rate for amoxicillin of 0.0079 min−1.164

The effect of AMX concentration, WO3 dosage, and pH was studied for the photocatalytic degradation of amoxicillin by solar-driven simulated irradiation.165 These findings revealed the complete removal of AMX under optimal conditions corresponding to an initial AMX concentration of 1.0 μM, catalyst dosage of 0.104 g L−1 and pH 4. Sol–gel-synthesized nano-NiO under optimal conditions efficiently degraded 96% of amoxicillin from pharmaceutical wastewater.166 The photodegradation process was found to follow pseudo-first-order kinetics (k: 0.084 min−1) for an amoxicillin concentration of 25 mg L−1.

3.2.2 Doped metal oxides. According to Klauson et al.,155 TiO2 doped with C (32 at%) and Fe (2.2 at%) under identical conditions of solar radiation in 2 h of treatment and pH 6.0 TiO2 showed about 83%, 73% and 75% degradation of amoxicillin, respectively. Mohammadi et al.167 used Sn (1.5 mol%) doped/TiO2 nanoparticles to carry out the photocatalytic decomposition of amoxicillin trihydrate in aqueous solutions under UV light. It showed high photocatalytic activity during the mineralization of AMX due to hydroxyl radicals and band gap energy. Sol–gel-synthesized Sn,Zn-co-doped TiO2 showed marked improvement in the photocatalytic degradation of amoxicillin trihydrate due to the synergistic actions of the dopants.168 According to Wahyuni et al.,169 doping of Cu in TiO2 shifts the light absorption into the visible region. Furthermore, doping of Cu in TiO2 increased the degradation of amoxicillin under visible light. Amoxicillin (10 mg L−1) exhibited about 90% photodegradation using 0.40 g L−1 of a Cu (4.56 mg g−1) doped TiO2 photocatalyst in 24 h at pH 6 under visible-light irradiation. In another study, the removal of amoxicillin from aquatic and pharmaceutical wastewater solution was studied using Fe3+-doped TiO2 under UVA radiation.170 These findings revealed removal efficiencies of 99.14% and 88.92% under the optimum conditions (pH: 11, initial concertation of amoxicillin: 10 mg L−1, catalyst: 90 mg L−1, contact time: 120 min) for synthetic and pharmaceutical water, respectively.

A Ce3+-doped TiO2 thin film, prepared using polyethylene glycol as the templating agent, acting as a catalyst succeeded in the removal of amoxicillin under UVA radiation from aqueous solution (pH 6.0).171 It was noted that the removal of amoxicillin increased from 28% to 67% (2 h) in the presence of Ce3+@TiO2, corresponding to a decrease in the initial concentration of amoxicillin from 15.0 to 0.5 mg L−1, respectively. The Ce3+@TiO2 thin film retained its photocatalytic stability more or less unaltered even after 6 cycles. It was suggested that cerium ions trapped the electron and hole pairs in the TiO2 catalyst to form hydroxyl and peroxy radicals that play a significant role in the degradation of amoxicillin. Mn-doped Cu2O nanoparticles synthesized using aloe vera leaf extract exhibited 92% degradation of amoxicillin under sunlight irradiation at pH 9, an initial concentration of amoxicillin of 15 mg L−1, and a photocatalyst dosage of 1 g L−1.172 In all likelihood, Mn doping in Cu2O delays rapid recombination by trapping the photogenerated electrons, accounting for its enhanced photocatalytic performance in amoxicillin degradation.

3.2.3 Metal dispersed on metal oxides. The photocatalytic degradation of amoxicillin antibiotic was investigated in the presence of La and Ce nanoparticles as co-catalysts dispersed on the surface of TiO2.173 These findings showed it had more than twice the activity of pure TiO2 in the removal of amoxicillin, which was attributed to the synergistic interaction between La and Ce nanoparticles loaded on TiO2. However, more work still needs to be carried out to explore the effect of different metals on the surface of TiO2 and ZnO for the photodegradation of antibiotics. UV-visible or visible illuminated TiO2 nanowire arrays (TNAs), TiO2 nanowires (TNWs)/TNAs, Au–TNAs and Au–TNWs/TNAs degraded amoxicillin completely in aqueous solution within 20 min due to the surface plasmonic effect and synergistic effects.174 The photodegradation of amoxicillin (and levofloxacin) was performed using an Ag/ZnO photocatalyst in aqueous solution under A-type ultraviolet irradiation (UVA 365 nm) to study its variation with solution pH, initial concentration of amoxicillin, catalyst dosage, and reaction time.175 According to this, maximum removal (93.7%) of amoxicillin was achieved under optimum conditions corresponding to Ag/ZnO concentration of 0.15 g L−1, pH 5, amoxicillin concentration of 5 mg L−1 and contact time of 120 min.
3.2.4 Metal oxide nanocomposites.
3.2.4.1 TiO2 nanocomposites. Bergamonti et al.176 studied the photocatalytic activity of TiO2 immobilized on a chitosan scaffold under UV/vis irradiation to examine the degradation of amoxicillin in wastewater under UV-vis irradiation. These findings showed high photodegradation efficiency compared to the direct photolysis of amoxicillin. A TiO2/PAC (powdered activated carbon) mixture in suspension removed 95% amoxicillin in 60 min owing to significant synergy.177 TiO2/zeolite-photocatalysis also presented a feasible methodology for the degradation of the AMX under UV radiation.178 It was noted that a material obtained by acid–alkaline pretreatment and calcination (300 °C) showed the best performance due to its favorable surface structure and TiO2 content.

Pastrana-Martínez and others179 prepared nanodiamond (ND) composites of pristine TiO2 (NDDT) to study its oxidative degradation of amoxicillin soluble in water under near-UV/vis irradiation. Their findings clearly revealed the complete degradation of amoxicillin by NDDT, owing to the generation of holes and better charge separation. In addition, specific surface area, functional groups introduced in ND and the porosity of NDDT compared to bare TiO2 also play an important role in the photocatalytic degradation efficiency of amoxicillin. Li and coworkers180 investigated the effect of Fe3O4 loading in TiO2–Fe3O4 composites, H2O2 concentration, different initial pH and light intensity on the degradation of amoxicillin. The separation showed the following trend towards the degradation of amoxicillin in 100 min under optimum conditions (amoxicillin: 30 mg L−1, UV irradiation: 200 W, [H2O2]: 4.24 mM, pH: 2.84): TiO2/15 wt% Fe3O4 + H2O2 > TiO2/20 wt% Fe3O4 + H2O2 > TiO2/25 wt% Fe3O4 + H2O2 TiO2/10 wt% Fe3O4 + H2O2 > TiO2 + H2O2. It was noted that the presence of H2O2 contributed to oxidation in a photo-Fenton process while the choice of the optimum pH of 2.84 is guided by the scrambling of Fe3+ between OH and H2O2. Furthermore, the reaction rate below 200 W increased remarkably with increasing light intensity due to the generation of electrons and holes. As a consequence, maximum AMX removal efficiency (∼88% in 100 min) was achieved for 0.4 g L−1 of TiO2/15 wt% Fe3O4/H2O2 (6 mM) under optimum conditions corresponding to an initial concentration of amoxicillin of 30 mg L−1 and catalyst loading of 0.4 g L−1. The highest performance for amoxicillin in the presence of TiO2/15 wt% Fe3O4 could be ascribed to the generation of more active ·OH. The proposed mechanism involved the rapid transfer of excited electrons from TiO2 to Fe3O4, reducing h+/e pair recombination and providing an additional ·OH generation pathway for amoxicillin degradation.

dela Rosa et al.181 studied the degradation and kinetic profiles of amoxicillin using solar/TiO2/Fe2O3/persulfate and the corresponding findings are displayed in Fig. 8(A) and (B), respectively. It was observed that AMX degradation was reduced from 70% (no scavengers) to 39%, 54% and 64% (50 min) in the presence of methanol (MeOH), tert-butanol (t-BuOH) and 1,4-benzoquinone, respectively. Based on the overall findings, arrangements of reactive oxygen species (ROS) for AMX degradation by a solar/TiO2–Fe2O3/PS process follows the order: h+ > SO4· > HO· > O2·. The overall amoxicillin degradation can be accounted for by considering the suppression of recombination of charges by the presence of PS as well as the generation of ROS at h+.


image file: d3lf00142c-f8.tif
Fig. 8 (A) Photocatalytic degradation of AMX under solar irradiation in the presence of scavengers; and (B) corresponding zero-order rate constants (kobs) (experimental conditions: [AMX] = 50 μm; initial pH = 4; [PS] = 334 μm, treatment time, t = 50 min). Reproduced from ref. 181 with permission from Wiley (2021).

TiO2 immobilized on activated carbon fabricated by a high-temperature impregnation method degraded amoxicillin, diclofenac and paracetamol by 100% (120 min), 85% (180 min) and 70% (180 min) in aqueous solution under solar irradiation.182 Li et al.183 reported the photocatalytic degradation of amoxicillin using TiO2 nanoparticles submerged on a porous ceramic membrane. TiO2 immobilized on sand has been used as a catalyst in a solar photocatalytic process for the removal of amoxicillin residues from aqueous solution.184 These findings showed 93.12% degradation of amoxicillin under the optimal conditions of pH 5, 7 5 mg L−1 of TiO2, 400 mg L−1 of H2O2, and 10 mg L−1 of AMX concentration at 150 min irradiation time. Furthermore, the removal of undesirable compounds follows a pseudo-second-order kinetic model. In addition, TiO2/Mg–Al-layered double hydroxide (LDH),185 Ag-ion-exchanged zeolite/TiO2,186 Fe-8-hydroxyquinoline-7-carboxylic/TiO2 flowers187 and TiO2–SiO2188 composites have also been used to remove amoxicillin from aqueous solutions.


3.2.4.2 ZnO-based nanocomposites. Thi et al.189 observed the enhanced photocatalytic activity of ZnO–TiO2 (10%) for the ozonation and perozone degradation of amoxicillin in water under visible-light irradiation. The visible-light-driven MIL-53(Al)/ZnO hierarchical photocatalyst produced 100% removal of amoxicillin corresponding to an initial amoxicillin concentration of 10 mg L−1, solution pH 4.5 and catalyst dose of 1.0 g L−1.190 Recently, Liu and others191 reported significantly high degradation efficiency of amoxicillin (93.10%) in wastewater using Bi2WO6/nano-ZnO (1[thin space (1/6-em)]:[thin space (1/6-em)]3) after 120 min in comparison to ZnO and Bi2WO6. It is anticipated that the reduction in band gap energy of Bi2WO6/nano-ZnO (1[thin space (1/6-em)]:[thin space (1/6-em)]3) could prevent the recombination of photogenerated charge carriers.
3.2.5 Graphitic-carbon-based nanocomposites.
3.2.5.1 g-C3N4-based nanocomposites. Carbon-rich g-C3N4 nanosheet samples were prepared by a combination of 20 g of urea and 60 mg, 90 mg and 120 mg of 1,3,5-cyclohexanetriol as starting materials (referred to as C-CN60, C-CN90 and C-CN120, respectively).192 They included plenty of carbon-rich functionalities and were examined for their photocatalytic activity for amoxicillin degradation under solar and visible light in the aqueous phase and the results are displayed in Fig. 9. The degradation of amoxicillin was found to follow the order: C-CN90 > C-CN60 > C-CN120 > g-C3N4. Photocatalyst C-CN90 showed nearly complete photocatalytic degradation of amoxicillin under solar light and visible light after 150 and 300 minutes, respectively. This has been attributed to the interaction between g-C3N4 and graphited conjugated construction narrowing the band gap and separating photogenerated electron–hole pairs.
image file: d3lf00142c-f9.tif
Fig. 9 Photocatalytic degradation kinetics of AMX by the synthesized materials under (a) simulated solar light, (b) visible light, and (c) AMX degradation rate constants under solar and visible light. Reproduced from ref. 192 with permission from Elsevier (2021).

Silva et al.193 synthesized metal-free polymeric carbon nitrides using melamine (CN-M), thiourea (CN-T) and their 1[thin space (1/6-em)]:[thin space (1/6-em)]1 mixture (CN-1M[thin space (1/6-em)]:[thin space (1/6-em)]1T) as precursors in a Teflon reactor comprising 25 mL of deionized water followed by heating of the products at 550 °C for 30 min. Their investigations revealed 100% degradation of AMX for CN-T followed by CN-M (65%) and CN-1M[thin space (1/6-em)]:[thin space (1/6-em)]1T (56%) after 48 h of visible-light exposure. The superior performance of CN-T was found to be directly related to the greater number of defects present in its structure, that can help in the separation of electron–hole pairs. An Ag/g-C3N4/ZnO nanorod (0.08 g L−1) nanocomposite has also acted as an efficient photocatalyst in the photocatalytic degradation of amoxicillin of high concentration (40 mg L−1) irradiated by visible light.194 V2O5-nanodot-decorated laminar C3N4 degraded amoxicillin under solar light, exhibiting 91.3% removal efficiency.195 It is suggested that such a V2O5/C3N4 S-scheme structure provides an internal electron channel at the interface and maintains the active sites with high potentials for the photodegradation of amoxicillin. Mesoporous g-C3N4/persulfate exhibited 99% degradation of AMX under visible-light irradiation within 60 min at pH 7 due to a synergistic effect.196 Graphitic-carbon–CuO–ZnO nanocomposites exhibited 49% efficiency in the photocatalytic degradation of amoxicillin under direct sunlight and followed pseudo-first-order kinetics.197 α-Fe2O3/g-C3N4,198 mesoporous g-C3N4,199 and CQDs/K2Ti6O13200 photocatalysts have also been reported in the photocatalytic degradation of amoxicillin.


3.2.5.2 Graphene-based nanocomposites. Changotra et al.201 prepared nanocomposites of varying FeS2 to GO weight to study the degradation of amoxicillin as a function of different parameters, such as solution pH value, optimal doses of H2O2 and catalyst, stability of the catalyst, and leaching effect of the catalyst, under optimal solar-Fenton treatment. These investigations showed the complete degradation of amoxicillin (∼99%) by FeS2/GO (4[thin space (1/6-em)]:[thin space (1/6-em)]3) in 180 min owing to the synergistic coupling of FeS2 and GO under the optimal conditions of [amoxicillin]init conc 25 mg L−1, [FeS/GO] 0.75 g L−1, 12 mM [H2O2] and pH 5. Further, HO· acted as dominant reactive species and no toxic secondary products were produced in the amoxicillin degradation. The photocatalytic degradation efficiency for amoxicillin by TiO2 nanoparticles loaded on graphene oxide under UV light was found to be >99% at pH 6, catalyst dose of 0.4 g L−1, amoxicillin concentration of 50 mg L−1 and intensity of 36 W (Fig. 10(a–d)).202
image file: d3lf00142c-f10.tif
Fig. 10 The effect of different operational factors on AMX photocatalytic degradation and kinetic constant (a–d). Reproduced from ref. 202 with permission from Springer (2021).

According to Song and others,203 KBrO3 added to graphene–TiO2 nanotubes achieved 100% photodegradation of amoxicillin under UVA-light irradiation. It is suggested that KBrO3 prevents electron–hole recombination and has a direct role as an oxidant in the degradation of amoxicillin. A visible-light-driven MIL-68(In)–NH2/graphene oxide (GO) composite photocatalyst (0.6 g L−1) exhibited 93% degradation (120 min) of amoxicillin in aqueous solution of pH 5 compared to pure MIL-68(In)–NH2.204 It is suggested that MIL-68(In)–NH2/GO acted as an electron transporter for suppressing photogenerated carrier recombination and also acted as a sensitizer for enhancing visible-light absorption. The proposed mechanism suggested that h+ and ·O2 are active species. In another study, a 2D/3D g-C3N4/BiVO4 hybrid photocatalyst decorated with rGO (1.2 wt%) degraded amoxicillin by 91.9% under optimized conditions with visible-light illumination.205

3.2.6 Heterostructures, heterojunctions and Z-scheme-based photocatalysts. Thuan et al.206 compared the superior performance of an InVO4@Ag@g-C3N4 ternary heterojunction in the photocatalytic degradation of amoxicillin in an aqueous environment at an initial AMX concentration of 10 ppm and catalyst dose of 0.5 g L−1 under visible light for 420 min: InVO4@Ag@g-C3N4 (∼99%) > InVO4@Ag@g-C3N4 (∼80%) > InVO4@ (∼43%) > g-C3N4 (∼37%). The choice of Ag in this work is mainly guided by its two-fold contribution in the InVO4@Ag@g-C3N4 ternary heterojunction. It accounts for the enhanced electron–hole separation of both g-C3N4 and InVO4 components. In addition, silver also acts as an electron mediator to improve electron transfer from the InVO4 conduction band to the g-C3N4 valence band. A CuI/FePO4 p–n heterojunction nanocomposite showed photodegradation efficiency of 90% for the elimination of amoxicillin under simulated sunlight radiation.207 A mesoporous SnO2/g-C3N4 nanocomposite exhibited degradation to the extent of 92.1% against amoxicillin and 90.8% for pharmaceutical effluent in 80 min.208 Such excellent performance is ascribed to the presence of a heterojunction, effective separation, good band structure and good light absorption.

El-Fawal et al.209 observed the better performance of an AgFeO2–graphene/Cu2(BTC)3 MOF heterojunction compared to AgFeO2/graphene and AgFeO2/Cu2(BTC)3 binary photocatalysts in achieving about 97% removal of amoxicillin and diclofenac after 150 min under sunlight irradiation, which exhibited excellent stability up to four cycles. Based on these findings, a direct Z-scheme heterojunction mechanism has been proposed for the separation of photo-induced charge carriers at the interface of these photocatalysts. The enhanced photocatalytic activity of the tertiary heterojunction photocatalyst was mainly attributed to its superiority for light absorption (up to 650 nm) with high photostability, accelerated e/h+ pair separation and increased lifetime of photogenerated charges. The heterojunction p-ZnO/CuO (50[thin space (1/6-em)]:[thin space (1/6-em)]50 wt%) assisted photocatalytic process removed amoxicillin (initial concentration: 50 mg L−1) from water (pH: 11) almost completely on exposure to solar irradiation for 4 h.210 The degradation of amoxicillin followed pseudo-first-order kinetics (k: 9.95 × 10−3 min−1).

Gao et al.211 deposited Ag nanoparticles on the surface of a TiO2/mesoporous g-C3N4 heterojunction and used it in the photocatalytic removal of amoxicillin under visible light. A photocatalyst fabricated in this manner achieved higher degradation efficiency for amoxicillin than a TiO2/mesoporous-g-C3N4 heterojunction, mesoporous-C3N4, or bulk-g-C3N4. Such photoactivity of an Ag/TiO2/M–g-C3N4 catalyst has been assigned to the synergistic effect accounting for the effective transfer of electrons and inhibition of electron–hole recombination. The effectiveness of this photocatalyst was also tested for the removal of amoxicillin in real situations. A WO3/Ag3VO4 Z-scheme heterojunction with enhanced separation efficiency of electron–hole and surface area was deposited on rGO and used as a photocatalyst in the degradation of amoxicillin under irradiation by visible light.212 The amoxicillin photocatalytic degradation followed the following order on irradiating it with visible light: Ag3VO4/WO3/r-GO (∼96%) > Ag3VO4/WO3 (∼37%) > WO3 > Ag3VO4 (∼32%). It is suggested that the presence of rGO, by increasing the surface area in Ag3VO4/WO3/rGO, facilitates amoxicillin adsorption and electron transfer for charge separation of Ag3VO4/WO3.

Investigations have also been made on the photodegradation of amoxicillin via a magnetic TiO2–graphene oxide–Fe3O4 composite213 and Pd nanoparticles anchored to anatase TiO2.214 Hajipour et al.215 fabricated heterojunctions of TiO2/CuO, adopting the surface modification of TiO2 with CuO, and investigated its application in the photocatalytic degradation of amoxicillin in wastewater. It should be noted that TiO2/CuO (7.5%) showed reduced photo-activity compared to a TiO2/CuO (10%) photocatalyst, which could be attributed to the partial blockage of the active sites in the TiO2 nanoparticles, In another study, a novel nanophotocatalyst of CuO nanoparticles and ZnO nanorods anchored on thermally-exfoliated g-C3N4 nanosheets established the complete removal of amoxicillin corresponding to a catalytic dosage of 0.9 g L−1 and pH 7.0 within 120 min under simulated sunlight illumination.216 Subsequently, a double Z-scheme mechanism as well as a tentative pathway were proposed in detail.

Table 3 records the performance data of different photocatalysts on the removal of amoxicillin from wastewater.

Table 3 The performance data on removal of amoxicillin in water using variety of photocatalysts
Photocatalyst Method of preparation AMX Catalyst dose pH Light source details Degradation (time) Rate constant
TiO2 nanoparticles (US3490)150 Commercial 15 mg L−1 2 g L−1 5 UV lamp (18 W) 27.6% (15 min)
ZnO nanoparticles (US3590)150 Commercial 15 mg L−1 2 g L−1 5 UV lamp (18 W) 48.6% (15 min)
GO–Fe3O4150 Ultrasonic mixing followed by reflexing 15 mg L−1 2 g L−1 Lamp (UV): 18 W 87.1% (15 min)
TiO2 (P25 Degussa)152 Commercial 10 mg L−1 (20 mL) 0.01 g UV 100% (15 min) 4.33 × 10−1 min−1
TiO2 (P25 Degussa)152 Commercial 10 mg L−1 (20 mL) 0.01 g Visible 99% (15 min)
ZnO (Hoechst)152 Commercial 10 mg L−1 (20 mL) 0.01 g UV 98% (15 min) 3.03 × 10−1 min−1
ZnO (Hoechst)152 Commercial 10 mg L−1 (20 mL) 0.01 g Visible 99% (15 min)
TiO2 (Fluka)153 Commercial 104 mg L−1 (500 mL) 1.0 g L−1 11 UV lamp: 6 W (365 nm) ∼71% (300 min) 0.007 min−1
TiO2 (H2O2: 100 m L−1)153 Commercial 104 mg L−1 (500 mL) 1.0 g L−1 5 UV lamp: 6 W (365 nm) 100% (20 min)
TiO2 (P25 Degussa)154 Commercial 0.01 g 10 mg L−1 (20 mL) UV lamp 100% (30 min) 0.433 min−1
TiO2 (Degussa P25)155 Commercial 25 mg L−1 1 g L−1, slurry 6 Solar light (16 mW cm−2) ∼83% (120 min)
Carbon (32%) doped TiO2 (Degussa P25)155 Commercial 25 mg L−1 1 g L−1, slurry 6 Solar light (16 mW cm−2) ∼73% (120 min)
Fe (2.2%) doped TiO2 (Degussa P25)155 Commercial 25 mg L−1 1 g L−1, slurry 6 Solar light (16 mW cm−2) ∼75% (120 min)
TiO2 (sigma Aldrich)156 Commercial 1.5 g L−1 17 mg L−1 9.5 Solar irradiation 84.12% (240 min)
ZnO162 Microwave assisted gel combustion method 10 mg L−1 (200 mL) 0.25 g L−1 10 UVC lamp (30 W) 100% (5 h) 0.014 min−1
WO3 (sigma Aldrich)165 Commercial 1.0 μM 0.104 g L−1 4 Xenon lamp (300 W) 99.99% (180 min) 2.908 × 10−2 min−1
NiO166 Sol–gel method 25 mg L−1 0.2 g L−1 Low mercury lamp (15 W) ∼96% (120 min) 0.084 min−1
Cu (4.54 mg g−1) doped TiO2169 Photoreduction method 10 mg L−1 40 mg 6 Wolfram lamp as visible light source ∼90% (24 h) 4 × 10−4 min−1
Fe3+ doped TiO2170 Sol–gel method 10 mg L−1 90 mg L−1 11 UV lamp of C type, 125 W, 247 nm Synthetic water: 99.14% (120 min), pharmaceutical water: 88.92% (120 min)
Mn-doped Cu2O172 Green synthesis 15 mg L−1 (100 mL) 1 g L−1 9 Sunlight irradiation (900 W m−2) 92% (180 min) 0.073 min−1
La–Ce (1 wt%) TiO2173 Sonochemical-assisted synthesis 10 mg L−1 (100 mL) Appropriate amount Halogen lamp (500 W) 75.7% (?)
Ag/ZnO175 Conventional method 5 mg L−1 0.15 g L−1 5 UVA, 365 nm 93.76% (120 min) 0.073 min−1
TiO2/chitosan176 3D printing 0.1 mM (40 mL) 15 layers (AMX/TiO2 molar ratio: 1/100) 6.7 Medium-pressure Hg vapour water jacket lamp (UV-vis), 125 W, 300–800 nm, 3.5 mW cm−2 ∼95% (2 h) 0.57 × 10−2 min−1
TiO2/PAC177 Suspension method 15 mg L−1 TiO2: 1 g L−1, PAC: 0.1 g L−1 6.5 UV-vis (540 W m−2) 90–97% (60 min) 0.034 min−1
TiO2/zeolite178 Modified reported method 30 mg L−1 (100 mL) 2 g L−1 4.05 Medium-pressure Hg lamp (47 W) with λ ≤ 290 nm cut-off 88% (240 min)
Functionalized nanodiamond-TiO2179 Liquid phase deposition 0.1 mM (7.5 mL) 1 g L−1 Medium-pressure hg vapor lamp 100% (60 min) 83.3 × 10−3 min−1
TiO2-15 wt% Fe3O4180 Hydrothermal 30 mg L−1, (H2O2: 24 mM) 0.4 g L−1 2.84 Low-pressure mercury vapor lamp: 100 W, 1200 mW cm−2 ∼88% (100 min)
TiO2@α-Fe2O3 film (PS: 334 μm)181 Spin coating 50 μm 4 Xenon lamp (450 W) 70% (50 min) 7.4 × 10−7 M min−1
TiO2 immobilized on activated carbon182 High-temperature impregnation method 50 mg L−1 (4 L) 1.2 g L−1 10 Solar irradiation 100% (120 min) 0.037 min−1
TiO2–sand184 Sol–gel dip-coating 10 mg L−1, H2O2, 400 mg L−1 75 mg L−1 5 Solar irradiation 93.12% (150 min) 0.0175 min−1
TiO2/Mg–Fe-LDH185 Direct co-precipitation method 30 mg L−1 2 g L−1 11 UVA light (λmax: 365 nm) ∼100% (240 min)
TiO2/Mg–Al-LDH185 Direct co-precipitation method 30 mg L−1 2 g L−1 5.5 UVA light (λmax: 365 nm) ∼95% (240 min)
Ag/zeolite/TiO2186 Liquid ion-exchange method One g L−1 (15 mL) 0.01 g 6.7 High-pressure Hg lamp (400 W), 120 mW cm−2 ∼25% (75 min)
TiO2(80%)–SiO2(20%)188 Sol–gel method 20 mg L−1 (100 mL) 4 g L−1 5 Hg lamp – UVA (15 W), 365 nm 88% (150 min) 0.0014 min−1
MIL-53 (Al)/ZnO190 Hydrothermal/chemical conditions followed 10 mg L−1 1.0 g L−1 4.5 Metal halide lamp: 400 W, 510 nm 100% (60 min)
g-C3N4193 Heating of aq. Thiourea in Teflon reactor 30 mg 50 mg L−1 (10 mL) pH ∼ 6 Visible light: 150 W, 16 mW cm−2 100% (48 h) 0.088 h−1
Ag/g-C3N4/ZnO nanorods194 Dispersion method 40 mg L−1 0.08 g L−1 (60 mL) Solar simulator lamp: 300 W (λ ≥ 420 nm) 41.36% (180 min) 0.01017 min−1
V2O5/C3N4195 Heating powdered NH4VO3/g-C3N4 mixture 20 mg L−1 0.5 g L−1 7 Simulated sunlight ∼91% (120 min) 0.0268 min−1
α-Fe2O3 (5%)/g-C3N4198 Solution method 20 mg L−1 0.02 g (60 mL) Neutral Solar simulator (300 W) with cut-off filter (λ > 420 nm) 46% (180 min) 40.20 × 10−4 min−1
Mesoporous g-C3N4199 Template-free method 2 mg L−1 100 g L−1 (100 mL) 9 Xenon lamp: 300 W (λ > 420 nm) 90% (60 min) 0.036 min−1
CQDs modified K2Ti6O13 nanotubes200 Hydrothermal method combined with calcination 1 mg L−1 (50 mL) 0.2 g L−1 6 Light-emitting diode, 10 mW cm−2, 365 nm 100% (90 min) 0.0495 min−1
GO/TiO2202 Chemical hydrothermal method 50 mg L−1 (100 mL) 0.4 g L−1 6 UV light (36 W) 99.84% (60 min) 0.105 min−1
Graphene@TiO2 nanotube/KBrO3 (0.20 g L−1)203 Reaction under autoclave 5 mg L−1 Light: UVA lamp: 19 W, λ = 369 nm 96.94% (180 min) 0.0186 min−1
MIL-68(In)–NH2/GrO204 Dispersion method 20 ppm (200 mL) 0.6 g L−1 5 Xenon lamp (300 W) with 420 nm cut-off filter 93% (120 min) 0.0187 min−1
1.2 wt% rGO@g-C3N4/BiVO4205 Wet impregnation method 10 mg L−1 (100 mL) 0.1 g (100 mL) Halogen lamp (500 W) 91.9% (180 min) 0.0023 min−1
InVO4/Ag/g-C3N4206 Hydrothermal 10 ppm 0.5 g L−1 Visible light (30 W bulb) >99% (420 min)
CuI/FePO4207 Reflux-assisted co-precipitation technique 20 mg L−1 (50 mL) 50 mg Visible light (400 W) 90% (120 min)
Mesoporous SnO2/g-C3N4208 Green modified technique 10 ppm (40 mL) 10 mg Xenon lamp: 300 W with a cut-off filter (λ > 400 nm) 92.1% (80 min)
AgFeO2–graphene/Cu2(BTC)3 MOF209 In situ solvothermal impregnation 5 mg L−1 5 g L−1 (50 mL) 8 Halogen lamp 500 W, 420–600 nm 97% (150 min) (6.4–8.7) × 10−2 min−1
p-CuO/n-ZnO (50:50 wt%)210 Chemical route 50 mg L−1 0.5 g L−1 11 Sunlight (109 mW cm−2) >87% (240 min) 9.95 × 10−3 min−1
1.94 wt% Ag/TiO2/mesoporous g-C3N4211 Photodeposition means 5 ppm (0.1 L) 0.1 g Xe lamp: 300 W (λ > 420 nm) 99% (60 min) 0.0614 min−1
WO3/Ag3VO4/rGO212 Multiple steps 20 ppm 0.5 g L−1 LED lamp (220 V, 30 W) ∼96% (420 min)
CuO and ZnO co-anchored on g-C3N4216 Via isoelectric point-mediated annealing 60 mg L−1 0.9 g L−1 7.0 Xenon lamp (250 W) simulated sunlight 100% (120 min) 0.0269 min−1


3.3 Sulfamethoxazole

Sulfamethoxazole is used to treat a wide variety of bacterial infections, including those of the urinary, respiratory, and gastrointestinal tracts.217 However, it has been frequently detected in wastewater and surface water in aquatic environments due to its extensive consumption, excretion and disposal. Therefore, several investigations have been made by many researchers focusing on the biodegradation of sulfamethoxazole during wastewater treatment following photocatalytic degradation of sulfamethoxazole in water using a variety of photocatalysts.218–291
3.3.1 Metal oxides.
3.3.1.1 TiO2. The photodegradation of sulfonamides has been studied in the UV/TiO2 system to study the effects of pH and salinity on sulfamethoxazole concentration and total organic carbon (TOC) during the removal of sulfonamides in a UV/TiO2 system.219 The photodegradation and mineralization rates of sulfonamides in the UV/TiO2 system satisfied pseudo-first-order kinetics. A TiO2 suspension has been used as a catalyst in a sunset solar simulator to examine the degradation of sulfamethoxazole in real municipal wastewater treatment plant effluent.220 It was inferred that hydrogen peroxide can be highly recommended for working with TiO2 at low concentrations. The photocatalytic degradation of sulfamethoxazole in surface and drinking water in the absence and presence of UV (265 nm) involving TiO2 nanoparticles after 60 minutes follow the order: UV (∼100%) > anatase TiO2 (∼92%) > rutile and commercial TiO2 (∼90%).221 The effects of different UV-LED (UVA, UVB, and UVC) wavelengths were studied in carrying out the photocatalytic decomposition of sulfamethoxazole by TiO2.222 These findings showed complete decomposition within 1 h by TiO2/UVC under the conditions of TiO2: 0.5 g L−1, natural pH, and initial concentration of sulfamethoxazole: 20 mg L−1. Sulfamethoxazole in an aqueous suspension of TiO2 (0.5 g L−1) showed 82% degradation of sulfamethoxazole under UV irradiation.223 In another study, the removal efficiency for the photocatalytic degradation of sulfamethoxazole (20 mg L−1) in aqueous solution (pH: 3) by TiO2 (0.08 g L−1) as a photocatalyst was found to be 96.5% in 60 min under UV light.224 In addition, investigations have also been reported on the degradation of sulfamethoxazole using TiO2,225–227 biochar-supported TiO2228 and immobilized TiO2229–231 as photocatalysts.
3.3.1.2 ZnO. ZnO nanoparticles prepared by a microwave-assisted gel combustion synthesis method showed complete removal of amoxicillin (and sulfamethoxazole) from contaminated water in six hours under UVC irradiation.162 It was inferred that the photocatalytic removal followed the Langmuir–Hinshelwood model in the range of concentration of 5–20 mg L−1. Mirzaei et al.232 achieved ∼97% removal of sulfamethoxazole by a zinc oxide photocatalyst in the presence of fluoride ions (F–ZnO) after 30 min of reaction illuminated by UV irradiation under optimum conditions and followed pseudo-first-order kinetics (k: 0.099 min−1). The hydrothermally synthesized ZnO at 200 °C for 8 h at pH 7.5 reached 84% removal of sulfamethoxazole after 60 min under UVA irradiation.233 In addition, TiO2 and WO3 nanoparticles have also been utilized in the removal of sulfamethoxazole by its photocatalytic degradation.234
3.3.2 Metal-modified metal oxide and mixed metal oxides. Tiwari et al.235 studied the removal of sulfamethoxazole aqueous solutions by means of Ag0(NP)/TiO2 thin film irradiated under UVA light (λmax: 330 nm) for 2 h by varying the solution pH (4.0–8.0) with an initial sulfamethoxazole concentration of 1.0 mg L−1. A decreasing trend in the removal (%) of sulfamethoxazole was noted from 59% to 50% with a variation in pH from 4 to 10. The percentage removal of sulfamethoxazole as a function of pollutant concentration of sulfamethoxazole (0.5 to 15.0 mg L−1) at constant pH of 6.0 under 2 h of UVA light showed a decreasing trend in the degradation of sulfamethoxazole from 57% to 20% with the sulfamethoxazole concentration increasing from 0.5 mg L−1 to 15.0 mg L−1. Borowska et al.236 investigated the solar photocatalytic degradation of sulfamethoxazole as a contaminant in water by Pt- and Pd-modified TiO2. Their findings established significantly enhanced absorption properties from surface modification achieved by 1%Pd/TiO2 and 1%Pt/TiO2. As a result, higher removal of sulfamethoxazole is observed compared to unmodified TiO2 in aqueous solution corresponding to a concentration of catalyst of ∼50 mg L−1 and a concentration of sulfamethoxazole of 1 mg L−1. This could be explained on the basis of their band gaps (1%Pd/TiO2: 2.92 eV, 1%Pt/TiO2: 3.18 eV).

TiO2 nanotube arrays (TNAs), TiO2 nanowires on nanotube arrays (TNWs/TNAs), Au-nanoparticle-decorated TNAs, and TNWs/TNAs efficiently degraded sulfamethazine amoxicillin, ampicillin, doxycycline, oxytetracycline, lincomycin, vancomycin and sulfamethoxazole irradiated in water under UV-vis and visible light.174 Among these, the Au–TNWs/TNAs photocatalyst showed the highest activity towards the degradation of all the antibiotics due to synergistic and surface plasmonic effects. In another study, Cu–TiO2 (at low mass ratios of 0.016–0.063 wt%) produced nearly complete degradation of sulfamethoxazole by visible light at pH 5.2 for a 4 mg L−1 initial concentration of sulfamethoxazole.237 Further studies revealed the highly stable photoactivity of Cu–TiO2, as evident from experiments comprising at least 4 cycles. Au, Ag, Cu, Au–Ag and Au–Cu nanoparticles deposited on TiO2 showed increased photocatalytic activity for the photocatalytic degradation of sulfamethoxazole using UVC light.238

3.3.3 Doped metal oxides. Tsiampalis et al.239 used iron-doped TiO2 (iron/titania ratios: 0–2%) as a photocatalyst to study the photocatalytic degradation of sulfamethoxazole under simulated solar radiation. These findings showed the highest photocatalytic efficiency (95%) for sulfamethoxazole in ultra-pure water with SMX concentration of 234 μg L−1, catalyst loading of 1 g L−1 and natural pH. The initial activity of the photocatalyst also retained half of its initial value after 5 consecutive experiments. F,Pd-co-doped TiO2 nanocomposites prepared by a microwave-assisted hydrothermal synthesis method under direct sunlight irradiation degraded ∼94.4% and 98.8% of sulfamethoxazole at 20 and 70 min, respectively.240 It was suggested that doping of TiO2 by F and Pd involved multiple processes.

F,Pt-co-doped photocatalysts have also been employed in photocatalytic degradation using direct solar light.241 Fluoride ions and Pt in the TiO2 lattice were chosen in order to control the growth of the photocatalytically active anatase phase and to introduce new energy levels between the valence and conductive bands of TiO2 to narrow its band gap. These findings demonstrated degradation of sulfamethoxazole under direct solar light and a solar simulator corresponding to about >93% (90 min) and 58% (360 min), respectively. An iodine (I)–potassium (K)–C3N4 photocatalyst removed nearly 100% of sulfamethoxazole within 45 min under visible-light irradiation.242 N,Cu-co-doped TiO2 decorated on SWCNTs demonstrated total removal of sulfamethoxazole under a pH of 6.0, catalyst dosage of 0.8 g L−1, light intensity of 200 W, US power of 200 W, and initial sulfamethoxazole concentration of 60 mg L−1 in 60 min.243

Ag metal has been used as a co-dopant in P-doped g-C3N4 in order to overcome its poor photocatalytic performance.244 The investigations of Ag (nano)–P-co-doped@g-C3N4 (Ag–P@UCN) as a photocatalyst in visible light followed the trend in the removal of sulfamethoxazole in water: Ag(nano)–P@g-C3N4 (>99%) > P-doped g-C3N4 (68%) > g-C3N4 (47%). The presence of silver nanoparticles Ag(nano)–P@g-C3N4 enhanced light absorption and also acted as photogenerated electron traps, thereby enabling the effective separation of electron and hole pairs. A mechanism has also been proposed for the degradation of sulfamethoxazole in presence of an Ag–P@UCN photocatalyst. In another study, multi-homojunction gradient-nitrogen-doped TiO2 exhibited enhanced performance in the removal of sulfamethoxazole from water compared to pristine TiO2 and non-gradient-doped TiO2 under simulated solar-light irradiation.245 Zammit et al.246 examined the removal of sulfamethoxazole using a cerium-doped zinc oxide (Ce–ZnO) photocatalyst and its comparison with ZnO and benchmark TiO2–P25 in immobilized form on a metallic support and found Ce–ZnO to be most effective under UVA irradiation. In another study,247 Zn (10 wt%)–TiO2/pBC (pretreated biochar) was investigated for the photodegradation of sulfamethoxazole under visible-light irradiation and a comparison with TiO2/pBC and TiO2 after 3 h took the following order: Zn–TiO2/pBC (80.81%) > TiO2/pBC (59.05%) > TiO2 (50.07%).

3.3.4 Metal oxide–metal oxide based composites. Fernández et al.248 focused on Fe3O4/ZnO nanocomposites on the photodegradation performance for sulfamethoxazole, trimethoprim, erythromycin and roxithromycin from surface water under UVA irradiation. Their studies showed complete removal of the antibiotics (100 mg L−1) after 70 min under optimal conditions of pH 7, [H2O2] 100 mg L−1 and catalyst dose of 100 μg L−1. In addition, a reusability evaluation of Fe3O4/ZnO after removing it by applying an external magnetic field showed no significant decrease in its performance even after 8 cycles. Investigations were also made on the solar photocatalytic removal of sulfamethoxazole and other micropollutants (carbamazepine, flumequine, ibuprofen) using TiO2 and its comparison with TiO2/Fe3O4 applied in a heterogeneous photo-Fenton process.249 Magnetically separable Fe2O3/WO3 nanocomposites were also successfully used as a peroxymonosulfate activator to efficiently degrade sulfamethoxazole under visible-light irradiation.250 Wang and others251 reported that photogenerated holes played an important role in achieving more than 99% photocatalytic degradation efficiency for sulfamethoxazole (initial solution pH: 3) in 30 min by irradiating a Bi2O3–TiO2/PAC (powdered activated carbon) ternary composite with solar light.

A composite comprising titania nanoparticles/activated carbon prepared by calcination at 400 °C exhibited much better performance in the removal of sulfamethoxazole from deionized water and seawater.252 Clay–TiO2 nanocomposites prepared via biomass-assisted synthesis showed fast degradation of sulfamethoxazole (>90%) in 30 min under sunlight.253 An LDH–TiO2 (10%) nanocomposite has been developed, keeping in view its possible reusability and regeneration after subjection to UVA radiation, to carry out the degradation of sulfamethoxazole.254 These findings established almost complete degradation after 360 min of UVA irradiation, corresponding to initial sulfamethoxazole concentration of 20 mg L−1, pH 10 and LDH–TiO2 catalyst loading of 50 mg. Recycling and reusability studies were also conducted by dissolving a mass of 50 mg of LDH–TiO2 in sulfamethoxazole (concentration: 20 mg L−1) and pH 10, irradiated for 8 h under UVA. Further investigations revealed no significant variation in sulfamethoxazole degradation efficiency from the first cycle (100%) to the fifth cycle (90.5%).

According to Długosz et al.,255 a floating TiO2-expanded perlite (referred to as EP-TiO2-773: where 773 is the calcination temperature in °C) photocatalyst enhanced the photodegradation of sulfamethoxazole in the aqueous medium over a wide range of pH values on irradiation from the near-UV spectral region. However, the fastest decrease in the concentration of sulfamethoxazole was observed for the system irradiated at pH 10. The degradation of sulfamethoxazole followed pseudo-first-order kinetics in accordance with the Langmuir–Hinshelwood model. Their findings also suggested the key role of hydroxyl radical formation in the degradation of sulfamethoxazole. Noroozi et al.256 synthesized copper doped TiO2 decorated with carbon quantum dots (CQDs) and observed its excellent performance in the degradation of SMX during 60 minute time under optimum conditions corresponding to initial SMX concentration, catalyst dosage, pH, visible light intensity and CQDs ratio in the composites of 20 mg L−1, 0.8 g L−1, 6, 75 Wm−2 and 4 wt% respectively. The photocatalytic degradation of sulfamethoxazole was found to be guided by a pseudo-first-order kinetic model with HO· and O2· as active species. Poly(ethylene terephthalate)–TiO2,257 BiVO4/SrTiO3,258 CuOx–BiVO4259 and TiO2@CuCo2O4260 were also used for the photocatalytic degradation of sulfamethoxazole.

3.3.5 Graphitic-materials-based composites.
3.3.5.1 MWCNT-based composites. WO3–MWCNT composites with different amounts of functionalized MWCNTs were prepared by a hydrothermal method (named WT-2, WT-4 and WT-8), and SMX degradation was studied under visible-light irradiation.261Fig. 11(a) shows the highest efficiency of 73.3% within 3 h for WT-8; however, WT-4 with efficiency of 65.2% was preferred due to its better dispersion in water. Further studies on SMX (10 mg L−1) degradation at different catalyst dosages of WT-4 in Fig. 11(b) showed its maximum efficiency (88.5%) corresponding to a loading of 2.00 g L−1. A possible degradation mechanism highlighting the role of O2 and OH· radicals during the photocatalytic process has also been proposed and is displayed in Fig. 11(c). Awfa et al.262 reported ∼60% photodegradation of sulfamethoxazole by magnetic carbon nanotube–TiO2 composites. Martini et al.263 observed almost complete reduction of toxicity using photocatalytic ozonation with H2O2 and Fe/CNT.
image file: d3lf00142c-f11.tif
Fig. 11 (a) SMX degradation under visible-light irradiation by WO3, WT-2, WT-4 and WT-8. Conditions: catalyst: 0.50 g L−1, SM: 10 mg L−1. (b) SMX degradation by WT-4 at different catalyst dosage (0.25, 0.50, 1.00 and 2.00 g L−1). Conditions: SMX: 10 mg L−1. (c) Schematic illustration of the proposed mechanism for the enhanced degradation of SMX by WO3-CNT composites under visible-light irradiation. Reproduced from ref. 261 with permission from Elsevier (2018).

3.3.5.2 g-C3N4-based composites. An Ag (5%)/P–g-C3N4 composite synthesized by thermal polymerization combined with a photodeposition method completely degraded sulfamethoxazole within 20 min under visible-light irradiation.264 This is attributed to the formation of holes and superoxide radicals acting as dominant active species. In addition, the surface plasmon resonance effect (Ag) and the formation of a Schottky barrier on the Ag/P–g-C3N4 interface could facilitate the enhanced generation of electrons/holes as well as accounting for the recombination of photogenerated electron–hole pairs. A magnetic ZnO@g-C3N4 composite under optimum conditions removed 90.4% of sulfamethoxazole after 60 min.265 In addition, core–shell g-C3N4@ZnO,266 peroxymonosulfate (PMS)/g-C3N4267 and Ag/g-C3N4268 have also been reported in the photocatalytic degradation of sulfamethoxazole.
3.3.5.3 Graphene-based composites. Visible-light-derived rGO–WO3 composites showed 98% removal of sulfamethoxazole within 3 hours.269 In another study, Ag@Ag2O–graphene nanocomposites comprising variable graphene concentrations (1.7, 2.5, and 3.4 wt%) were prepared to study the degradation of sulfamethoxazole under simulated solar light (λ > 280 nm) and visible-light irradiation (λ > 400 nm), including the stability of the photocatalyst and the mechanism of photocatalytic degradation.270 These findings indicated higher activity and comparable stability for the first and second cycles in an Ag@Ag2O–graphene photocatalyst loaded with 2.5 wt% graphene. Possible charge transfer processes were suggested to take place under visible-light irradiation, and holes were major active species for Ag@Ag2O–graphene photocatalytic degradation while Ag0 acted as an electron capture center. Lin et al.271 observed 92% degradation of sulfamethoxazole after subjecting an immobilized TiO2–reduced graphene oxide (rGO) nanocomposite on optical fibers to 180 min of UV irradiation. A visible-light-driven Cu2O/rGO photocatalyst successfully degraded sulfamethoxazole.272

Nawaz et al.273 used graphene oxide and titanium dioxide in combination with sodium alginate to synthesize a reduced graphene oxide–TiO2/sodium alginate (rGOT/SA) aerogel. They observed more than 99% removal of these contaminants taking place within 45–90 min under UVA light, corresponding to an optimal mass ratio of TiO2 nanoparticles with respect to graphene oxide of 2[thin space (1/6-em)]:[thin space (1/6-em)]1 in an rGOT/sodium alginate aerogel in the presence of 1 wt% sodium alginate solution. Zhou et al.274 investigated the photocatalytic decomposition of SMX by Ag3PO4, Ag3PO4–graphene and Ag/Ag3PO4–graphene under simulated solar-light irradiation. They observed that the photocatalytic activities of Ag3PO4–graphene and Ag/Ag3PO4–graphene were no better than pure Ag3PO4. However, these studies indicated the enhanced structural stability of Ag/Ag3PO4–graphene, which would be more practical in real treatment processes.

3.3.6 Heterojunction and Z-scheme-based photocatalysts. WO3–g-C3N4 (WCN) photocatalysts with different g-C3N4 amounts (referred to as WCN-4, WCN-6 and WCN-8) were prepared by a hydrothermal method and evaluated for SMX degradation under visible light.275 In view of this, Fig. 12(a) and (b) show the degradation of SMX by (a) WCN-8 at various pH and (b) WCN-8 at different catalyst dosages under visible light. The optimized WO3–g-C3N4 composite (dosage: 1.0 g L−1) showed 91.7% removal efficiency for SMX as a result of Z-scheme heterojunctions between g-C3N4 and WO3 to account for the separation between photogenerated electron–hole pairs. Alternatively, the role of the larger surface area and better visible-light absorption capability of the photocatalyst in enhancing the removal efficiency of SMX cannot be ruled out. Fig. 12(c) is a schematic illustration of the SMX photodegradation process over WCN composites under visible-light irradiation. Rodrigues et al.276 observed 97% (120 min) photocatalytic efficiency for sulfamethoxazole using Ce0.8Gd0.2O2−δ/TiO2 under UV light.
image file: d3lf00142c-f12.tif
Fig. 12 (a) Degradation of SMX by WCN-8 at various pH values under visible light: Conditions: catalyst = 0.5 g L−1, SMX = 10 mg L−1. (b) Degradation of SMX by WCN-8 at different catalyst dosages under visible light: Conditions: SMX: 10 mg L−1, no pH adjustment. (c) Schematic illustration of SMX photodegradation process over WCN composites under visible-light irradiation. Reproduced from ref. 275 with permission from RSC (2017).

In another study, Ag2S/Bi2S3/g-C3N4 heterojunctions exhibited 97.4% degradation of sulfamethoxazole in 90 min in aqueous solution under visible light.277 The stable hierarchical Fe2O3/Co3O4 heterojunction on nickel foam exhibited enhanced photocatalytic degradation of sulfamethoxazole.278 The photocatalyst was also studied to evaluate its effectiveness in surface water, hospital wastewater, and wastewater treatment. A magnetic quaternary BiOCl/g-C3N4/Cu2O/Fe3O4 nano-heterojunction exhibited 99.5% photodegradation of sulfamethoxazole (100 μM) in 60 and 120 min under visible and natural sunlight, respectively.279 Photocatalysts comprising graphene-supported p–n heterojunction rGO@Cu2O/BiVO4 composites with different Cu2O loadings (l, 5, 10, 15 and 20 wt%) were prepared to study their photocatalytic degradation activity for sulfamethoxazole oxidation under LED light at neutral pH.280 All the composites were found to be effective in sulfamethoxazole oxidation owing to the electrical conductivity of rGO and the p–n heterojunction between Cu2O and BiVO4.

Zhang et al.281 evaluated the performance of a Bi2WO6/TiO2 heterojunction for photocatalytic ozonation degradation of sulfamethoxazole under simulated sunlight. They attained 97.1% removal rate of sulfamethoxazole corresponding to a catalyst dosage of 0.2 g L−1, ozone concentration of 1.5 mg L−1, sulfamethoxazole concentration of 10 mg L−1 and pH 5.25. These studies also established excellent recyclability and stability, as evidenced through 5 cycle experiments. They also proposed a new Z-scheme transfer pathway for electrons and a degradation mechanism. A direct Z-scheme MIL-53(Co/Fe)/10 wt% MoS2 heterojunction composite photocatalyst displayed 99% removal of sulfamethoxazole (10 mg L−1) in aqueous solution (pH: 6) following visible-light-driven activation of peroxymonosulfate (initial concentration: solution 0.2 g L−1).282 Bi2O3/C3N4/TiO2@C quaternary hybrids (fabricated by a hydrothermal and calcination two-step method) exhibited high photocatalytic activity, degrading 100% sulfamethoxazole (SMZ, 5 mg L−1) within 100 min under visible-light irradiation.283 These investigations further revealed the photocatalytic degradation rates of SMZ by a Bi2O3/C3N4/TiO2@C junction to be 5.12, 2.87, and 1.35 times higher than those with Bi2O3/C3N4, C3N4/TiO2@C, and Bi2O3/TiO2@C junctions, respectively.

Ren et al.284 examined Ag (0.5, 1 and 2 wt%) nanoparticles/g-C3N4/Bi3TaO7 as Z-scheme photocatalysts prepared by combining hydrothermal and photodeposition for visible-light-driven performance in the degradation of sulfamethoxazole. It should be noted that the removal efficiency for sulfamethoxazole by Ag (1 wt%)/g-C3N4/Bi3TaO7 was found to be about 98% after 25 min and adopted the following order compared to g-C3N4, Bi3TaO7, g-C3N4–Bi3TaO7 and other Ag/g-C3N4/Bi3TaO7 composites: Ag (1 wt%)/g-C3N4/Bi3TaO7 > Ag (2 wt%)/g-C3N4/Bi3TaO7 > Ag (0.5 wt%)/g-C3N4/Bi3TaO7 > g-C3N4/Bi3TaO7 > g-C3N4 > Bi3TaO7. Such improved performance of Ag (1 wt%)/g-C3N4/Bi3TaO7 is attributed to the effective separation/transfer of photo-excited electrons and holes. In another study, an in situ prepared Ag3PO4/Bi4Ti3O12-20% heterojunction composite photocatalyst under visible-light irradiation exhibited much better photocatalytic activity in degrading sulfamethoxazole and stability compared to Ag3PO4 or pure Bi4Ti3O12.285 This is attributed to the formation of a direct Z-scheme improving the stability and activity of the Ag3PO4/Bi4Ti3O12 composite.

An Ag2O–KNbO3 (0.15Ag–Nb) composite fabricated by an in situ deposition method exhibited improved degradation of sulfamethoxazole under visible-light irradiation compared to the corresponding pure KNbO3 and Ag2O.286 The apparent rate constant of the composite was found to be 0.40 and 8 times those of KNbO3 and Ag2O, respectively. According to these studies, a type-I heterojunction formed between KNbO3 and Ag2O significantly enhanced the separation of photo-induced holes and electrons and accounted for sulfamethoxazole degradation. The rate constant value of the visible-light-driven optimal 0D/1D AgI/MoO3 (0.13 min−1) Z-scheme heterojunction photocatalyst in sulfamethoxazole degradation was found to be ∼22.4 times and 32.5 times those of MoO3 (0.0058 min−1) and AgI (0.0040 min−1), respectively.287 In addition, Z-scheme Ag3PO4/g-C3N4,288 Fe3O4–ZnO@g-C3N4,289 CeO2/g-C3N4 (CeO2: 5% mass ratio)290 and S-scheme-based N–SrTiO3/NH4V4O10291 photocatalysts have also been evaluated for the removal of sulfamethoxazole from water.

Table 4 records the performance data of different photocatalysts on the removal of sulfamethoxazole in wastewater.

Table 4 Performance data on removal of sulfamethoxazole in water using variety of photocatalysts
Photocatalyst Preparative method SMX Catalyst dose pH Light source and other details Degradation/removal (time) Rate constant
TiO2: mainly of anatase (80%), (P25 Degussa)219 Commercial 20 mg L−1 1 g L−1 5 Xenon lamp: 400 W (200 nm < λ < 700 nm) 96% (180 min) 0.026 min−1
TiO2, P-25 Degussa222 Commercial 20 mg L−1 0.5 g L−1 Natural UV lamp equipped with UV C (260 nm) 100% (180 min)
TiO2 Degussa P25223 Commercial 100 mgL−1 1.0 g L−1 5 Xenon lamp (1000 W) with λcut-off < 290 nm 88% (360 min) 0.054 min−1
TiO2 Merck224 Commercial 20 mg L−1 0.08 g L−1 3 Low-pressure mercury vapour lamp (15 W) 96.5% (60 min)
Biochar supported TiO2228 Sol–gel method 10 mg L−1 (0.1 L) 0.5 g 4 UV lamp-UVC (15 W), λ: 254 nm 91% (6 h)
TiO2 immobilized on glass spheres229 By dip coating on glass 100 μg L−1 0.335 g L−1 7.82 Solar UV radiation (λ < 400 nm) 100% (120 min) 0.030 min−1 (first cycle)
F–ZnO232 Commercial 1 mM (NH4F: 2.505 mM) 1.48 g L−1 4.7 UVC lamp: 10 W 97% (30 min) 0.099 min−1
ZnO233 Hydrothermal 10 mg L−1 200 mg L−1 7.5 UVA lamp 84% (60 min) 0.030 min−1
TiO2 nanoparticles (sigma-Aldrich)234 Commercial 50 mg L−1 500 mg L−1 4 UV lamp 100% (90 min) 0.0356 min−1
WO3 commercial (sigma-Aldrich)234 Commercial 25 mg L−1 750 mg L−1 3 UV lamp 100% (90 min) 0.0093 min−1
Pd/TiO2 (1%)236 UV-reduction 1 mg L−1 ∼50 mg L−1 Natural sunlight 100% (10 min) 52.1 ± 5.1 × 10−2 min−1
Pt/TiO2 (1%)236 UV-reduction 1 mg L−1 ∼50 mg L−1 Natural sunlight ∼90% (10 min) 7.6 ± 501 × 10−2 min−1
Cu (0.045 wt%)–TiO2237 Microwave assisted impregnation method 4 mg L−1 (20 mL) 1 g L−1 5.2 Lamps: 8 W, 77 mW cm−2 100% (120 min) 0.0506 min−1
TiO2 Evonik P25238 Sol–gel procedure 30 mg L−1 0.5 g L−1 UVC 100% (90 min) 0.046 min−1
TiO2 Evonik P25238 Sol–gel procedure 30 mg L−1 0.5 g L−1 Simulated solar light 100% (240 min) 0.022 min−1
1.5% au/TiO2238 Deposition precipitation method 30 mg L−1 0.5 g L−1 UVC light (254 nm) 100% (90 min) 0.071 min−1
1.5% au/TiO2238 Deposition precipitation method 30 mg L−1 0.5 g L−1 Simulated solar light 100% (180 min) 0.039 min−1
1.5% Ag/TiO2238 Deposition precipitation method 30 mg L−1 0.5 g L−1 UVC light (254 nm) 100% (45 min) 0.201 min−1
Simulated solar light 100% (240 min) 0.027 min−1
1.0% Cu/TiO2238 Deposition precipitation method 30 mg L−1 0.5 g L−1 UVC light (254 nm) 100% (90 min) 0.186 min−1
Simulated solar light 100% (240 min) 0.028 min−1
Au–Ag/TiO2238 Deposition precipitation method 30 mg L−1 0.5 g L−1 UVC light (254 nm) 100% (45 min) 0.143 min−1
Simulated solar light 100% (240 min) 0.025 min−1
Au–Cu/TiO2238 Deposition precipitation method 30 mg L−1 0.5 g L−1 UVC light (254 nm) 100% (45 min) 0.145 min−1
Simulated solar light 100% (240 min) 0.026 min−1
Fe-doped Titania (Fe/Ti molar ratio: 0.04%)239 Co-precipitation method 234 mg L−1 1 g L−1 Natural pH Xenon ozone free lamp (100 W) 95% (120 min) 29 × 10−3 min−1
F–Pd co-doped TiO2240 Microwave assisted hydrothermal method 30 mg L−1 1 g L−1 Sunlight 98.4% (40 min)
F–Pd co-doped TiO2240 Microwave assisted hydrothermal method 30 mg L−1 1 g L−1 Solar simulator 98.5% (220 min)
F–Pt co doped TiO2241 Microwave assisted hydrothermal method 20 mg L−1 (50 mL) 50 mg ∼5.1 Solar light >93% (90 min)
F–Pt co-doped TiO2241 Microwave assisted hydrothermal method 20 mg L−1 (50 mL) 50 mg ∼5.1 Simulated solar light ∼58% (360 min)
N–Cu co doped TiO2@f-SWCNT243 Sol–gel method 60 mg L−1 0.8 g L−1 6 Xenon lamp (200 W) 100% (60 min) 0.0512 min−1
Ag,P-co-doped g-C3N4244 Pyrolysis method 5 mg L−1 1000 mg L−1 9.0 Visible lamps (8 W each), λ: 465 ± 40 nm >99% (30 min) 2.06 × 10−1 min−1
Ce-doped ZnO246 Spray coating 6.332 μg L−1 Catalyst immobilized on 11.5 cm dia discs of area 104 cm2 6.28 UVA lamp (36 W) 1.09 × 10−2 min−1
Zn–TiO2/pBC247 Modified sol–gel method 10 mg L−1 (160 mL) 0.2 g 5.03 Xenon lamp (50 W) with 420 nm cut-off filter 80.8% (180 min) 0.0085 min−1
Fe3O4/ZnO, (H2O2:100 mg L−1)248 Polyol-mediated preparation 100 μg L−1 (20 mL) 200 mg L−1 7 UVA lamp (15 W), λ: 365 nm, 4 mW cm−2 ∼100% (240 min)
Bi2O3–TiO2/PAC251 Two-stage calcination method 20 mg L−1 (250 mL) 0.05 g 11 Solar light–xenon arc lamp (300 W) ∼100% (30 min) 0.159 min−1
LDH–TiO2254 Impregnation process 20 mg L−1 (100 mL) 50 mg 10 UVA lamp (λ: 300–400 nm, 300 W) 100% (360 min)
Poly(ethylene terephthalate)-10% TiO2257 Solvent casting method 1 mg L−1 (100 mL) 50 mg L−1 Xenon lamp (simulated solar light): 1.5 kW, 500 W m−2 98% (360 min) 0.015 min−1
BiVO4/SrTiO3 (1%)258 Self-template method under hydrothermal condition 10 mg L−1 (50 mL) 0.05 g Xenon lamp (500 W) 91% (60 min)
0.75 CuOx–BiVO4259 Polyol-reduction method 0.5 mg L−1 500 mg L−1 (persulfate: 100 mg L−1) Simulated solar light 100% (30 min) 0.0991 min−1
WO3–MWCNT261 Hydrothermal method 10 mg L−1 2.0 g L−1 Solar simulator–xenon arc lamp (300 W), 420–630 nm 88.5% (180 min)
Magnetic ZnO@g-C3N4265 In situ growth 30 mg L−1 (1000 mL) 0.65 g L−1 5.6 UVC lamp (10 W) 90.4% (60 min) 0.0384 min−1
5 wt% Ag/g-C3N4268 Photo-reduction method 10 μM (100 mL) 5 mg Natural pH Xenon lamp (300 W) with a 400 nm cut-off filter 97.5% (60 min)
rGO–WO3269 Hydrothermal method 10 mg L−1 1.0 g L−1 No pH adjustment Xenon arc lamp: 200 W (420–630 nm) >98% (180 min) 1.607 h−1
Ag@Ag2O-2.5 wt% graphene270 Precipitation method 1 mg L−1 0.05 g L−1 Xenon lamp (300 W) with a cut-off filter (λ > 280 nm), 37.7 mW cm−2 ∼100% (90 min) 0.038 min−1
Immobilized TiO2-2.7% rGO271 Polymer assisted hydrothermal deposition method 5 mg L−1 Bundle of thirty 10 cm photocatalyst-coated SOF (25 mL) placed on a petri disc 6 High-pressure UV mercury vapor lamp (160 W) 92% (180 min) 0.757 h−1
Cu2O/rGO-80 (80 refers amount of GO (mg) used in preparation of rGO)272 Wet chemical method 5 mg L−1 (80 mL) 20 mg Xe lamp: 300 W (420 nm cut-off filter) 50% (120 min) 0.00525 min−1
rGO–TiO2/sodium alginate (1[thin space (1/6-em)]:[thin space (1/6-em)]3)273 Hydrothermal method 10 ppm (200 mL) 0.5 g L−1 Neutral High-pressure mercury lamp (100 W) >99% (45–90 min) 0.108 min−1
WO3–g-C3N4 (referred as WCN-8)275 Hydrothermal method 10 mg L−1 1.0 g L−1 No pH adjustment Xenon arc lamp (300 W), 420–630 nm 91.7% (240 min)
Ce0.8Gd0.2O2−δ/TiO2276 Modified Pechini method 25 mg L−1 (300 mL) 30 mg Mercury lamp (15 W) 97% (120 min) 0.2959 mg−1 min−1
Ag2S/Bi2S3/g-C3N4277 Hydrothermal 20 mg L−1 0.25 mg mL−1 7 Xenon lamp (visible light): 300 W 97.3% (90 min) 0.0642 min−1
BiOCl/g-C3N4/Cu2O/Fe3O4279 Co-precipitation method 100 μM 0.2 g L−1 6.5 Xenon lamp 99.5% (60 min) 0.0543 min−1
rGO@Cu2O/BiVO4280 Solution method 0.5 mg L−1 100 mg (250 ml) 7 LED light (30 W) ∼98.5% (270 min)
Bi2WO6/TiO2281 Hydrothermal method 10 mg L−1, [ozone]: 1.5 mg L−1 0.2 g L−1 5.25 Simulated sunlight 97.1% (180 min) 1.83 × 10−2 min−1
MIL-53(Co/Fe)/10 wt% MoS2282 Hydrothermal through in situ growth 10 mg L−1, (peroxymonosulfate: 0.2 g L−1) 0.01 g L−1 6 Visible light 99% (60 min)
Bi2O3/C3N4/TiO2@C283 Hydrothermal and calcination 5 mg L−1 1 g L−1 5 Visible light 100% (100 min)
Ag (1 wt%)/g-C3N4/Bi3TaO7284 Photo deposition method 5 mg L−1 25 mg (50 mL) Xenon lamp (300 W) 98% (25 min) 0.1499 min−1
Ag3PO4/Bi4Ti3O12-20%285 In situ growth method 5 ppm (50 mL) Xenon lamp: 300 W (λ > 400 nm) ∼77% (40 min) 0.0372 min−1
Ag2O–KNbO3 (Ag–Nb molar ratio: 0.15/1)286 In situ growth 5 ppm 0.3 mg mL−1 Visible-light irradiation 91% (40 min) 0.0603 min−1
97.9% Ag3PO4/2.1% g-C3N4288 In situ precipitation method 1 mg L−1 (100 mL) 5 mg Neutral pH Xenon lamp (300 W), λ > 400 nm, 138.7 mW cm−2 ∼99% (90 min) 0.063 min−1
Fe3O4–ZnO@g-C3N4289 In situ growth 30 mg L−1 0.5 g L−1 7 UVC lamp (10 W) 95% (90 min) 0.0351 min−1


3.4 Ibuprofen

Ibuprofen (IPF) is a drug belonging to the class of propanoic acid derivatives and is extensively used in the treatment of fever, pain in human beings, inflammatory disorders, muscle problems, including migraines, rheumatoid arthritis, analgesic and painful menstrual periods.22,292 It is slightly soluble in water, stable, is eliminated from the body through urine and does not undergo biodegradation. As a result, it can be found in water samples of different origins originating from municipal wastewater treatment plant effluents, groundwater through leaching and natural water and cannot be treated through conventional wastewater treatments. The presence of ibuprofen even in low concentration through water affects the reproduction of aquatic animals, including the photosynthesis of aquatic plants. Ibuprofen can leach into ground water and soil in daily life. In view of this, several studies have been made using metal oxide and graphitic material related photocatalysts to make wastewater free from ibuprofen.293–357
3.4.1 Metal oxides. The photocatalytic degradation of ibuprofen has been reported in the literature using TiO2, ZnO and other metal oxides.294–306 Jallouli et al.294 used a TiO2/UV-LED system to study the photocatalytic degradation of ibuprofen present in ultrapure water (UP), the secondary treated effluent of a municipal wastewater treatment plant (WWTP) and pharmaceutical industry wastewater (PIWW). They observed the removal of ibuprofen below the detection limit in the case of UP and PIWW compared to municipal water. Their investigations inferred the higher degradation of IBU at near natural pH (5.3) of UP and PIWW compared to acidic (3.0) and alkaline (9.0) pH. In another study, the photocatalytic degradation of ibuprofen in water was carried out using TiO2 nanoparticles/UV light.295 The emerging findings established the faster depletion of ibuprofen with TiO2/UV (pH: 5.05) and followed pseudo-first-order kinetics (k: 1.0 min−1). TiO2 (0.03 g) resulted in almost 100% (5 min) photodegradation of ibuprofen in aqueous solution (pH: 5.0) on irradiation by a mercury lamp (125 W).296

The photodegradation of ibuprofen has been tested as a function of catalyst type (TiO2 and ZnO), loading (50–500 mg L−1), initial drug concentration (10, 40, 80 mg L−1) and wavelength (200–600 nm) of irradiation.297 The photocatalytic efficiency was found to be greater than 90% in 15 min under UVA and visible-light irradiation corresponding to an initial concentration of ibuprofen of 10 mg L−1 and amount of photocatalysts (TiO2 and ZnO) of 100 mg L−1. These findings also indicated over 90% conversion of the drug within 8 min with k-values of 0.382 and 0.326 min−1 under UVA for TiO2 and ZnO, respectively, and it correspondingly decreased to 0.199 and 0.144 min−1 under visible light. Tanveer and others298 used UV and solar irradiation to compare the photocatalytic degradation of ibuprofen in water using TiO2 and ZnO. A much higher rate of degradation was observed in UV for TiO2 (99%) compared to ZnO (86%) after 15 min compared to solar degradation.

The degradation of ibuprofen using a heterogeneous ZnO photocatalyst irradiated with UVC achieved 82.97% removal efficiency within a reaction time of 95 min under optimized conditions (pH: 6.7, ZnO loading: 583 mg L−1, initial IBP concentration: 1.5 mg L−1, humic acid concentration: 54 mg L−1).299 The reactive species responsible for oxidizing ibuprofen were found to be h+, O2·, H2O2, and ·OH. In another experiment, ZnO–Ce (0.50 g L−1) showed 60% removal of ibuprofen (20 ppm) under acidic conditions after 120 min under UVC irradiation.300 Holes played a vital role in the degradation process of ibuprofen and it displayed good degradation activity even after 3 cycles under UV light. Hexagonal α-Fe2O3 flakes have removed up to 80% of ibuprofen in a combination of adsorption treatment followed by UV (265 nm) irradiation.301 TiO2 immobilized on glass coupled with simulated solar irradiation also eliminated ibuprofen and its derivatives.302 Investigations on the photocatalytic activity of TiO2,303,304 ZnO,304,305 and ZnO membrane306 have also been reported in the remediation of water from ibuprofen.

3.4.2 Doped metal oxides. N,S-co-doped TiO2 exhibited high photocatalytic activity in the degradation of ibuprofen under simulated solar irradiation due to the synergistic effects of N and S co-doping in TiO2 owing to the separation of photogenerated electrons and holes and higher visible-light adsorption.307 Reusability tests of the N,S–TiO2 photocatalyst showed that its catalytic activity was not significantly altered even after 6 cycles. C,N,S-co-doped TiO2 prepared by thermally treating hydrothermally prepared mesoporous TiO2 (anatase/brookite) and thiourea in a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 wt. ratio demonstrated complete degradation of ibuprofen under visible light within 5 h in contaminated water.308

Bi (0.25 wt%) and Ni (0.5 wt%) doped TiO2 photocatalysts synthesized by a sol–gel method under irradiation of solar light for 6 h achieved degradation of ibuprofen by 89% and 78% repectively.309 The degradation of ibuprofen followed kinetics in accordance with the Langmuir–Hinshelwood model. In addition, La3+-doped TiO2 monolith,310 Cu-doped LaFeO3,311 Cu2O-doped TiO2 nanotube arrays,312 C,N-co-doped mesoporous TiO2313 and TiO2 co-doping with urea and functionalized CNT314 photocatalysts also displayed enhanced photocatalytic degradation of ibuprofen in aqueous solution.

3.4.3 Metal oxide–metal oxide composites. Lin et al.315 prepared TiO2 nanofibers wrapped in BN nanosheets by an electrospinning method, which were examined as a photocatalyst for the removal of ibuprofen from contaminated water under UV irradiation. The ibuprofen was almost completely removed after 2 h owing to wrapping of the BN nanosheets to facilitate improved light absorption and efficient separation of the electron–hole pairs. Investigations were also made on the reusability and regeneration capability of the prepared photocatalyst on the degradation of ibuprofen. Activated carbon (90 wt)% impregnated with TiO2 showed 92% removal efficiency for ibuprofen solution under UV light within 4 h due to the synergy of adsorption and photodegradation.316 FeO,317 Fe3O4@MIL-53(Fe),318 Fe3O4/Bi2WO6,319 BiOBr0.9I0.1/Fe3O4@SiO2,320 and Ag/Ag2O321 nanocomposites also displayed enhanced removal of ibuprofen under visible-light irradiation.

Ag and Fe3O4 co-modified WO3−x (Ag/Fe3O4/WO3−x) composites were fabricated by hydrothermal and photodeposition processes and showed almost complete photocatalytic-Fenton degradation of ibuprofen (and diclofenac), as evident from (Fig. 13(a) and (b)).322 This is attributed to the surface plasmon resonance effect of Ag, separation of photogenerated carriers and heterostructures of Ag/Fe3O4/WO3−x. In addition, the possibility of absorption of light greatly improving the photocatalytic-Fenton degradation efficiency cannot be ruled out. The fabricated Ag/Fe3O4/WO3−x also exhibited good photocatalytic-Fenton stability in the photodegradation of ibuprofen (and diclofenac), as indicated by the almost unchanged degradation rate of the antibiotic in (Fig. 13(c) and (d)). The degradation and charge transfer mechanism involved in the removal of the ibuprofen and diclofenac have also been proposed and are displayed in Fig. 13(e).


image file: d3lf00142c-f13.tif
Fig. 13 Photocatalytic-Fenton degradation of (a) ibuprofen and (b) diclofenac by Fe3O4, WO3–x, Fe3O4/WO3–x, and Ag/Fe3O4/WO3–x samples. (c and d) Corresponding recycling study and stability of Ag/Fe3O4/WO3−x. (e) Schematic illustration of the possible catalytic degradation mechanism and charge transfer of Ag/Fe3O4/WO3–x under light irradiation (modified image). Reproduced from ref. 322 with permission from ACS (2021).

Lenzi et al.323 showed that the photocatalytic degradation of ibuprofen (10 ppm) solution (pH: 7) by 0.3 g L−1 of Ag/ZnO/CoFe2O4 (5 wt%) exhibited removal efficiencies of 80% and 47% under artificial and solar radiation, respectively. These studies also confirmed the recovery and reuse of the catalyst after 3 cycles without significant loss of catalytic activity. Visible-light-driven mesoporous hierarchical BiOBr/Fe3O4@SiO2 (dose: 1 g L−1) photocatalyst degraded ibuprofen (initial concentration: 2 mg L−1) almost completely in 60 min.324 Further studies have shown BiOBr/Fe3O4@SiO2 maintaining its initial photocatalytic activity (∼80%) even after five cycles. In another study, a magnetically separable Fe3O4–SiO2-coated TiO2 composite demonstrated excellent photocatalytic activity.325 An immobilized TiO2/ZnO-sensitized copper(II) phthalocyanine heterostructure displayed about 80% degradation of ibuprofen (initial conc.: 5 mg L−1) after 4 h of irradiation under 365 nm UV.326 The studies revealed a small decline in the IBF degradation (77%) after the 5th cycle. PANI-coated WO3@TiO2,327 polyacrylonitrile (PAN)–MWCNT/TiO2–NH2,328 TiO2 nanoparticles and C-nanofiber-modified magnetic Fe3O4 nanospheres (TiO2@Fe3O4@C-NF),329 carbon dots/Fe3O4@carbon sphere pomegranate-like composites,330 PVDF–ZnO/Ag2CO3/Ag2O,331 and PAN–MWCNT nanofiber crosslinked TiO2–NH2 nanoparticles332 have also been examined for their photodegradation performance for ibuprofen.

3.4.4 Graphitic materials. Hernández-Uresti et al.333 observed the following order for the degradation of different pharmaceutical compounds in aqueous solution (pH ∼ 5.5) using g-C3N4 under UV-vis irradiation: tetracycline (86%) > ciprofloxacin (60%) > ibuprofen (20%). Wang and coworkers334 undertook investigations on the degradation of pharmaceutical contaminants by bubbling a gas-phase surface discharge plasma combined with g-C3N4 photocatalysis. These findings disclosed 82% and 100% removal of ibuprofen and tetracycline hydrochloride after 25 min, corresponding to initial concentrations of 60 and 200 mg L−1, respectively. A photocatalytic study of hydrothermally prepared reduced-graphene-oxide-loaded HoVO4–TiO2 revealed enhanced photodecomposition efficiency of rGO–HoVO4–TiO2 (∼96%) compared to rGO–HoVO4 (75%), HoVO4 (67%), rGO–TiO2 (30%) or TiO2 (10%) in the removal of ibuprofen over 60 min.335 The findings also showed ibuprofen decomposition to depend mainly on superoxide radicals photogenerated from rGO–HoVO4–TiO2 under visible-light illumination.

Acidified g-C3N4/polyaniline/rGO@biochar (0.5 mg L−1) nano-assemblies degraded ibuprofen (20 mg L−1) to the extent of 98.4% in 50 min under exposure to visible light.336 Such significant performance is attributed to multiple reasons, such as highly separated charges, enhanced visible absorption and diffusion. The major reactive species in the degradation process for ibuprofen involved hydroxyl and superoxide radical anions. Akbarzadeh et al.337 explored the photodegradation of ibuprofen solution in the presence of a hydrothermally fabricated g-C3N4/Ag/AgCl/BiVO4 microflower composite as photocatalyst under visible light and compared its performance with BiVO4, g-C3N4/BiVO4 and Ag/AgCl/BiVO4. These findings revealed remarkably enhanced degradation efficiency of g-C3N4/Ag/AgCl/BiVO4 (94.7%) compared to g-C3N4 (6.5%), BiVO4 (11.4%), g-C3N4/BiVO4 (68.6%), or Ag/AgCl/BiVO4 (88.3%) in 1 h corresponding to a photocatalyst dosage of 0.25 g L−1 and initial concentration of 2 mg L−1. The reduced band gap energy and recombination rate of the g-C3N4/Ag/AgCl/BiVO4 photocatalyst are ascribed to charge transfer along the heterojunction. The photocatalytic degradation performance of IPF increases with the (121)/(040) XRD plane intensity ratio of BiVO4, Ag/AgCl/BiVO4, g-C3N4/BiVO4 and g-C3N4/Ag/AgCl/BiVO4 and is found to be in good agreement with the photoluminescence findings.

A hierarchical assembly of Ag (7%)/g-C3N4/kaolinite composite fabricated following an in situ calcination and photodeposition process exhibited 99.9% degradation of ibuprofen (k: 0.01128 min−1) after 5 h under visible-light irradiation compared to g-C3N4, g-C3N4/kaolinite and Ag/g-C3N4.338 This outcome is due to the stronger adsorption property, efficient separation and transfer of electron–hole pairs. In addition, the presence of monodispersed Ag nanoparticles in the g-C3N4/kaolinite sheets led to more active sites, accounting for this. The efficient photocatalytic degradation of ibuprofen has also been reported in aqueous solution using graphene quantum dots/AgVO3 nanoribbons,339 g-C3N4/MIL-68(In)–NH2 composites,340 graphene oxide and TiO2 heterostructures doped with F,341 reduced-graphene-oxide–TiO2/sodium alginate 3-dimensional structure aerogel273 and Fe3O4/graphene/S-doped g-C3N4342 also exhibited enhanced visible-light photocatalytic activity for the degradation of ibuprofen.

3.4.5 Heterojunction and Z-scheme-based photocatalysts. A TiO2/g-C3N4 (5%) photocatalyst exhibiting a sea urchin morphology with interface effects was synthesized by a solvothermal method.343 Its application in the photocatalytic degradation of ibuprofen showed significantly enhanced performance under irradiation by visible light for 60 min. The formed superoxide radicals and holes were assigned as the main active species involved in the photodegradation of ibuprofen. The photocatalytic performance of this catalyst after 5 cyclic experiments indicated its good stability. Wang et al.344 fabricated atomic-scale g-C3N4/Bi2WO6 comprising ultrathin g-C3N4 nanosheets and monolayer Bi2WO6 nanosheets (1[thin space (1/6-em)]:[thin space (1/6-em)]4 mol ratio) by a hydrothermal reaction. Such an assembly of 2D/2D heterojunctions removed 96.1% ibuprofen under visible-light irradiation within 60 min due to a synergistic effect.

Kumar and others345 synthesized a magnetically recyclable direct-contact Z-scheme g-C3N4/TiO2/Fe3O4@SiO2 heterojunction nanophotocatalyst and recorded 97% removal of ibuprofen solution (pH: 3) after 15 min under irradiation by visible light (∼330 W m−2). Such excellent performance of a magnetically recyclable direct-contact Z-scheme nanophotocatalyst was attributed to the low recombination rate of photogenerated e and h+. Visible-light-assisted persulfate activation by an SnS2 (0.5%)/MIL-88B(Fe) Z-scheme heterojunction achieved 100% removal of ibuprofen in 120 min.346 This was found to be 54 and 4 times higher than SnS2 and SnS2 (0.5%)/MIL-88B(Fe), respectively. Such findings could be ascribed to the structure and crystallinity of the photocatalysts. In another reported study, an optimized Z-scheme based 1D/2D FeV3O8/g-C3N4 composite comprising 10% FeV3O8 achieved a maximum degradation rate for ibuprofen of 95% at 85 min under visible-light irradiation.347 Kinetic studies established that the rate constant is 4 times that of g-C3N4 nanosheets. However, the presence of 30% FeV3O8 in g-C3N4 decreased the degradation efficiency to 52.8%.

Heterostructure g-C3N4/Bi2WO6/rGO nanocomposites prepared by microwave- assisted treatment for 120 min in a hydrothermal method undertook the maximum photocatalytic degradation of ibuprofen (93.9%) under visible-light illumination.348 In addition, g-C3N4@NiO/Ni@MIl-101,349 Bi5O7I–MoO3,350 AgSCN/Ag3PO4/C3N4,351 N–TiO2@SiO2@Fe3O4,352 g-C3N4/CQDs/CdIn2S4,353 direct Z-scheme Co3O4/BiOI,354 a double Z-scheme system of α-SnWO4/UiO-66(NH2)/g-C3N4,355 CdS/Fe3O4/TiO2356 and Ag2CO3/Ag2O/ZnO357 heterojunctions also exhibited excellent photocatalytic degradation of ibuprofen.

Table 5 records the performance data of different photocatalysts on the removal of ibuprofen from wastewater.

Table 5 Performance data on removal of ibuprofen in waste in water using variety of photocatalysts
Photocatalyst Preparation method IPF Catalyst dose pH Light source % degradation Rate constant
TiO2 Degussa P25 (80% anatase and 20% rutile)294 Commercial 213 mg L−1 2.5 g L−1 5.0–5.3 UV-LEDs (10 W), 365 nm, 375 W m−2 100% (5 min) 24 × 10−3 min−1
TiO2 nanoparticles (Degussa P25)295 Commercial 5 μg mL−1 (50 mL) 134.5 mg 5.5 UV light: 15 W, 365 nm 100% (10 mi) 1.0 min−1
TiO2 (Vetec, 98% of purity)296 Commercial 10−4 M (100 mL) 0.03 g 5 Mercury lamp (125 W) 100% (5 min)
TiO2 P-25 Degussa (75[thin space (1/6-em)]:[thin space (1/6-em)]25 w/w mixture of anatase[thin space (1/6-em)]:[thin space (1/6-em)]rutile)297 Commercial 10 mg L−1 100 mg L−1 4 UVA ∼100% (18 min) 0.382 min−1
ZnO Sigma Aldrich297 Commercial 10 mg L−1 100 mg L−1 4 UVA ∼100% (18 min) 0.326 min−1
TiO2 P-25 Degussa (75[thin space (1/6-em)]:[thin space (1/6-em)]25 w/w mixture of anatase[thin space (1/6-em)]:[thin space (1/6-em)]rutile)297 Commercial 10 mg L−1 100 mg L−1 4 Visible ∼94% (18 min) 0.199 min−1
ZnO Sigma Aldrich297 Commercial 10 mg L−1 100 mg L−1 4 Visible ∼90% (18 min) 0.144 min−1
TiO2 (Sigma-Aldrich)298 Commercial 20 mg L−1 1.5 g L−1 3 UV lamp (40 W) 99% (15 min) 0.54 min−1
ZnO (Sigma-Aldrich)298 Commercial 20 mg L−1 1.0 g L−1 7 UV lamp (40 W) 86% (15 min) 0.31 min−1
ZnO (Nano pars Spadana)299 Commercial 5 mg L−1 (humic acid: 50 mg L−1) 500 mg L−1 7 125 W medium-pressure Hg lamp (UVC) 98% (100 min)
ZnO–Ce300 Precipitation method 20 ppm 0.5 g L−1 3 UV light: 125 W Hg without bulb 60% (120 min) 6.86 × 10−3 min−1
ZnO–Ce; H2O2: 0.5 m mole per L300 Precipitation method 20 ppm 0.5 g L−1 3 UV light: 125 W Hg without bulb 70% (120 min)
TiO2 (Degussa P25) dispersed powder302 Commercial 25 mg L−1 0.2 g L−1 4.5 Solar simulator exposed to xenon lamp irradiation ∼95% (150 min) 0.2378 mg L−1 min−1 (zero order), 0.0251 min−1 (first order), 0.0034 L mg−1 min−1 (second order)
TiO2 immobilized on the active coated glass302 Chemical vapour deposition 25 mg L−1 0.2 g L−1 4.5 Solar simulator and exposed to xenon lamp irradiation 100% (1480 min) 0.0124 mg L−1 min−1 (zero order), 0.0012 min−1 (first order), 0.0001 L mg−1 min−1 (second order)
TiO2 Degussa (P-25)303 Commercial 4 mg L−1 20 mg L−1 7.8 125 W Hg vapor lamp, 10.75 mW cm−2 >98% (30 min)
TiO2 Degussa P25 (ref. 304) Commercial 5 mg dm−3 50 mg dm−3 Mercury lamp (150 W), λ < 300 nm ∼89% (60 min) 0.0425 min−1
ZnO Degussa P25 (ref. 304) Commercial 1 mg dm−3 50 mg dm−3 Mercury lamp (150 W), λ < 300 nm 60% (30 min) 0.0328 min−1
ZnO nanoparticles305 Chemical method 60 ppm 10 mg L−1 Four UV-vis solarium lamps (60 W) 24% (180 min) 0.055 min−1
PVDF- ZnO/Ag2CO3/Ag2O membrane306 Casting solution using wet phase inversion method 10 ppm (300 mL) 1.96 wt% (membrane area: 12.56 cm2) White light-emitting diode lamp (λ > 400 nm, 100 W) 49.96% (180 min)
N,S-co-doped TiO2 nanoparticles307 Sol–gel and hydrothermal methods 5 mg L−1 (50 mL) 2.0 g L−1 6 Simulated solar radiation: 350 W xenon lamp 85% (90 min) 0.062 min−1
C–N–S co-doped TiO2308 Thermal treatment method 20 ppm (200 mL) 0.5 g L−1 LED lamp (λmax: 420 nm, 1 mW cm−2) ∼100% (300 min) 0.021 min−1
Bi (0.25 wt%) doped TiO2309 Sol–gel method 25 ppm 2 g L−1 6 UV (36 W, 254 nm) 89% (360 min) 0.0064 min−1
Ni (0.5 wt%) doped TiO2309 Sol–gel method 25 ppm 2 g L−1 6 UV (36 W, 254 nm) 78% (360 min) 0.0046 min−1
La3+(2%)-doped TiO2 monolith310 Sol–gel method 50 mg L−1 (70 mL) 0.1 g 5 Sunlight 96.9% (150 min) 2.2 × 10−2 min−1
C,N-co-doped mesoporous TiO2313 Hydrothermal method 20 ppm (220 mL) 0.5 g L−1 High-pressure Hg lamp (150 W), λmax: 254 nm 98.9% (120 min) 0.0377 min−1
C,N-doped mesoporous TiO2313 Hydrothermal method 20 ppm (220 mL) 0.5 g L−1 LED lamp (visible light, λmax: 420 nm, 1 mW cm−2) 100% (120 min) 0.0207 min−1
N doped CNT COOH/TiO2 (anatase/rutile: 20/80)314 Hydrothermal 5 mg L−1 ppm 400 mg L−1 Natural pH LED light: 240 W, 40 mW cm−2 and 410 nm 85–86% (120 min) 4.45 × 10−3–1.22 × 10−2 min−1
Activated carbon impregnated with TiO2316 Sol–gel method 25 mg L−1 (20 mL) 1.6 g L−1 4.3 UV lamp: 15 W, 254 nm 92% (240 min)
Fe3O4@MIL-53(Fe)318 Calcination (400 °C) 10 mg L−1 (50 mL), H2O2 (20 mM) 0.4 g L−1 Xenon lamp (500 W with 420 nm cut-off filter) 99% (60 min) 4.71 × 10−2 min−1
Fe3O4/Bi2WO6319 Two-step approach 10 mg L−1 (70 mL) 70 mg 4.7 Solar light >80% (120 min) 0.0144 min−1
Ag/Fe3O4/WO3−x/H2O2 (10 mM)322 Simultaneous calcination 10 mg L−1 (30 mL) 30 mg Xenon lamp (500 W) with optical filter (λ ≥ 420 nm) ∼100% (90 min)
Ag/ZnO/CoFe2O4323 Coating CoFe2O4 with ag/ZnO using Pechini method 10 ppm 0.3 g L−1 7 UV light (125 W medium-pressure Hg lamp) 80% (60 min) 0.03905 min−1
BiOBr/Fe3O4@SiO2324 Solvothermal 2 mg L−1 1 g L−1 (50 ml) 7 Fluorescent lamp (visible light) ∼99% (60 min) 0.08 min−1
TiO2/ZnO/copper phthalocyanine (CuPc)326 Multiple steps 5 mg L−1 (50 mL) Film 6.5 Hg lamp with 365 nm cut-off filter, 1,2 W cm−2 80% (240 min) 0.42 h−1
PAN–MWCNT/TiO2–NH2328 Electrospinning 5 mg L−1 (100 mL) 15 mg L−1 2 UVA lamp (315–400 nm) of 40 W ∼100% (120 min)
Carbon dots/Fe3O4@carbon sphere (in presence of persulfate)330 Solvothermal method 50 μmol L−1 0.3 g L−1 Xenon lamp (350 W) with a glass filter (λ > 420 nm) 96% (120 min)
PAN–MWCNT/TiO2–NH2 composite nanofibers332 Multiple steps 5 mg L−1 (100 mL) 15 mg 2 Xenon lamp (125 W) with cut-off filter (λ > 400 nm), 0.1 W cm−2 100% (210 min)
g-C3N4333 Polycondensation 20 mg L−1 (200 mL) 200 mg 5.5 Xenon lamp (35 W) 20% (4 h)
Reduced graphene oxide–HoVO4–TiO2335 Hydrothermal 10 mg L−1 40 mg L−1 7 Tungsten lamp (150 W), (λ > 4900 nm) ∼96% (60 min)
g-C3N/ag/AgCl/BiVO4337 Hydrothermal 2 mg (50 mL) 0.25 g L−1 4 Visible light 94.7% (60 min)
Ag (7%)/g-C3N4/kaolinite338 Two steps 5 ppm (50 mL) 50 mg Xenon lamp (500 W with 400 nm cut-off filter) 99.9% (300 min) 0.01128 min−1
Graphene quantum dots (3 wt%)/AgVO3339 Hydrothermal 10 mg L−1 (50 mL) 0.01 g Xenon lamp (350 W with λ > 420 nm) ∼100% (180 min) 0.1678 min−1
g-C3N4 (10 wt%)/MIL-68(In)–NH2 composites340 In situ solvothermal assisted by ultrasonication 20 mg L−1 0.15 g 4 Xenon lamp (300 W with λ > 420 nm) 93% (120 min) 0.01739 min−1
Graphene oxide/TiO2 doped with F (BrO3 100 μg L−1)341 Hydrothermal 100 μg L−1 0.05 g L−1 5.2 Low-pressure Hg lamp (10 W), (26 μW cm−2) ∼100% (60 min) 0.4504 min−1
rGO–TiO2/sodium alginate273 Hydrothermal 10 ppm (200 mL) 0.5 g L−1 7 High-pressure Hg lamp (100 W), (13.5 W m−2) ∼100% (90 min) 0.047 min−1
TiO2/5% g-C3N4343 Solvothermal 5 mg L−1 50 mg 7 Xenon lamp (259 W) ∼90% (60 min) 0.03833 min−1
g-C3N4/Bi2WO6 (1[thin space (1/6-em)]:[thin space (1/6-em)]4 molar ratio)344 Hydrothermal 25 μM 0.2 g L−1 Xenon lamp (300 W) with 420 nm cut-off filter ∼96.1% (60 min) 0.062 min−1
g-C3N4/TiO2/Fe3O4@SiO2345 Sol–gel method 2 mg L−1 (50 mL) 50 mg 7 Visible light, 330 W m−2 97% (15 min)
FeV3O8 (10%)/g-C3N4347 Dispersion, grinding and calcination 10 ppm (30 mL) 10 mg Xenon lamp (300 W) with UV cut-off filter (λ: 420 nm) 95% (85 min) 0.03 min−1
g-C3N4/Bi2WO6/rGO348 Microwave assisted hydrothermal preparation 5 mg L−1 1.0 g L−1 4.3 Xenon lamp (300 W), λ > 420 nm 93% (240 min) 0.011 min−1
g-C3N4/Bi2WO6/rGO348 Microwave assisted hydrothermal preparation 5 mg L−1 1.0 g L−1 4.3 Sunlight 98.6% (240 min)
AgSCN/Ag3PO4/C3N4 (molar % of AgSCN: 11.3)351 Precipitation reaction 5 mg L−1 (100 mL) 50 mg Sunlight (500 W halide lamp) 91% (6 min) 0.46 min−1
N–TiO2@SiO2@Fe3O4352 Sol–gel method 2 mg L−1 (50 mL) 50 mg Fluorescent lamps (9 W), 320 μW cm−2 94% (300 min)
g-C3N4/CQDs/CdIn2S4353 Hydrothermal 80 mg L−1 (100 mL) 0.1 g 300 W xenon lamp with 420 nm cut-off filter, 200 mW cm−2 91% (60 min)
Co3O4/BiOI (1[thin space (1/6-em)]:[thin space (1/6-em)]2)354 Solvothermal 10 ppm (50 mL) 40 mg 11.3 60 W LED lamp with 420 nm cut-off filter 93.87% (60 min) 0.0945 min−1
α-SnWO4/UiO-66(NH2)/g-C3N4355 Solvothermal 10 mg L−1 (100 mL) 50 mg Simulated sunlight using high-pressure 300 W xenon lamp 95.5% (120 min) 0.017 min−1


3.5 Norfloxacin

Norfloxacin (NOR) is an effective antibacterial agent of the fluoroquinolone family and is widely used as a drug in clinical treatments for bacterial infections of urinary, biliary, and respiratory tracts, and gastrointestinal and skin infections.358–360 Norfloxacin has frequently been detected in municipal/wastewater treatment plants, is difficult to biodegrade and is predicted to be a potential risk to human beings and the environment. Therefore, it is considered a potential threat to the water environment and human health.361–422
3.5.1 Metal oxides. Reduced TiO2 (TiO2−x) samples comprising Cat.I-A (anatase), Cat.II-R (rutile) Cat.III-B (brookite) and a series of Cat.IV-A&R (anatase/rutile phases) mixed in different ratios showed about ∼100% photocatalytic degradation of norfloxacin in visible light (>400 nm).361 Such degradation of norfloxacin is guided by the specific surface area, concentration of Ti3+ and the density of oxygen vacancies of the photocatalysts. Haque and Muneer362 reported Degussa P25 (anatase: 75%, rutile: 25%) acting as an efficient photocatalyst for the photodegradation of norfloxacin in aqueous suspensions compared to other TiO2 powders. Cu2O particles prepared by a hydrothermal method showed a high degradation rate for norfloxacin (79.8%) with ·OH and ·O2 species playing major roles.363 The removal of norfloxacin has also been explored in a broad operating pH range via simulated solar-light-mediated bismuth tungstate Bi2WO6.364
3.5.2 Metal–metal oxides composites. Sayed et al.365 prepared immobilized {001}-faceted TiO2/Ti film by placing Ti plate water/2-propanol solvent and 0.02 M HF (pH: 2.62) under hydrothermal conditions at 180 °C for 3 h and exhibited the following order for the degradation of norfloxacin (10 mg L−1) under UV irradiation: Milli-Q-water (70.5%, k: 0.0504 min−1) > tapwater (∼55.1%, k: 0.03 min−1) > river water (44.9%, k: 0.009 min−1) > synthetic wastewater (39.89%, k: 0.005 min−1). Triangular silver nanoplates (T-Ag)/ZnO nanoflowers significantly enhanced the photocatalytic degradation of norfloxacin under visible light due to synergistic effects in the different water matrices.366 It was concluded that the degradation efficiency for norfloxacin by T-Ag/ZnO nanoflowers is guided by the choice of water source. In another report, Zhang et al.367 prepared triangular Ag nanoplate coated ZnO nanoflowers by a hydrothermal/dual-reduction method and studied its performance in the photocatalytic degradation of NF in aqueous solutions under visible-light irradiation. It should be noted that the improved photocatalytic degradation of NF activity could be ascribed to the synergetic effect and the unique surface plasmon resonance of triangular silver nanoplates in T-Ag/ZnO. In addition, photogenerated holes are considered to be the main oxidative species that account for the photocatalytic degradation of NF by T-Ag/ZnO composites under visible light. A chemically doped Prussian blue in CeO2 (doping ratio: 10%) photo-Fenton catalyst showed 88.93% degradation of norfloxacin in 30 min with ·OH acting as the major reactive species.368
3.5.3 Doped metal oxides. The effect of ion doping on the properties of photocatalysts has been receiving considerable attention in exploring their better performance for wastewater treatment applications.369 In this regard, the photocatalytic degradation of norfloxacin has been studied using an N-doped TiO2 catalyst under visible-light irradiation. Jin et al.370 also fabricated TiO2 doped with nitrogen to enhance its optical response through reduction in the band gap and carried out the photocatalytic degradation of norfloxacin under visible-light irradiation. These investigations indicated almost complete removal of norfloxacin within 30 min under optimum conditions (pH: 6.37, catalyst dose: 0.54 g L−1, norfloxacin: 6.03 mg L−1). Al-doped TiO2 achieved 93% norfloxacin removal in 2 h which was found to be ∼5 times higher than undoped TiO2 nanoflakes under visible light.371 The norfloxacin was completely degraded by visible-light-mediated C-doped TiO2 in 20 min corresponding to a concentration of 0.0313 mM and catalyst dosage of 2.0 g L−1.372 It was established that the hydroxyl radical plays an important role in the degradation process.

The photocatalytic degradation of norfloxacin (and ciprofloxacin) was found to be 90–93% under optimized conditions in B and Ce doped TiO2, irradiated by sunlight.373 Bi3+ and Fe2+ ion doped ZnO showed significant photocatalytic degradation of norfloxacin with the addition of HSO5 under solar irradiation and followed pseudo-first-order kinetics.374 The co-doped ZnO exhibited a lower band gap, which accounted for the increased absorption of solar irradiation and reduced electron and hole recombination, which facilitated high norfloxacin degradation compared to undoped ZnO. Fe-doped CeO2 exhibited about 95% photocatalytic degradation of norfloxacin in aqueous solution (pH: 8.0) within 180 min corresponding to an initial norfloxacin concentration of 2.5 mg L−1 and catalyst dose of 0.1 g L−1.375 An Ag-doped TiO2/CFA (coal fly ash) photocatalyst has also been used to monitor the photocatalytic degradation of norfloxacin.376

3.5.4 Metal oxide–metal oxide composites. A mesoporous Fe2O3–TiO2 photocatalyst showed complete norfloxacin removal from aqueous solution (pH: 7) within 120 min under UV illumination with a stoichiometric amount of H2O2.377 Trang et al.378 used an ordered SBA-15 mesoporous silica support synthesized by a sol–gel method using the triblock copolymer Pluronic P123 and immobilized with different amounts of photocatalyst TiO2 (TiO2[thin space (1/6-em)]:[thin space (1/6-em)]SiO2 ratios of 0, 0.25, 1.0 and 5.0). Subsequent investigations on the removal of norfloxacin revealed the better photocatalytic activity of 1.0TiO2/SBA-15 hybrid material in achieving 96.6% degradation of norfloxacin in 150 min under UV-light irradiation. Fe-complex/TiO2 composites comprising [FeII(dpbpy)2 (H2O)2]/TiO2, [FeII(dpbpy)(phen)2]/TiO2 and [FeII(dpbpy)(bpy)2]/TiO2 (dpbpy: 2,2′-bipyridine-4,4′-diphosphoric acid, phen: 1,10-phenanthroline, bpy: 2,2-bipyridyl) photocatalysts exhibited 98.5% degradation of norfloxacin in water under visible-light irradiation after 3 h.379 Further, the photocatalytic performance and cyclic stability of these composites were found to be much better than those of pure TiO2 or P25. An Ag2O/TiO2-zeolite composite fabricated through a modified sol–gel method exhibited high performance in the decomposition of norfloxacin under simulated solar-light illumination.380 This is a consequence of the narrow band gap of the photocatalyst, its enhanced light absorbance ability in the visible region and high charge separation efficiency.

FeVO4/Fe2TiO5 (2[thin space (1/6-em)]:[thin space (1/6-em)]1) synthesized via a one-pot hydrothermal method exhibited high photocatalytic activity and excellent stability for the removal of norfloxacin in aqueous solution under visible-light irradiation.381 This is ascribed to the synergistic effect of photogenerated electron–holes with radical OH· and h+. MIL-101(Fe)–NH2 immobilized on an α-Al2O3 sheet has also been investigated for effective norfloxacin elimination via a photo-Fenton process.382 Ag/AgCl–CeO2 composite photocatalysts fabricated by in situ interspersal of AgCl on CeO2 and subsequent photoreduction of AgCl to Ag exhibited enhanced photocatalytic activity in the photodegradation of norfloxacin under visible-light irradiation.383Fig. 14(a) shows the highest degradation efficiency (91%) for norfloxacin achieved by sample Ag/AgCl–CeO2 composites with an Ag mass ratio of 13.94 wt% (denoted AC-3) within 90 min under visible-light irradiation. It is also apparent from Fig. 14(b) and (c) that the photodegradation process followed a pseudo-first-order kinetic model with the highest rate constant (0.02279 min−1) for AC-3 compared to CeO2, Ag/AgCl, Ag/CeO2 and other AC composites. Fig. 14(d) shows the time-dependent UV-vis spectra of NOF solution for the AC-3 sample. ZnO/ZnS@biochar,384 ZnFe2O4/hydroxyapatite–Sn2+,385 (BiO)2CO3–Bi–TiO2,386 and Ag/AgCl/Ag2MoO4387 composites have also been reported as promising photocatalysts in the degradation of norfloxacin in water under UV irradiation.


image file: d3lf00142c-f14.tif
Fig. 14 (a) Photocatalytic degradation NOF curves; (b) kinetic curves of NOF degradation; (c) apparent rate constants for the degradation of NOF; (d) time-dependent UV–vis spectra of NOF solution for AC-3 sample (Ag/AgCl–CeO2). Reproduced from ref. 383 with permission from Elsevier (2017).
3.5.5 Graphitic composites.
3.5.5.1 g-C3N4-based composites. Fei et al.388 investigated the photocatalytic degradation of norfloxacin in the presence of a sunlight-driven mesoporous g-C3N4. The results showed 90% decomposition of norfloxacin in 1.5 h under simulated sunlight irradiation. Co/g-C3N4, Co/g-C3N4/H2O2 and Co/g-C3N4/PMS composite photocatalysts exhibited better performance compared to pure g-C3N4 in the photocatalytic degradation of norfloxacin under visible-light irradiation.389 The optimization and variations of different parameters have been used to study the photocatalytic degradation of norfloxacin in the presence of ZnO/g-C3N4/Fe3O4 under visible light.390 These findings indicated a removal rate of norfloxacin greater than 90% in 120 min for a catalyst concentration of 1.43 g L−1, solution pH 7.12 and norfloxacin concentration of <8.61 mg L−1. Shuttle-like CeO2/g-C3N4 combined with persulfate391 and NiWO4 nanorods anchored on g-C3N4 nanosheets392 also exhibited enhanced degradation of norfloxacin under visible light.
3.5.5.2 Graphene-based composites. A TiO2/Bi2WO6/rGO (0.5%) photocatalyst attained about 87.79% removal of norfloxacin in water under visible-light irradiation after 60 min and was found to be superior to its individual components under optimal conditions.393 Such enhanced catalytic activity of TiO2/Bi2WO6/rGO arises due to the ligand–metal electron transfer mechanism. According to Zhao et al.,394 an rGO/Bi2WO6 composite exhibited outstanding photocatalytic activity for norfloxacin degradation in an aquatic environment under visible-light irradiation, as evident from the time-dependent-UV spectrum and time-dependent-HPLC spectrum displayed in Fig. 15(a) and (b), respectively. Fig. 15(c) and (d) indicate about 87.49% degradation of norfloxacin within 180 min compared to Bi2WO6, under visible-light irradiation. Additional investigations revealed ·OH and e playing dominant roles in the photocatalytic degradation of norfloxacin. N-doped TiO2/graphene exhibited enhanced photocatalytic degradation under UV-light irradiation.395 It is suggested that graphene acts as an efficient “electron pump”, thereby promoting the separation of carriers to account for the observed photodegradation.
image file: d3lf00142c-f15.tif
Fig. 15 (a) The time-dependent UV spectrum, (b) the time-dependent-HPLC spectrum, (c) the photodegradation curve, and (d) photocatalytic degradation rate of norfloxacin. Reproduced from ref. 394 with permission from Elsevier (2021).

Wu et al.396 reported a UV-assisted nitrogen-doped reduced graphene oxide/Fe3O4 composite by a simple hydrothermal–co-precipitation method and investigated the degradation of norfloxacin with activated peroxodisulfate. These findings demonstrated 100% degradation efficiency of norfloxacin (pH: 3.0) within 13 min due to an excellent synergistic effect at m(NGO–Fe3O4)[thin space (1/6-em)]:[thin space (1/6-em)]m(PDS) of 4[thin space (1/6-em)]:[thin space (1/6-em)]1, and concentrations of NOR and S2O82− of 100 mg L−1 and 1 mM, respectively. According to this, in situ generated ·OH was considered to be the main active free radical. rGO-coupled manganese oxynitride,397 immobilized Ag3PO4/GO on 3D nickel foam398 and γ-Fe2O3-MIL-53(Fe)–GO399 photocatalysts also displayed efficient degradation of norfloxacin.

3.5.6 Heterojunction, Z- and S-scheme-based composites. Ni-doped ZnO/MWCNTs were tested for complete degradation of norfloxacin corresponding to initial concentrations in mg L−1 (time in min) of 10 (30), 20 (60), 50 (120), 100 (160) and 10 (40), 20 (70), 50 (150), 100 (200) under visible and UV radiation, respectively.400 The findings also suggested that MWCNTs can act as a charge transfer channel for accelerating electron transfer between Ni and ZnO nanoparticles. This could subsequently effectively decrease the recombination of electron–hole pairs in the Ni-doped ZnO/MWCNTs composite, accounting for the degradation of norfloxacin by the Ni-doped ZnO/MWCNTs photocatalyst. A Bi-containing glass–ceramic defect-rich heterojunction photocatalyst originating from the removal of chloride ions achieved 98%, 73%, and 36% degradation of norfloxacin under UV-vis–NIR, vis–NIR, and NIR irradiation, respectively.401 Guo et al.402 prepared Co3O4/Bi2MoO6 p–n heterostructure photocatalysts via an in situ calcination process and applied them to activate peroxymonosulfate (PMS) in the degradation of norfloxacin under irradiated visible light. These findings indicated 87.68% removal of norfloxacin within 30 min by selecting a 5 wt% Co3O4/Bi2MoO6/PMS photocatalyst owing to the synergistic effect. A CoTiO3/UiO-66-NH2 p–n junction mediated heterogeneous photocatalyst showed 90.13% degradation of norfloxacin in 1 h under optimized conditions and followed a type-II p–n heterojunction charge transfer mechanism.403 An LaOCl/LDH Z-scheme heterojunction catalyst containing oxygen vacancies showed a 82.5% (150 min) removal rate for norfloxacin owing to the synergistic effect of the Z-scheme heterojunction and oxygen vacancies.404 Further, the degradation of norfloxacin followed pseudo-first-order kinetics with the rate constant of LaOCl/LDH twice that of the individual components.

Z-Scheme ternary heterojunctions comprising phosphate-doped BiVO4/graphene quantum dots/P-doped g-C3N4 (BVP/GQDs/PCN) produced an 86.3% degradation rate for norfloxacin under visible light.405 Such an excellent performance of the photocatalyst is guided by interfacial charge transfer efficiency and a broadened visible-light response range compared to binary type-II heterojunction phosphate-doped BiVO4/PCN. CoWO4 nanoparticles assembled with g-C3N4 nanosheets fabricated by a hydrothermal method showed 3.18 and 2.69 times higher photocatalytic degradation of norfloxacin under visible light compared to g-C3N4 and CoWO4, respectively.406 Such enhanced performance of CoWO4/g-C3N4 is attributed to the synergism between CoWO4 and g-C3N4 inhibiting the fast recombination of photogenerated electron–hole pairs. Investigations involving radical scavengers suggested that ·OH rather than O2˙ plays a dominant role in the degradation of norfloxacin. Fig. 16 shows the possible mechanism responsible for the photodegradation of norfloxacin by this synthesized CoWO4/g-C3N4, a phenomenon driven through a Z-scheme mechanistic pathway.


image file: d3lf00142c-f16.tif
Fig. 16 Schematic illustration of possible Z-scheme photocatalytic mechanism. Reproduced from ref. 406 with permission from Elsevier (2019).

A Bi2Sn2O7/heated perylene diimide (PDIH) Z-scheme heterojunction photocatalyst reached 98.71% degradation of norfloxacin in 90 min under visible light.407 The apparent rate constant of norfloxacin was found to be 3.65 and 20 times those of PDIH and Bi2Sn2O7, respectively. The fabricated Bi2Sn2O7/PDIH heterojunction catalyst also facilitated the separation of charge carriers and preserved the redox capability. In another study, piezo-photocatalytic degradation of norfloxacin by the S-scheme heterojunction BaTiO3/TiO2 was found to be 91.7% (60 min) with a rate constant of 43 × 10−3 min−1.408 Free radical trapping investigations indicated h+ and ·OH to be the main active species in the degradation process. The heterojunction also showed excellent stability and cyclability, as evident after 5 cycles. An LaFeO3/g-C3N4 heterojunction showed 95% photocatalytic degradation of norfloxacin under visible light in 180 min, which was found to 9.32 times higher than pristine g-C3N4.409 Zhang et al.410 prepared an optimized AgBr (3%)/LaNiO3 (30%)/g-C3N4 (100%) dual Z-scheme composite system via ultrasound-assisted hydrothermal method considering energy band matching and observed 92% photodegradation of norfloxacin within two hours under visible light owing to a synergistic effect. These studies also indicated an almost unaltered photodegradation rate (>90%) even after six cycles.

Ag3PO4/CNTs exhibited an efficiency of about 93% for the photoelectrocatalytic degradation of NOR within 30 min.411 This is explained based on the Z-scheme mechanism that significantly promoted the separation of electron–hole pairs. Further, h+ and ·O2 made a major contribution to the degradation process to oxidize NOR. An oxygen-vacancy-rich CuWO4/BiOCl composite exhibited excellent photocatalytic degradation of norfloxacin (96.69%) in 120 min under a 300 W xenon lamp due to a Z-scheme structure compared with pure CuWO4 and oxygen-vacancy-rich BiOCl.412 A dual Z-scheme mechanism has been proposed for Ag (0.3 wt%)@BiPO4/BiOBr/BiFeO3 that enabled 98.1% and 99.1% degradation of norfloxacin (20 mg L−1) in 90 min and in less than 45 min under visible and UV light exposure, respectively.413 It is suggested that the synergistic effects of ternary nanoheterostructures heterojunctions, electron capture and the surface plasmon resonance effect of Ag lead to such high photocatalytic activity. Immobilized Z-scheme CdS/Au/TiO2 nanobelts displayed 64.67% (60 min) degradation of norfloxacin under xenon-light-simulated sunlight irradiation which was ascribed to the synergistic effect.414

The formation of an S-scheme in the heterojunction of a photocatalyst facilitates the separation of photogenerated electron–hole pairs and reduces the recombination of charge carriers. In view of this, an S-scheme heterojunction comprising N–ZnO/g-C3N4 prepared by calcining ZIF-L/g-C3N4 in a mass ratio of 15% showed more than 90% degradation of norfloxacin in 90 min under a visible system.415 The corresponding rate constant was 4.15 times and 4.65 times higher than g-C3N4 and N–ZnO, respectively. The effective light capture capacity and migration and separation of carriers accounted for such behavior. Further, holes and superoxide radicals are reported to be the active species in the photodegradation of norfloxacin. The degradation rate of norfloxacin on a 10% g-C3N4/Bi8(CrO4)O11 heterojunction photocatalyst is about 1.38 and 2.33 times higher than that of pure Bi8(CrO4)O11 and g-C3N4, respectively.416

Efficient photocatalytic performance for norfloxacin degradation has also been reported in chitosan/TiO2@g-C3N4,417 AgI/MFeO3/g-C3N4 (M: Y, Gd, La),418 Bi2Sn2O7/g-C3N4,419 Ag/graphitic carbon nitride quantum dots (CNQDs)/g-C3N4,420 BiOBr/iron oxides,421 and CdS QDs/CaFe2O4@ZnFe2O4422 photocatalysts.

Table 6 records the performance data of different photocatalysts on the removal of norfloxacin from wastewater.

Table 6 Performance data on removal of norfloxacin in water using various photocatalysts
Photocatalyst Preparation NOR Catalyst dose pH Light type Degradation (time) Rate constant
TiO2−x361 Combustion method 100 μM L−1 0.1 g L−1 7 Xenon lamp: 300 W (>400 nm) ∼100% (240 min) 0.0361 min−1
Cu2O363 Hydrothermal 20 mg L−1, (50 mL) 50 mg Xenon lamp (500 W) 79.87% (210 min) 0.0081 min−1
Bi2WO6 with [Fe3+]: 0.3 mmol L−1364 Ultrasonic spray pyrolysis 0.0313 mM L−1 (100 mL) 0.5 g L−1 9 Xenon lamp: 300 W 89.7% (20 min) 0.1006 min−1
TiO2/Ti film with exposed {001} facets (HF: 0.02 M)365 Hydrothermal 10 mg L−1 2.62 Low-pressure mercury lamp (10 W), λmax: 254 nm 70.5% (90 min) 0.0504 min−1
ZnO nanoflowers366 Sol–gel method 10 mg L−1 0.1 g L−1 11 Fluorescent lamp: 8 W (0.55 mW cm−2) ∼72% (100 min) 3.93 × 10−2 min−1
Triangular Ag nanoplates coated ZnO nanoflowers366 Sol–gel method 10 mg L−1 1.0 g L−1 11 Fluorescent lamp (8 W), 0.55 mW cm−2 ∼97% (100 min) 3.93 × 10−2 min−1
Triangular Ag nanoplates coated ZnO nanoflowers367 Hydrothermal method and dual-reduction method 10 ppm (3 mL) Fluorescent lamp (8 W), 0.55 mW cm−2 92.2% (270 min) 9.2 × 10−3 min−1
Prussian blue doped CeO2 (ratio: 10%) with H2O2: 9 mM368 Physical and chemical loading approaches 16 mg L−1 (50 mL) 0.6 g L−1 6 W fluorescent lamp (0.55 mW cm−2) 88.93% (30 min)
N doped TiO2370 Hydrothermal method 6.03 mg L−1 0.54 g L−1 6.37 Xenon lamp (300 W), 350–780 nm, 150 mW cm−2 99.53% (30 min)
Al (1 Mol%)-doped TiO2 nanoflakes371 Solvothermal 2 × 10−4 M 15 mg (50 ml) 10.1 Visible light 93% (120 min) 0.0143 min−1
C–TiO2372 Solution phase carbonization method 0.0094 mM 0.2 g L−1 Neutral Low-pressure mercury lamps (420 nm) ∼100% (70 min) 5.44 × 10−4 [NFX]0-1 + 0.10 [C–TiO2] − 1.99 × 10−2 min−1
Bi3+ and Fe2+ doped ZnO374 Sol–gel method 10.0 mg L−1 1.0 g L−1 8 Xenon lamp (300 W), 45.2 mW cm−2 80% (120 min)
Bi3+ and Fe2+ doped ZnO (0.2 mM HSO5)374 Sol–gel method 10.0 mg L−1 1.0 g L−1 8 Xenon lamp (300 W), 45.2 mW cm−2 99% (120 min) 9.8 × 109 M−1 s−1 (·OH), 9.0 × 109 M−1 s−1 (SO4·)
[FeII(dpbpy)(phen)2]/TiO2379 Hydrothermal 0.313 mM 1 g L−1 5 Xenon lamp (300 W), λ > 420 nm, 140 mW cm−2 98.5% (180 min) 0.0412 min−1
Ag2O/TiO2–zeolite380 Sol–gel method 5 mg L−1 (100 mL) 50 mg Xenon lamp (35 W), 6.7 mW cm−2 98.7% (60 min)
FeVO4/Fe2TiO5 (2[thin space (1/6-em)]:[thin space (1/6-em)]1)381 One-pot hydrothermal method 10 mg L−1 (50 mL) 0.05 g 500 W Xe lamp 95% (30 min)
Ag/AgCl–CeO2 (Ag mass ratio: 13.94 wt%)383 Via urea hydrolysis and calcination 10 mg L−1 (50 mL) 30 mg Xe lamp: 300 W (equipped with a UV cut-off filter) 91% (90 min) 0.02279 min−1
ZnO/ZnS@biochar (ZnSO4/poplar sawdust ratio: 1[thin space (1/6-em)]:[thin space (1/6-em)]1)384 Impregnation-roasting method 0.025 g L−1 (50 mL) 0.5 g L−1 7 UV-light 95% (180 min) 0.021 min−1
Ag/AgCl/Ag2MoO4387 In situ photoreduction 10 mg L−1 (50 mL) 30 mg Xenon lamp: 300 W, (λ > 420 nm) ∼65% (90 min)
ZnO/g-C3N4–Fe3O4390 Hydrothermal 8.61 mg L−1 1.43 g L−1 7.12 Xenon lamp with 280 nm UV filter >90% (120) min 0.0117 min−1
CeO2/g-C3N4 (mass ratio of CeO2 to g-C3N4:5 and PS: 5 mM)391 Mixing method 10 mg L−1 (50 mL) 0.05 g 2 150 W high-pressure xenon lamp with cut-off λ of 420 nm 88.6% (60 min) 0.03573 min−1
NiWO4 nanorods/g-C3N4392 Hydrothermal followed by sonication 10 mg L−1 50 mg (100 mL) W lamp (visible light), 150 mW cm−2 97% (60 min) 0.0547 min−1
rGO/Bi2WO6394 Hydrothermal 10 mg mL−1 (100 mL) 50 mg Xenon lamp (300 W) 87.79% (180 min)
N–TiO2/graphene395 Three-step method 30 mg L−1 (20 mL) Mercury lamp (250 W), 365 nm 50% (160 min) 0.0051 min−1
N-doped rGO/Fe3O4 [m(N–GO–Fe3O4)[thin space (1/6-em)]:[thin space (1/6-em)]m(peroxodisulfate) = 4[thin space (1/6-em)]:[thin space (1/6-em)]1]396 Hydrothermal-co-precipitation 100 mg L−1, S2O82−: 1 mM 1 g L−1 3 UV lamp: 15 W, 254 nm, 44 μW cm−2 100% (13 min) 0.238 min−1
Ni foam supported Ag3PO4/GO (16.78 wt%)398 Dip-coating 15 mg L−1 (120 mL) Xenon lamp (250 W) with 400 nm cut-off filter, 100 mW cm−2 83.68% (100 min) 0.426 min−1
γ-Fe2O3-MIL-53(Fe)–GO399 Multiple steps 10 mg L−1 20 mg 500 W Xe lamp (100 mW cm−2), (420 nm cut-off filter) 92.8% (90 min)
Ni-doped ZnO/MWCNTs400 Dispersion method 100 mgL−1 (100 mL) 6.8 UV 100% (200 min)
Visible 100% (160 min)
Bi contained glass–ceramic401 Multiple steps 20 mg L−1 (20 mL) 20 mg UV-vis–NIR ∼53% (180 min) 6.76 × 10−3 min−1
Bi contained glass–ceramic401 Multiple steps 20 mg L−1 (20 mL) 20 mg Visible ∼35% (180 min) 2.52 × 10−3 min−1
Bi contained glass–ceramic401 Multiple steps 20 mg L−1 (20 mL) 20 mg UV ∼52% (180 min) 4.05 × 10−3 min−1
LaOCl/LDH404 Precipitation method 10 mg L−1 (50 mL) 20 mg 7 Xenon lamp: 300 W 85% (80 min) 0.014 min−1
Phosphate-doped BiVO4/graphene quantum dots/P-doped g-C3N4405 Hydrothermal 20 mg L−1 (50 mL) 50 mg 9.6 Xenon lamp (300 W) with a 420 nm cut-off filter 86.3% (120 min) 0.0148 min−1
CoWO4/g-C3N4406 Hydrothermal method, followed by ultrasonication 10 mg L−1 (100 mL) 50 mg 250 W halogen lamps (visible light) 91% (80 min) 0.0283 s−1
LaFeOx/g-C3N4409 Ultrasound assisted hydrothermal method 20 mg (100 mL) 20 mg L−1 Xenon lamp with 420 nm cut-off filter 95% (180 min) 0.01371 min−1
3 wt% AgBr/30 wt% LaNiO3/100% g-C3N4410 Ultrasound-assisted hydrothermal method 20 mg L−1 (100 mL) 20 mg 7 Xenon lamp (500 W) with a 420 nm cut-off filter 92% (120 min) 0.01790 min−1
0.3 wt% ag@BiPO4/BiOBr/BiFeO3413 Precipitation-wet impregnation-photo deposition method 20 mg L−1 0.3 g 7.3 Visible 98.1% (90 min) 0.04123 min−1
0.3 wt% ag@BiPO4/BiOBr/BiFeO3413 Precipitation-wet impregnation-photo deposition method 20 mg L−1 0.3 g 7.3 UV 99.1% (45 min) 0.07023 min−1
Immobilized CdS/au/TiO2414 Multiple steps 5 mg L−1 (35 mL) 4 cm3 Xenon lamp (35 W) 64.67% (60 min) 0.018 min−1
AgI/LaFeO3/g-C3N4418 Ultrasound-assisted hydrothermal approach 20 mg L−1 (100 ml) 0.2 g Xenon lamp (500 W), (40 mW cm−2) 95% (180 min) 0.0188 min−1
20% Bi2Sn2O7/g-C3N4419 Ultrasound-assisted hydrothermal method 20 mg L−1 (100 mL) 0.02 g 500 W xenon lamp with a UV cut-off filter 94% (180 min) 0.01261 min−1
BiOBr/iron oxides421 In situ co-precipitation method 10 mg L−1 (50 mL) 0.5 g ∼7 800 W xenon lamp with 420-nm cut-off filter 99.8% (90 min) ∼0.076 min−1


3.6 Ciprofloxacin

Ciprofloxacin (CIP) is a synthetic antimicrobial agent of the fluoroquinolone class and considered to be a very promising and efficacious drug for use in the treatment of various community-acquired and nosocomial infections.360,423,424 It is not easily biodegradable and is considered a potential risk to human health. The presence of ciprofloxacin in water acts as pollutant and can be removed by means of a photocatalytic approach.425–524
3.6.1 Metal oxides.
3.6.1.1 TiO2. The photocatalytic degradation of ciprofloxacin as a micropollutant in water has been receiving considerable attention in the presence of metal oxides. Zeng et al.424 used carbon-dot-doped TiO2 to investigate the kinetics, mechanism and pathway following heterogeneous photocatalytic ozonation degradation of ciprofloxacin. It was noted that 1.0 wt% introduction of carbon dots enhanced the degradation of CIP by 91.1% compared to pristine TiO2 (64%) in 30 min. Several studies have been made on ciprofloxacin degradation using commercial TiO2 as a photocatalyst irradiated with simulated solar light,425,426 artificial sunlight,426 simulated sunlight427 and UVA/LED428 and UVC radiation.429 TiO2 nanoparticles irradiated with UVA light demonstrated removal of ciprofloxacin (300 μg L−1) from water in less than 6 minutes.430 The hydrothermally synthesized mesoporous TiO2 exhibited 96% photocatalytic degradation of ciprofloxacin hydrochloride (CIP·HCl) under artificial sunlight compared to that prepared by calcination of a titanium glycolate precursor and subsequent hydrothermal-calcination.431 This is ascribed to the higher electron–hole separation and charge transfer capability.

Li et al.432 fabricated 3D tripyramid TiO2 (TP-TiO2) architectures and rod-like morphology of TiO2 (RL-TiO2) and studied their application in the photocatalytic degradation of ciprofloxacin hydrochloride under UV-vis-light irradiation. They observed relatively superior removal efficiency (90% within 60 min) for ciprofloxacin and its significantly higher rate constants in the presence of TP-TiO2 compared to RL-TiO2. This is ascribed to the key role played by superoxide radicals and photogenic holes in the degradation of ciprofloxacin. Usman et al.433 used TiO2 nanoparticles (50 mg) in the ∼91% degradation of ciprofloxacin aqueous solution (pH: 5.5) on irradiation by a white mercury UV lamp for 5 hours.


3.6.1.2 ZnO and other oxides. ZnO (125 nm) is found to be a very effective photocatalyst in removing 300 μg L−1 ciprofloxacin from aqueous solution treated by UVA in less than 6 minutes.430 ZnO nanoparticles prepared by a chemical precipitation method on irradiation with UV light (365 nm) for 60 min degraded ciprofloxacin (∼48%) in aqueous solution (pH: 10) and also followed pseudo-first-order kinetics (∼0.00437 min−1).434 ZnO nanoparticles synthesized by a sol–gel method were used to examine the degradation of ciprofloxacin in contaminated water under UVC light.435 These findings showed complete photodegradation in 140 minutes corresponding to an initial concentration of ciprofloxacin of 10 mg L−1, pH 5, ZnO loading of 0.15 g L−1 and irradiation time of 140 min. According to Ulyankina et al.,436 UVA-irradiated ZnO nanoparticles synthesized by a pulse alternating current electrochemical method reached 93.6% removal efficiency in 30 min under optimal conditions (initial CIP concentration: 5 mg L−1, pH: 6.5, catalyst dosage: 0.5 g L−1, UV light intensity: 2.0 mW cm−2). Such performance of ZnO nanoparticles is attributed to their higher surface area and increased charge carrier separation compared to commercial ZnO. In another study, ZnO nanoparticles prepared by chemical precipitation immobilized on a glass plate showed 69.5% degradation efficiency for an aqueous solution (pH: 6.8) of ciprofloxacin (10 mg L−1) under UVC irradiation (180 min).437 A ZnO nanostructure prepared by a pyrolysis method achieved 95.5% ciprofloxacin degradation in 60 min under visible light.438

A ZnO nanotube photocatalyst on irradiation with the terrestrial solar spectrum showed about 2.9 times faster degradation of ciprofloxacin compared to TiO2 Degussa P25.439 The flower-like ZnO architectures assembled with nanorods displayed 96% efficiency (240 min) for the degradation of ciprofloxacin (initial conc.: 0.015 μM) in aqueous solution under a UV lamp as a light source.440 Finčur et al.441 undertook comparative studies by examining the photocatalytic properties of TiO2, ZnO and MgO nanopowders prepared by a sol–gel method in the removal of ciprofloxacin from water under UV/simulated sunlight. The corresponding efficiencies of 93.4%, 86.9% and 59.6% suggested TiO2 to be most efficient nanopowder for this. The photocatalytic activity of CdO nanoparticles synthesized via a green route imparted 95% degradation of ciprofloxacin in aqueous media under sunlight (60 minutes).442 In another work, ZnO nanorod irradiated with UV lamp recorded 92% degradation of ciprofloxacin in 60 minutes.443

3.6.2 Metal–metal oxides. A photocatalyst of mesoporous TiO2 modified with Fe (1.5%) and N (2.5%) degraded nearly 70% of ciprofloxacin under visible light in 6 h.444 Ag (0.5 to 4%) nanoparticles grown on the surface of TiO2 exhibited highly enhanced degradation of ciprofloxacin under solar light at low pH.445 A mechanism has also been proposed based on the formation of intermediates identified during the oxidation of ciprofloxacin. A simple reduction method has been used to prepare Cu@TiO2 hybrids of varying Cu/TiO2 wt. ratios (0.1–50) and their photocatalytic performance was examined for ciprofloxacin hydrochloride under sunlight simulated by a 500 W xenon lamp.446 These findings revealed its complete removal in 3 h, corresponding to a Cu/TiO2 wt. ratio of 0.1 in Cu@TiO2 due to the best charge separation and transfer efficiency of photogenerated electrons and holes compared to pure TiO2.

TiO2 modified with monometallic and bimetallic nanoparticles comprising 1.5%-Au/TiO2, 1.5%-Ag/TiO2, 1.0%-Cu/TiO2, 1%Au–0.5%Ag/TiO2 and 1.0%Au–0.5% Cu/TiO2 were fabricated by a deposition–precipitation method and used as photocatalysts in the degradation of ciprofloxacin in pure water under UVC-light irradiation.447 These investigations revealed 100% degradation of ciprofloxacin for all these modified TiO2 catalysts corresponding to 60, 30, 60, 90 and 45 min, respectively. This is ascribed to the lower recombination of the hole–electron pairs arising from the electron trap effect by metal nanoparticles.

3.6.3 Doped metal oxides. The removal of ciprofloxacin from water has been studied in the presence of metals, nonmetals and conducting polymers as dopants in metal-oxide-based photocatalysts. Suwannaruang et al.448 used a hydrothermal method to synthesize nitrogen (12.5%) doped TiO2 particles by selecting urea as a source of nitrogen. Subsequent investigation of its photocatalytic activity showed maximum degradation of ciprofloxacin (94.29%) after 4 h of UV-light irradiation. This is attributed to the integration of nitrogen into the TiO2 lattice and the increased formation of OH radicals. Nitrogen-doped TiO2 (N/Ti wt. ratio: 0.34%) prepared by a sol–gel method and immobilization on glass spheres resulted in 93.5% removal of ciprofloxacin in 90 min under visible-light irradiation.449 The photodegradation of ciprofloxacin followed first-order-kinetics and the photocatalyst exhibited excellent stability even after 5 cycles. Visible-light-irradiated P-doped TiO2 with surface oxygen vacancies (SOVs) exhibited 100% degradation efficiency for ciprofloxacin.450 This is explained on the basis of the synergistic effect as a result of P doping and SOVs on TiO2 significantly enhancing the transfer and separation efficiency of photogenerated charge carriers. Polyaniline (PANI)-doped ZrO2 on UV-light irradiation showed 96.6% photodegradation of ciprofloxacin under optimum conditions (PANI/ZrO2: 30 mg, ciprofloxacin conc: 4 × 10−5 M) in 120 min.451

A ZnO-modified g-C3N4 photocatalyst removed 93.8% ciprofloxacin from water, corresponding to an amount of 0.05 g L−1 and pH value of 8.452 Further studies have shown the degradation rate of ciprofloxacin by ZnO-doped g-C3N4 to be 4.9 times faster than that of undoped g-C3N4. The photocatalyst also exhibited high reusability, as evident from 89.8% efficiency after 3 cycles. Boron-doped TiO2 and cerium-doped TiO2 demonstrated about 90–93% photocatalytic degradation of ciprofloxacin and norfloxacin under solar light.373 Such enhanced photocatalytic activity was explained on the basis of the narrowed band gap and electron–hole separation. In addition, metal-doped metal oxides, such as Fe0/TiO2,453 Fe-doped ZnO454 Zn-doped Cu2O,455 and Cu-doped ZnO,456 have also been successfully reported in the photodegradation of ciprofloxacin.

Several investigations have also been reported on co-doped metal oxides for their applications as photocatalysts in the removal of ciprofloxacin from water. According to Nguyen and others,457 the UV-visible-light-driven photocatalytic degradation of ciprofloxacin hydrochloride (30 mg L−1) by N,S-co-doped TiO2 exhibited a removal efficiency of 78.7% at pH 5.5 for a catalyst dose of 0.05 g. The synthesized N,C-co-doped TiO2 under optimum conditions demonstrated the highest photocatalytic activity in the removal of ciprofloxacin in water under visible light.458 It was concluded that photogenerated holes and superoxide radicals play an active role in the degradation of ciprofloxacin. ZnO nanowires doped with copper and cerium oxides displayed 88.9% removal of ciprofloxacin under UV irradiation.459

3.6.4 Metal oxide composites. In recent years, several studies have been reported on the photodegradation of ciprofloxacin using a variety of composite materials.460–471 A graphitized mesoporous carbon–TiO2 nanocomposite facilitated an almost complete photocatalytic performance in the degradation of ciprofloxacin under UV irradiation.460 A Co/Mn oxide photocatalyst (1.00 g L−1) prepared by a sol–gel method displayed maximum discoloration (56.3%) of ciprofloxacin (10.00 mg L−1) in water (pH: 4) at about 120 min under sunlight.461 TiOF2/TiO2 prepared at 160 °C under hydrothermal conditions exhibited 95.3% degradation of ciprofloxacin hydrochloride under simulated solar light after 90 min.462 In all likelihood, such a combination of TiO2 and TiOF2 composites generates more charge carriers, including an improvement in the transmission and separation efficiency of photogenerated electron–hole pairs. TiO2/Montmorillonite,463 3D γ-Fe2O3@ZnO core–shell464 and rGO–BiVO4–ZnO465 photocatalysts have also shown enhanced degradation of ciprofloxacin.

Teixeira et al.466 made an assessment of the optimization and reusability of Fe3O4/SiO2/TiO2 magnetic photocatalytic particles in the degradation of ciprofloxacin. These studies have shown 95% degradation of ciprofloxacin (pH: 5.5) after 90 min under UV with no significant loss even after five uses. Ternary core–shell Fe3O4/SiO2/TiO2 nanocomposite photocatalysts showed good synergistic properties on the removal efficiency for ciprofloxacin under UVA-light irradiation.467 The photocatalytic degradation of ciprofloxacin hydrochloride by Ag–SrTiO3/TiO2 composite nanostructures under simulated sunlight resulted in 97.6% degradation of ciprofloxacin due to an increase in the carriers and separation between electron–hole pairs.468

Metal oxide/hydroxyapatite,469 CuFe2O4@methyl cellulose,470 TiO2-modified Bi2MoO6471 and Ag2O/Ag2CO3/MWNTs472 have also been examined successfully as composite photocatalysts for the enhancement of ciprofloxacin degradation in water under UV, UVC and visible light, respectively.

3.6.5 Carbonaceous-material-based composites.
3.6.5.1 g-C3N4 and carbon-dot-based composites. Hernández-Uresti et al.333 used polymeric g-C3N4 powder and observed 60% degradation of ciprofloxacin in aqueous solution (pH: 5.5) after 240 min under UV-vis irradiation. Recent studies on exfoliated g-C3N4 (2 g L−1) showed 78% degradation of ciprofloxacin (20 ppm) irradiated under solar light for 1 h.473 In another finding, a 3D g-C3N4/TiO2/kaolinite heterogeneous composite displayed ∼92% degradation efficiency for ciprofloxacin in 240 min under visible-light irradiation.474 This is ascribed to the larger surface area and the availability of more reactive sites, and the efficient separation and longer lifetimes of photogenerated electron–hole pairs. Chuaicham et al.475 observed 98% decomposition of ciprofloxacin (10 mg L−1) within 120 min after irradiation with visible light of a Zn–Cr layered double oxide/fly ash composite photocatalyst in aqueous conditions. The formation of new electronic levels accounted for such enhanced photocatalytic performance. In situ synthesized 3D g-C3N4/La–N–TiO2 also showed complete degradation of ciprofloxacin (5 mg L−1 starting concentration) at a pH of about 6.5 in about 60 min under exposure to simulated solar light.476 Carbon dots/Bi4O5Br2477 nanocomposites also displayed improved visible-light photocatalytic degradation of ciprofloxacin.
3.6.5.2 Composites of graphene oxide and graphene. Graphene oxide and reduced graphene have been used to fabricate binary and ternary composites and they have been used as photocatalysts in the removal of ciprofloxacin from water. Sponza et al.478 prepared nano-GO–Fe3O4 nanocomposites by adding water-dispersed Fe3O4 nanoparticles to an aqueous solution of GO. This irradiated with sunlight produced 80% removal efficiency for ciprofloxacin in water under optimum conditions (initial conc. of ciprofloxacin: 1 mg L−1, original pH: 6.5, nano-GO/M concentration: 2 g L−1, irradiation time: 250 min). ZnO-particle-coated carboxyl-enriched GO (ZnO@cGO) degraded almost 100% ciprofloxacin in water (pH: 7) within about 5 min under visible irradiation (initial concentration of CIP: 25 μg mL−1, catalyst: 0.5 mg mL−1).479 It was concluded that degradation of ciprofloxacin depends mainly on O2 and h+. An rGO-supported BiVO4/TiO2 heterostructure nanocomposite achieved 80.5% degradation rate for ciprofloxacin in acidic ambient (pH: 5) within 150 min, 2.06 times higher than BiVO4/TiO2.480 A nanostructured ZnO–CdO incorporated rGO photocatalyst showed degradation of ciprofloxacin of around 99.28% in 75 min under UV light.481 This is attributed to the effective separation of charge carriers consequential on the production of more reactive oxygen species after incorporation of rGO nanosheets with ZnO–CdO.

The performance of ZnAl mixed metal-oxide (MMO)/rGOx (x: wt% of rGO) composites was tested and compared with ZnAl MMO and pure ZnAl MMO in the photodegradation of ciprofloxacin hydrochloride in aqueous solution under visible light.482 It was found to show the following order of photodegradation efficiency at the end of 2 h of irradiation time: ZnAl MMO/rGO20 (∼90.58%) > ZnAl LDH/rGO20 (∼67.74)% > ZnAl MMO (50.96%) > ZnAl LDH (36.47%). Such enhanced performance of the ZnAl MMO/rGO20 photocatalyst has been ascribed to the synergistic effect of the heterogeneous structure. The degradation mechanism of ciprofloxacin has been clearly explained based on the heterostructure that accounts for efficient charge separation and inhibition of the recombination of photogenerated carriers. It is believed that O2· radicals and h+ predominantly contribute to the degradation of ciprofloxacin. TiO2 (64.3 wt%)-pillared multilayer graphene nanocomposites showed better photodegradation efficiency of 78% than TiO2 (42%) under light-emitting diode irradiation for 150 min.483 The photodegradation followed pseudo-first-order kinetics with the rate constant of graphene/TiO2 composite about 3.89 times that of pristine TiO2. The graphene/TiO2 composite also exhibited high stability and reusability even after five consecutive photocatalytic cycles. Urus et al.484 used a GO@Fe3O4@TiO2-type core@shell@shell nanohybrid (10 mg) as a catalyst to remove 91.5% of ciprofloxacin (10 ppm) from water solution (pH: 7) after 240 min. In addition, the photocatalytic removal of ciprofloxacin has also been evaluated using 3D-structured flower-like bismuth tungstate/magnetic graphene nanoplates485 and Ag2CrO4/Ag/BiFeO3@rGO photocatalysts.486

Huo et al.487 synthesized an N-doped ZnO/CdS/graphene oxide ternary composite via a two-step method and tested its photocatalytic activity in the degradation of ciprofloxacin hydrochloride under visible light and compared it with pure CdS, N–ZnO, and N–ZnO[thin space (1/6-em)]:[thin space (1/6-em)]CdS (2[thin space (1/6-em)]:[thin space (1/6-em)]1, 1[thin space (1/6-em)]:[thin space (1/6-em)]1, 1[thin space (1/6-em)]:[thin space (1/6-em)]2, 1[thin space (1/6-em)]:[thin space (1/6-em)]3). The highest degradation rate of about 86% was shown for the 2[thin space (1/6-em)]:[thin space (1/6-em)]1 molar ratio of N–ZnO and CdS. This is explained in terms of heterostructure and the contribution from GO in N–ZnO/CdS promoting photogenerated electron transfer and suppressing the recombination of electron–hole pairs. The proposed schematic suggested that charge transfer and holes played a major role in the photocatalytic system.

3.6.6 Heterostructures, heterojunctions and Z-scheme-based photocatalysts. An Ag3PO4/TiO2 heterojunction has been fabricated following the corn-silk-templated synthesis of TiO2 nanotube arrays with Ag3PO4 nanoparticles.488 Its application as a photocatalyst in the removal of ciprofloxacin showed degradation efficiency of 85.3% within 60 minutes under simulated solar-light irradiation. Deng et al.489 observed 92.6% removal efficiency for ciprofloxacin by Ag-modified P-doped g-C3N4/BiVO4 nanocomposites under visible-light irradiation (>420 nm). It was suggested that a synergistic effect could account for such improvements as a result of reduced electron–hole recombination. ZnO–Ag2O/porous g-C3N4 ternary composites achieved 97.4% degradation efficiency for ciprofloxacin compared to ZnO (8.2%), g-C3N4 (25.4%), Ag2O (42.3%), and ZnO–Ag2O (69.4%) within 48 min under visible-light irradiation.490

Magnetic g-C3N4/MnFe2O4/graphene composites have been examined for the photocatalytic degradation of ciprofloxacin in the presence of persulfate as an oxidant under visible-light irradiation.491 Graphene-layer-anchored TiO2/g-C3N4 showed enhanced photocatalytic performance (degradation rate: 61.7%, k: 0.01675 min−1) under visible light compared to graphene-layer-anchored TiO2, g-C3N4 and g-C3N4/TiO2.492 This is explained on the basis of accumulation of g-C3N4 electrons with high reduction capability and TiO2 holes with high oxidation capability. Enhanced photocatalytic activity has also been displayed by a visible-light-driven mesoporous TiO2@g-C3N4 hollow core@shell heterojunction in the degradation of ciprofloxacin.493

A heterostructure comprising Ag nanoparticles deposited on the surface of ZnO nanoplates and Fe2O3 nanorods exhibited superior solar-light-driven photocatalytic activity in ciprofloxacin degradation (76.4%) under optimized conditions (initial ciprofloxacin concentration: 10 mg L−1; pH 4; catalyst loading: 0.3 g L−1).494 The e, h+, ·OH and ·O2 played important roles as reactive species in the photocatalytic degradation process. The efficient separation of charge carriers and migration of e/h+ across the heterostructure interface accounted for this. Zhao et al.495 achieved 95.6% removal of ciprofloxacin under visible-light irradiation for 40 min by a ternary Mn2O3/Mn3O4/MnO2 (molar ratio of 3[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]2) valence state heterojunction with dual heterostructures under visible light. Such a performance is derived from its enhanced surface area, light absorption and charge separation of the Mn2O3/Mn3O4/MnO2 heterostructure. Further studies established that holes and superoxide radicals play an important role in the degradation of ciprofloxacin. Other studies comprising a unique 2D/3D/2D rGO (3%)/Fe2O3 (4%)/g-C3N4 heterojunction showed almost 100% degradation of ciprofloxacin (pH: 7) compared to pristine g-C3N4 nanosheets under visible-light irradiation for 40 minutes.496 Such photocatalytic properties of a heterojunction nanocomposite system are accounted for in terms of enhanced charge migration and separation.

Chen et al.497 noted the enhanced degradation of ciprofloxacin over Bi2O3/(BiO)2CO3 heterojunctions compared to pristine (BiO)2CO3 and Bi2O3 in the presence of simulated solar light. The decay process for ciprofloxacin followed pseudo-first-order kinetics with the rate constant increasing with decreasing concentration of CIP. In addition, CdS/BiOBr,498 Cu2O/Cu2(PO4)(OH),499 Sm-doped g-C3N4/Ti3C2MXene,500 CeO2/La2O3/TiO2,501 g-C3N4/NH2-MIL-88B(Fe)502 and a polypyrrole-sensitized ZnFe2O4/g-C3N4 n–n heterojunction503 have also displayed enhanced photocatalytic degradation of ciprofloxacin.

Costa et al.504 observed ∼98% photodegradation of ciprofloxacin (initial concentration: 5 ppm) at neutral pH in the presence of a Z-scheme TiO2/SnO2 nanostructure photocatalyst. These findings also revealed the active role of oxygen singlets, holes, and superoxide radicals as the main species in the photodegradation of ciprofloxacin. Li et al.505 prepared an oxygen-vacancy-rich TiO2/Ta3N5 composite by a solvothermal method and used it as a direct Z-scheme heterojunction photocatalyst. They observed 95.7% (90 min) degradation rate of ciprofloxacin hydrochloride under visible-light irradiation. It was suggested that oxygen vacancies form an intermediate energy level in TiO2 that accounts for the separation of photogenerated electrons and holes. In addition, the formation of a Z-scheme energy band structure by oxygen-vacancy-rich TiO2 and Ta3N5 is likely to enable more photogenerated carriers to participate in the photocatalytic reaction. This was also inevitable from the excellent photocatalytic degradation of ciprofloxacin delivered by an oxygen-vacancy-rich TiO2/Ta3N5 composite under visible light. CeO2/ZnO nanocomposites prepared by a co-precipitation method displayed twice the activity in the photocatalytic degradation of ciprofloxacin compared to undoped ZnO and was ten times more active than pristine CeO2.506 Such enhanced formation of a Z-scheme heterojunction is attributed to the migration of photo-excited electrons from the conduction band of ZnO to the valence band of CeO2.

N-doped carbon quantum dot (NCQD)-decorated Bi2O2CO3 heterojunction nanosheets exhibited remarkably enhanced photocatalytic activities for ciprofloxacin photodegradation mediated by radiation in the ultraviolet to near-infrared region.507 It is suggested that NCQDs act as photosensitizers (hole reservoirs) to harvest solar light and a type-II heterojunction facilitates efficient charge carrier separation to account for this. The mechanisms and pathways of ciprofloxacin degradation mediated by different lights were also discussed. N-doped carbon dots (NCDs) decorated onto a Bi2MoO6/g-C3N4 (BMCN) nanocomposite photodegraded ciprofloxacin by 98% (30 min) under visible-light irradiation.508 It is proposed that NCDs play a role as a mediator to transfer electrons from the conduction band to the valence band of Bi2MoO6 and g-C3N4, respectively. The findings also revealed ·OH and ·O2 radicals acting as the dominant reactive species. The photocatalyst also displayed good stability and reusability after five consecutive cycles of ciprofloxacin photodegradation.

A Z-scheme involving a TiO2 nanorod/g-C3N4 (30 wt%) nanosheet nanocomposite showed 93.4% degradation of ciprofloxacin (initial concentration: 15 mmol L−1) aqueous solution (pH: 6.3) under simulated sunlight irradiation in 60 min.509 It was also concluded that h+ and ·OH played a major role in the degradation of ciprofloxacin. In another study, a biochar@ZnFe2O4/BiOBr Z-scheme heterojunction photocatalyst prepared by a solvothermal method under visible-light irradiation (λ > 420 nm) showed no significant degradation efficiency for ciprofloxacin (65.26%).510 Wen et al.511 fabricated CeO2–Ag/AgBr composite photocatalysts with a Z-scheme configuration by following the in situ interspersal of AgBr on CeO2 and subsequently studied the photodegradation of ciprofloxacin under visible-light irradiation (Fig. 17(a)). According to this, CeO2 itself has almost no ability to degrade ciprofloxacin, though it can be partly eliminated in the presence of pristine Ag/AgBr. However, CIP concentration decreased further to some extent for CeO2 decorated with Ag/AgBr in CeO2–Ag/AgBr composites with 21.26 wt% of Ag (denoted CAB-21.26) exhibiting the most pronounced photocatalytic activity. This is ascribed to the accelerated interfacial charge transfer process and the improved separation of the photogenerated electron–hole pairs. Furthermore, the kinetic behavior followed pseudo-first-order kinetics and exhibited higher k-values for the CeO2–Ag/AgBr hybrids (Fig. 17(b)). Another Z-scheme-based AgBr/Ag/Bi2WO6 heterostructure achieved 57% (5 h) photocatalytic degradation of ciprofloxacin under visible-light irradiation in pure water.512 Such a performance was ascribed to the synergistic effect of the AgBr/Ag/Bi2WO6 heterostructure compared to its single components.


image file: d3lf00142c-f17.tif
Fig. 17 (a) Photocatalytic degradation CIP curves and (b) apparent rate constants for the degradation of CIP solution for a CAB-21.26 sample. Reproduced from ref. 511 with permission from Elsevier (2018).

Z-Scheme-guided g-C3N4/Bi2WO6,513 Fe3O4/Bi2WO6,514 g-C3N4/Ti3C2/MXene/black phosphorus,515 g-C3N4/Ag3PO4/chitosan,516 Ag/AgVO3/g-C3N4,517 CeO2/Co3O4 p–n hetrojunctions,518 Bi nanodots/2D Bi3NbO7 nanosheets,519 Bi2WO6/Ta3N5,520 g-C3N4@Cs0.33WO3,521 ZnO/SnS2,522 g-C3N4/rGO/WO3,523 and CuS/BiVO4524 have also displayed enhanced photocatalytic degradation of ciprofloxacin.

Table 7 records the performance data of different photocatalysts on the removal of norfloxacin from wastewater.

Table 7 Data on performance data on removal of ciprofloxacin using different photocatalysts
Photocatalyst Preparative method CIPa/CIP·HClb Catalyst dose pH Light source Degradation (time) Rate constant
P25 TiO2 (anatase[thin space (1/6-em)]:[thin space (1/6-em)]rutile = 80[thin space (1/6-em)]:[thin space (1/6-em)]20), [H2O2]: 82.5 mg L−1[thin space (1/6-em)]425 Commercial 0.030 mmol L−1[thin space (1/6-em)]a (500 mL) 0.5 g L−1 6 Simulated solar irradiation, 800 W xenon lamp ∼100% (90 min) 0.022 min−1
Degussa P-25 TiO2 (80:20% w/w anatase-to-rutile)426 Commercial 100 mg L−1[thin space (1/6-em)]b 1 g L−1 9 Simulated solar irradiation (850 W cm−2) ∼100% (160 min) 0.108 min−1
Degussa P-25 TiO2 (80[thin space (1/6-em)]:[thin space (1/6-em)]20% w/w anatase-to-rutile)428 Commercial 20 mg L−1[thin space (1/6-em)]a (100 mL) 100 mg L−1 6.0 UVA/LED lamp (3 W), 10 mW cm−2, λ > 365 nm 0.2217 ± 0.0179 min−1
TiO2 (80% anatase and 20% rutile) immobilized on glass plates429 Multiple steps 60 μmol L−1[thin space (1/6-em)]a (500 mL) TiO2 (7.5 g L−1) 9 UVC lamp: 15 W 254 nm ∼98% (120 min) ∼25 × 10−3 min−1
TiO2 P25 and ZnO430 Commercial 300 μg L−1[thin space (1/6-em)]a (50 mL) 1 g L−1 UVA (1.6 to 1.7 mW cm−2) 100% (6 min)
Mesoporous TiO2 nanoparticles431 Hydrothermal 160 mg L−1[thin space (1/6-em)]b (40 ml) 0.01 g Xenon lamp (500 W), 200–1000 nm 96.05% (360 min) 0.45 min−1
3D tripyramid TiO2 architectures432 Hydrothermal method 32.6 μMa (50 mL) 5 mg UV-vis light 90% (60 min) 4.03 × 10−2 min−1
ZnO nanoparticles434 Chemical precipitation method 4 mg L−1[thin space (1/6-em)]a (3 mL) 20 mg L−1 10 Xenon lamp (365 nm) ∼48% (60 min) 0.0043 ± 0.003 min−1
ZnO nanoparticles435 Sol–gel method 10 mg L−1[thin space (1/6-em)]a 0.15 g L−1 5 Low-pressure mercury-vapour lamps (9 W) 100% (140 min) 0.032 min−1
Nano-ZnO436 Pulse electrochemical synthesis 5 mg L−1[thin space (1/6-em)]a 0.5 g L−1 6.5 UV light (2.0 mW cm−2) 93.6% (30 min)
Immobilized ZnO nanoparticles437 Heat attachment method 10 mg L−1[thin space (1/6-em)]a 14 × 14 × 5 cm3 6.8 UV lamp (15 W, 42 W m−2) 69.5% (180 min) ∼0.008 min−1
ZnO nanotubes439 Modified published protocol 2 × 10−5 mol L−1[thin space (1/6-em)]a (0.4 L) 14 mg 8.0 300 W xenon lamp with AM1.5 filter (1000 W m−2) 12% (120 min) 9.61 × 10−4 min−1
Flower-like ZnO440 Thermionic vacuum arc 0.015 μMa ZnO deposited on 2 × 2 cm2 (Si wafer) UV lamp, 1 W m−2, 253.7 nm 96% (240 min) 14.8 × 10−3 min−1
TiO2441 (NH4)2S2O8: 0.125 mM Sol–gel method 0.05 mMa 0.5 mg mL−1 High-pressure Hg lamp (125 W), 1.4 × 10−2 W cm−2 in UV region 93.4% (60 min)
ZnO441 Sol–gel method 0.05 mMa 0.5 g L−1 High-pressure Hg lamp (125 W) in UV region, 1.4 × 10−2 W cm−2 86.9% (60 min)
CdO442 Green approach 10 ppma (50 mL) 50 mg Sunlight 95% (60 min) 0.04722 min−1
ZnO–Ag-Graphite443 Hydrothermal method 5 mg L−1[thin space (1/6-em)]a (50 mL) 0.3 g L−1 24 W UV lamp, λ: 254 nm 98% (60 min) 0.05983 min−1
2.5% N-1.5% Fe–TiO2444 Hydrothermal method 20 mg L−1[thin space (1/6-em)]a (300 mL) 0.3 g LED illumination source 70% (360 min) 5.52 × 10−3 min−1
Ag nanoparticles@TiO2445 Sonicating TiO2 and aq. AgNO3 + aq. Na2CO3 1.0 mMa (100 mL) 1.0 mg L−1 7 UV light (120 W Hg lamp) 85.21% (14[thin space (1/6-em)]500 s) 1.53 mM s−1
Ag nanoparticles@TiO2445 Sonicating TiO2 and aq. AgNO3 + aq. Na2CO3 1.0 mMa (100 mL) 1.0 mg L−1 7 Sunlight 75.58% (14[thin space (1/6-em)]500 s) 1.210 mM s−1
Mesoporous Cu (0.1 wt%)@TiO2446 Reduction method 40 mg L−1[thin space (1/6-em)]b (40 mL) 0.01 g 500 W xenon lamp (sunlight) ∼100% (3 h) 1.16 h−1
1.5%-Au/TiO2447 Deposition–precipitation method 30 mg L−1[thin space (1/6-em)]a (250 mL) 0.5 g L−1 UVC light irradiation (15 W low-pressure Hg lamp, 254 nm 44 W m−2) 100% (60 min) 0.06 min−1
1.5%-Ag/TiO2447 Deposition–precipitation method 30 mg L−1[thin space (1/6-em)]a (250 mL) 0.5 g L−1 UVC light irradiation (15 W low-pressure Hg lamp, 254 nm 44 W m−2) 100% (30 min) 0.117 min−1
1.0%-Cu/TiO2447 Deposition–precipitation method 30 mg L−1[thin space (1/6-em)]a (250 mL) 0.5 g L−1 UVC light irradiation (15 W low-pressure Hg lamp, 254 nm 44 W m−2) 100% (60 min) 0.072 min−1
1% Au–0.5% Ag/TiO2447 Deposition–precipitation method 30 mg L−1[thin space (1/6-em)]a (250 mL) 0.5 g L−1 UVC light irradiation (15 W low-pressure Hg lamp, 254 nm 44 W m−2) 100% (90 min) 0.053 min−1
1.0% Au–0.5% Cu/TiO2447 Deposition–precipitation method 30 mg L−1[thin space (1/6-em)]a (250 mL) 0.5 g L−1 UVC light irradiation (15 W low-pressure Hg lamp, 254 nm 44 W m−2) 100% (45 min) 0.099 min−1
N (12.9%) doped–TiO2 nanorice particles448 Hydrothermal method 20 ppma 0.3 g L−1 5.5 UVA lamps: 20 W, 365 nm, 0.493 mW cm−2 94.29% (240 min)
N doped–TiO2 (N/Ti wt ratio:0.34%) immobilized on glass spheres449 Sol–gel method followed by immobilization 20 mg L−1[thin space (1/6-em)]a (20 mL) 3 g L−1 Xenon lamp: 500 W and λ > 420 nm 93.5% (90 min) 0.02859 min−1
P-doped TiO2 (using 50 mg NaH2PO2)450 Heat treatment under flowing NH3 5 ppma (50 mL) 25 mg Visible-light irradiation 100% (60 min) 0.065 min−1
Polyaniline doped ZrO2451 In situ oxi. Polym. 4 × 10−5 Ma (100 mL) 30 mg UV-light irradiation (λ > 400 nm) 96.6% (120 min)
TiO2/Fe0453 Liquid-phase reduction process 30 mg L−1[thin space (1/6-em)]a 1.0 g L−1 3.0 UV-lamp: 10 W, 254 nm, 2.0 W m−2 94.6% (60 min)
Fe doped ZnO nanoparticles454 Precipitation route 5 mg L−1[thin space (1/6-em)]b 150 mg L−1 9 Sunlight, 650 W m−2, 80[thin space (1/6-em)]000 ± 3000 lux ∼80% (210 min)
Zn-doped Cu2O (by adding 0.05 g of ZnCl2)455 Solvothermal method 20 mg L−1[thin space (1/6-em)]a (50 mL) 30 mg 500 W metal halide lamp, λ < 400 nm filter 94.6% (240 min) 0.0038 min−1
N-S-doped TiO2457 Sol–gel method 30 ppma 0.05 mg 5.5 Halogen lamp: 500 W (360–780 nm) 78.7% (220 min) 0.0065 min−1
Graphitized mesoporous carbon–TiO2460 Extended resorcinol-formaldehyde method 15 mg L−1[thin space (1/6-em)]a (200 mL) 70 mg 14 W UV lamp, 254 nm 100% (120 min) 0.102 min−1
Mo/co oxides461 Sol–gel method 10 mg L−1[thin space (1/6-em)]a 1 g L−1 4 Sunlight 56.3% (180 min) 7.9 × 10−2 min−1
TiOF2/TiO2462 Hydrothermal (160 °C) 20 mg L−1[thin space (1/6-em)]b (50 mL) 50 mg Xenon lamp: 300 W with a UV-cut-off filter (420 nm) ∼95% (90 min) 0.034 min−1
Core–shell 3D γ-Fe2O3@ZnO464 Hydrothermal-sintering and atomic layer deposition 10 mg L−1[thin space (1/6-em)]a (100 mL) 0.5 g L−1 5.8 Xenon lamp (300 W) 92.5% (60 min) 0.0419 min−1
rGO–BiVO4–ZnO465 Hydrothermal method 4 × 10−5 Ma (100 mL) 30 mg W lamp (150 mW cm−2), (λ < 400 nm) 98.4% (60 min)
Fe3O4/SiO2/TiO466 Sol–gel synthesis (calcined at 600 °C) 5 mg L−1[thin space (1/6-em)]a 1 g L−1 5.5 UV irradiation, (365 nm, 1.6 mW cm−2) 95% (90 min) 0.032 min−1
Core–shell Fe3O4/SiO2/TiO2(100 °C)467 Microwave-assisted synthesis 10 mg dm−3[thin space (1/6-em)]a (100 cm3) 50 mg 6.5 UVA lamp (365 nm) 94.0% (120 min) 0.0158 min−1
Ag–SrTiO3/TiO2468 Hydrothermal/photoreduction 20 mg L−1[thin space (1/6-em)]b (50 mL) 20 mg 300 W xenon lamp 97.6% (60 min) 0.070 min−1
TiO2/hap (with 40% by wt% of oxide:Hap)469 Soft chemical method 20 ppma (100 mL) 2 g L−1 HPK 125 W lamp- UV light 100% (15 min)
ZnO/HAp (with 40% by wt% of oxide:Hap)469 Soft chemical method 20 mg L−1[thin space (1/6-em)]a 2 g L−1 HPK 125 W lamp-UV light 100% (20 min)
CuFe2O4@methyl cellulose470 Microwave-assisted method 3 mg L−1[thin space (1/6-em)]a 0.2 g 7 UVC lamps (low pressure, 6 W, Philips) 72.87% (90 min): real sample 0.902 min−1
TiO2/Bi2MoO6 (TiO2 content: 0.41 wt%)471 Solvothermal–calcination process 10 mg L−1[thin space (1/6-em)]a (50 mL) 30 mg Xenon lamp 350 W with a UV cut-off filter 88% (150 min) ∼8 × 10−3 min−1
Ag2O/Ag2CO3/MWNTs472 Calcination (10 min) 10 mg L−1[thin space (1/6-em)]a (100 mL) 0.05 g Xenon lamp: 300 W (visible light) 76% (60 min)
g-C3N4333 Polycondensation of melamine 10 mg L−1[thin space (1/6-em)]a 200 mg (200 mL) Xenon lamp (35 W): UV-vis radiation source 60% (240 min) 4 × 10−5 s−1
Exfoliated g-C3N4473 Green route 20 ppma 1 gL−1 Solar-light irradiation 78% (60 min) 23 × 10−3 min−1
g-C3N4/TiO2/kaolinite474 Sol–gel method/chemical stripping/self-assembly 10 ppma (100 mL) 0.2 g Xenon lamp (90 mW cm−2 with 400 nm cut-off filter) ∼92% (240 min) 0.00813 min−1
Zn–Cr LDH/fly ash (molar ratio = 2[thin space (1/6-em)]:[thin space (1/6-em)]1)475 Coprecipitation method followed by dispersion method 10 ppma (50 mL) 1.0 g L−1 Xenon lamp (500 W) with UV cut-off filter ∼98% (150 min)
g-C3N4/La–N–TiO2476 In situ synthetic method 10 mg L−1[thin space (1/6-em)]a 0.75 g L−1 ∼6.5 Xenon lamp; (300 W), λ > 420 nm 96.8% (60 min)
Nano graphene oxide–magnetite478 Mixing and dispersion 1 mg L−1[thin space (1/6-em)]a 2 g L−1 6.5 Sunlight irradiation at 80 W power 80% (250 min)
ZnO–CdO/rGO481 Refluxing method 10 mg L−1[thin space (1/6-em)]a (50 mL) 10 mg 7 UV light, 800 W xenon lamp with 420-nm cut-off filter 99.28% (75 min)
ZnAl mixed metal oxides/rGO482 Hydrothermal combined with calcination 10 mg L−1[thin space (1/6-em)]a (50 ml) 10 mg 800 W xenon lamp with 420 nm cut-off 90.58% (120 min) 0.01893 min−1
TiO2 (64.3 wt%)-pillared multilayer graphene (35.7 wt%)483 Hydrothermal 15 mg L−1[thin space (1/6-em)]a (40 mL) 20 mg 5.8 LED lamp (5 W), λ > 420 nm 78% (150 min) 0.99111 min−1
GO@Fe3O4@TiO2484 In situ method 10 ppma (100 mL) 10 mg 7–8 Solar simulator: 300 W 91.5% (240 min) 0.0079 min−1
Ag2CrO4/Ag/BiFeO3@8% wt ratio of rGO486 Dispersion method 10 mg L−1[thin space (1/6-em)]a 0.2 mg mL−1 7 Xenon lamp (300 W) with 400 nm cut-off filter, 450 mW cm−2 96% (60 min) 0.0638 min−1
N–ZnO/CdS/GO487 Hydrothermal 15 mg L−1[thin space (1/6-em)]a (100 mL) 50 mg Xenon lamp (300 W) with λ > 420 nm 86% (60 min)
0.6Ag3PO4/TiO2 nanotube arrays (600 °C)488 In situ growth method 10 mg L−1[thin space (1/6-em)]a (40 mL) 40 mg Xenon lamp (300 W), 200 mW cm−2 85.3% (60 min) 0.02499 min−1
P-doped ultrathin g-C3N4/BiVO4489 Impregnated process 10 mg L−1[thin space (1/6-em)]a 1 g L−1 6.72 Visible-light irradiation (λ > 420 nm) 92.6% (120 min) 0.0203 min−1
ZnO–Ag2O/porous g-C3N4490 Hydrothermal 20 mg L−1[thin space (1/6-em)]a (100 mL) 50 mg W lamp (500 W), λ ≥ 420 nm 97.4% (48 min) 0.057 min−1
Graphene layers anchored TiO2/g-C3N4492 In situ calcination method using 40 g of Ti3C2 3 mg L−1[thin space (1/6-em)]a (100 mL) 60 mg Xenon lamp (300 W), λ > 400 nm, 300 mW cm−2 61.7% (60 min) 0.01675 min−1
Ag/Fe2O3/ZnO494 Ultrasonic-assisted hydrothermal method 10 mg L−1[thin space (1/6-em)]a (100 mL) 0.3 g L−1 4 Solar illumination 76.4% (210 min) 0.3036 h−1
Mn2O3/Mn3O4/MnO2495 Hydrothermal and in situ method 10 mg L−1[thin space (1/6-em)]a (120 mL) 0.2 g L−1 7 Xenon lamp (300 W), 900 mW cm−2 95.6% (40 min)
rGO/Fe2O3/g-C3N4496 Embedding approach 50 mg L−1[thin space (1/6-em)]a 100 mg 7 Halogen lamp: 500 W ∼100% (40 min) 1.0878 min−1
Bi2O3/(BiO)2CO3497 Hydrothermal/calcination 10 mg L−1[thin space (1/6-em)]a 0.5 g L−1 (100 mL) 7 Xenon lamp: 300 W, 0.641 W cm−2 93.4% (30 min) 0.476 min−1
CdS/BiOBr-1[thin space (1/6-em)]:[thin space (1/6-em)]3498 Solvothermal route 10 mg L−1[thin space (1/6-em)]a (200 mL) 50 mg 7 Sunlight 99.1% (240 min) 0.00692 min−1
Cu2O/Cu2(PO4)(OH)499 Reflex method 20 mg L−1[thin space (1/6-em)]a (100 mL) 100 mg Direct sunlight irradiation ∼98% (120 min)
CeO2/La2O3/TiO2501 Sol–gel followed by calcination 10 ppma (50 mL) 50 mg 6–7 Visible light using tungsten lamp (300 W cm−2) 100% (120 min)
TiO2/SnO2504 Hydrothermal and ion exchange 2.5 × 10−3 g L−1[thin space (1/6-em)]a 2.5 × 10−3 g Neutral UVC lamps with 35 W each (253 nm) 92.8% (120 min) 22.4 × 10−3 min−1
CeO2/ZnO506 Co-precipitation method 15 mg L−1[thin space (1/6-em)]a (100 mL) 0.25 g L−1 3.2 200 W mercury-xenon lamp with 365 nm filter ∼60% (60 min) 0.0130 min−1
5 wt% N-doped carbon quantum dots decorated Bi2O2CO3507 Hydrothermal method 10 mg L−1[thin space (1/6-em)]a (80 mL) 40 mg UV-vis light 91.1% (60 min) ∼0.0325 min−1
Visible 92.8% (60 min) ∼0.02 min−1
Bi2MoO6/g-C3N4508 Hydrothermal method 5 mg L−1[thin space (1/6-em)]a 1.0 g L−1 8 Visible lamps (77 mW cm−2) 98% (30 min) 0.12 min−1
TiO2 nanorod/30 wt% g-C3N4 nanosheets509 Mixing followed by ultrasonication 15 μmol L−1[thin space (1/6-em)]a (50 mL) 10 mg 6.3 Xenon lamp: 500 W 93.4% (60 min) 0.0381 min−1
5 wt% biochar@ZnFe2O4/BiOBr510 Solvothermal/photodeposition/precipitation 15 mg L−1[thin space (1/6-em)]a 50 mg (100 mL) Xenon lamp: 300 W 65.26% (60 min)
CeO2-21.26 wt% Ag/AgBr511 In situ 10 mg L−1[thin space (1/6-em)]a (50 mL) 50 mg Xenon lamp (300 W) with a UV cut-off filter 93.05% (120 min) 0.02011 min−1
Bi2WO6/ag/AgBr512 Precipitation followed by dispersion 30 mg L−1[thin space (1/6-em)]a (250 mL) 125 mg Phillips lamp (50 W), λ = 380–800 nm 57% (5 h)
g-C3N4/Bi2WO6513 Solvothermal and grind calcination method 15 mg L−1[thin space (1/6-em)]a (100 mL) 0.1 g Xenon lamp:300 W, λ < 400 nm 98% (120 min)
Fe3O4/Bi2WO6 (4% iron content)514 Hydrothermal method 10 mg L−1[thin space (1/6-em)]a (100 mL) 30 mg Visible-light irradiation (λ > 420 nm) ∼99.7% (15 min)
g-C3N4/Ti3C2 MXene/black P515 Calcination process 20 mg L−1[thin space (1/6-em)]a (100 mL) 20 mg Xenon lamp: 300 W, λ > 420 nm >99% (60 min) 0.048 min−1
g-C3N4/Ag3PO4/chitosan516 Multiple steps 20 mg L−1[thin space (1/6-em)]a 2.0 mg 7 Visible light 90.34% (60 min) 0.01771 min−1
0.5 wt% Ag/AgVO3/g-C3N4517 Wet-impregnation method 10 ppma (100 mL) 0.1 g Halogen lamp (500 W): visible light 82.6% (120 min)
Bi (7%) nanodots/Bi3NbO7 nanosheets519 Two-step wet chemical reaction 10 mg L−1[thin space (1/6-em)]a (100 mL) 50 mg Xenon lamp: 300 W with 400 nm cut-off filter 86% (120 min) 0.01427 min−1
Bi2WO6/Ta3N5 (1.0/1 mole ratio)520 Electrospinning–calcination–solvothermal route 20 mg L−1[thin space (1/6-em)]a (100 mL) 40 mg 3 Xenon lamp: 300 W with a cut-off filter (λ > 400 nm), 97 mW cm−2 81.1% (120 min) 0.0105 min−1
g-C3N4@Cs0.33WO3521 Solvothermal 20 ppm 20 mg (100 mL) 3 Xenon lamp (500 W), λ: 230–2500 nm, 0.25 W cm−2 97% (145 min) 14.9 × 10−3 min−1
g-C3N4/rGO/WO3523 Photo reduction method 20 mg L−1[thin space (1/6-em)]a (50 mL) 10 mg High-pressure xenon arc lamp with 400 nm cut-off filter and 100 mW cm−2 85% (180 min)
CuS/BiVO4 (mass ratio: 7%)524 In situ 10 mg L−1[thin space (1/6-em)]a (100 mL) 100 mg Xenon lamp (300 W) with a 420 nm cut-off filter 86.7% (90 min) 0.02151 min−1


3.7 Tetracycline

Tetracycline (TC) is invariably used as an antibiotic against different bacterial infections, such as urinary tract infections, acne, gonorrhea, chlamydia, mycoplasma, rickettsia, cholera, brucellosis, plague and syphilis.52 It finds extensive application in the medical field, for veterinary purposes, and as a feed additive in the agricultural sector. However, extensive applications of tetracycline mean its presence in surface water, groundwater, wastewater, domestic wastewater and other source-related environments, causing a serious threat to the environment. Therefore, several approaches have been made to develop a highly efficient approach to remove antibiotics by a photocatalysis approach.525–662
3.7.1 Metal oxides.
3.7.1.1 TiO2. Several investigations have been reported using TiO2 as a photocatalyst in water treatment for the removal of tetracycline. According to Palominos et al.,527 an aqueous suspension of TiO2 has been used to facilitate the photocatalytic oxidation of tetracycline on irradiation with simulated solar light. Studies indicated the rapid degradation of tetracycline, undergoing 100% completion after 15 min under optimum conditions (tetracycline: 20 mg L−1, TiO2: 1.5 g L−1, pH: 8.7). The mechanism of photocatalytic tetracycline oxidation involved active roles for holes and OH radicals. The nanosized TiO2 achieved more than 95% removal of tetracycline within 40 min under UV irradiation for a tetracycline concentration of 40 mg L−1 and catalyst dose of 1000 mg L−1.528 Safari et al.529 also used nanosized TiO2 (1.0 g L−1) to study the degradation kinetics of a tetracycline hydrochloride (TC·HCl) aqueous solution (55 mg L−1, pH: 5) under ultraviolet irradiation. They observed 100% degradation after 30 min on adding H2O2 (100 mg L−1) compared to 91.4% degradation after 90 min for TiO2/UV. The photocatalytic degradation of tetracycline over commercial TiO2–P25 showed 94.8% (120 min) removal efficiency under visible light (λ = 700 nm).530 Recently, a crosslinking method has been followed for immobilizing TiO2 (P25) nanoparticles in chitosan film, which showed promising photocatalytic activity in the purification of water containing tetracycline hydrochloride under UV irradiation.531 The stability and reusability of this composite film in four consecutive cycles revealed a significant decrease in removal efficiency after the second run, from 87% to 57%. Tetracycline hydrochloride degradation has also been studied using a green and low-cost approach, involving the preparation of immobilized titania samples by depositing two successive TiO2 layers on two different commercial supports.532
3.7.1.2 ZnO and other oxides. Palominos et al.527 carried out the photocatalytic oxidation of tetracycline in an aqueous suspension containing ZnO and found its performance comparable to TiO2 (∼100% degradation) under simulated solar light. According to the suggested mechanism, the contribution towards photocatalytic tetracycline oxidation on ZnO is mainly guided by hydroxyl radicals. UV-irradiated ZnO/peroxymonosulfate has shown about 95.6% degradation of tetracycline (10 mg L−1, pH: 7) in 90 min compared to UV/ZnO (50.14%), attributed to the formation of SO4·.533 In addition, HSO5 acts as an electron acceptor and inhibits electron–hole pair recombination, thereby allowing the formation of more ·OH radicals. Iron oxide nanoparticles,534 nanospherical α-Fe2O3 supported on 12-tungstosilicic acid,535 SnO2 hollow microspheres,536 polyaniline coated on magnetic MoO3537 and BiFeO3538 have also been studied in the photocatalytic degradation of tetracycline aqueous solutions.
3.7.2 Metal-loaded metal oxides. A solution casting method has been used to fabricate membranes by mixing previously prepared core–shell Au (0.1, 0.3, 0.5 g)–TiO2 nanocomposites and PVDF and they were examined for their performance in the degradation of tetracycline under the influence of visible light.539 It is inferred that an Au (0.3)–TiO2/PVDF nanocomposite enhanced the photocatalytic degradation rate by 75% within 120 min under visible light. These findings clearly ensured first-order kinetics for Au–TiO2/PVDF composites, following the order: Au (0.3)–TiO2/PVDF (0.00599 min−1) > Au (0.1)–TiO2/PVDF (0.00449 min−1) > Au (0.5)–PVDF (0.01212 min−1). Excellent regeneration stability and its easy separation have also been achieved by this method. Gold-containing zinc–titanium oxide films540 and Ag/Bi2O3541 have also been reported in the photocatalytic degradation of tetracycline in aqueous media.

Liu et al.542 studied the photoactivity of an Au–ZnO nanomotor system based on vertically aligned ZnO in the photocatalytic degradation of tetracycline as a function of different photocatalysts, UV light intensity and cycling tests, as presented in Fig. 18(a)–(c), respectively. The findings revealed that the respective degradation rates of tetracycline within 30 min and rate constants corresponding to pseudo-first-order kinetics follow the order: Au–ZnO nanorod motors (Au–ZnO-M): 99.3% > Au–ZnO nanorod array (Au–ZnO-A): 95.5% > ZnO (86.5%), and k ‘Au–ZnO nanorod motors (k (Au–ZnO-M): 0.1451 min−1 > k (Au–ZnO-A): 0.1120 min−1 > k (ZnO): 0.0542 min−1). It was suggested that the Au layer in the Au–ZnO heterojunction nanoarrays acted as an electron reservoir to facilitate charge separation, thereby lowering the possibility of photogenerated carrier recombination. A possible photocatalytic mechanism for the photocatalytic degradation of tetracycline by Au–ZnO nanomotors under UV-light irradiation is displayed in Fig. 18(d). According to this, electrons in Au could react with O2 to form ·O2, accounting for the degradation of tetracycline. In contrast, h+ in the VB of ZnO could directly degrade tetracycline to a stable product.


image file: d3lf00142c-f18.tif
Fig. 18 Photocatalytic degradation of TC. (a) Dynamic curves of different photocatalysts (initial conditions: 40 mg L−1, TC, 0.2 g L−1 photocatalyst, and 350 mW cm−2 UV light). (b) The impact of UV light intensity (initial conditions: 40 mg L−1 TC and 0.2 g L−1 Au–ZnO nanomotors). (c) Cycling tests (initial conditions: 30 mg L−1 TC: 0.2 g L−1, Au–ZnO nanomotors, and 350 mW cm−2 UV light). (d) Proposed photocatalytic mechanism for TC degradation. Reproduced from ref. 542 with permission from RSC (2022).
3.7.3 Doped photocatalysts.
3.7.3.1 Doped TiO2 and ZnO. Several studies have been carried out on the performance of doped TiO2 and ZnO photocatalysts and subsequently used in the removal of tetracycline from water.543–546 Red mud and modified red mud originating from industrial solid waste discharged from the aluminum industry have been investigated as low-cost, effective photocatalysts under irradiated visible light.543 Xu et al.544 developed a C-doped TiO2–polymethylsilsesquioxane (PMSQ) aerogel followed by thermal treatment at 400 °C in air. They used it to achieve 98% removal of tetracycline hydrochloride from aqueous solution in 180 min and ascribed it to enhanced charge separation. In another study, hydrothermally prepared carbon (3 wt%)-doped TiO2 with metal (Ni/Co/Cu) nitrate hydroxide was used as a nanocomposite photocatalyst.545 The photocatalytic activity of this catalyst displayed 97% removal of tetracycline hydrochloride within 60 min. TiO2 doped with acetylene black,546 N-doped TiO2/diatomite,547 P-doped carbon nitride tubes combined with peroxydisulfate (PDS),548 N,S-doped TiO2 and N,S-doped ZnO modified chitosan,549 and C,N,S-tri-doped TiO2550 photocatalysts have also been investigated for the degradation of tetracycline.

Metal-doped photocatalysts have also received a lot of attention for their application in the photocatalytic degradation of tetracycline in aqueous solution. Nb-doped ZnO (Nb[thin space (1/6-em)]:[thin space (1/6-em)]Zn molar ratio: 1[thin space (1/6-em)]:[thin space (1/6-em)]1) showed 93.2% degradation efficiency for tetracycline in 180 min under visible light and also possessed superior recyclability and stability.525 Zhang et al.551 fabricated Ag-doped TiO2 (Ag+[thin space (1/6-em)]:[thin space (1/6-em)]Ti4+ mole ratio: 0, 0.5, 1.0, 2.0, 3.0, and 5.0%) hollow microspheres following an applied hydrothermal process by a template-free method. It was noted that Ag-doped TiO2 (Ag+[thin space (1/6-em)]:[thin space (1/6-em)]Ti4+ mole ratio: 3.0%) exhibited maximum removal of tetracycline hydrochloride following first-order kinetics with OH· and h+ playing an active role. Ce (2%)-doped TiO2/halloysite nanotubes and Ce (2%)–TiO2/halloysite nanotubes enabled about 78% tetracycline removal within 60 min under visible-light irradiation.552 TiO2 composite nanofibers doped with CuO were also studied for the photocatalytic degradation of pharmaceutical wastewater.553 Bembibre et al.554 used Ca-doped ZnO nanoparticles in the removal of tetracycline under a visible-light-driven sonocatalytic process.


3.7.3.2 Doped graphitic materials. Doped graphitic materials have attracted a lot of attention as photocatalysts in the removal of tetracycline from water.555 Nitrogen-self-doped g-C3N4 nanosheets prepared by a combination of N-self-doping and thermal exfoliation showed higher photocatalytic activity for tetracycline degradation than bulk g-C3N4, N-self-doped g-C3N4 or g-C3N4 nanosheets.556 This is attributed to the enlarged visible-light absorption ability, reduced recombination and prolonged lifetime of photogenerated charge carriers. Chen et al.557 reported the removal of tetracycline hydrochloride from wastewater (pH: 5) using an S–g-C3N4/PTFE membrane under irradiated visible light. These findings indicated 98.1% photocatalytic degradation corresponding to an initial concentration of tetracycline hydrochloride of 10 mg L−1, catalyst dosage of 1 g L−1, and S–g-C3N4 loading of 50 mg. Further, the S–g-C3N4/PTFE membrane displayed good recovery performance and photocatalytic stability. Ba (2%)-doped g-C3N4 demonstrated significant influence on the photocatalytic activity owing to its low band gap and the effective separation of photo-induced e–h+.558

Er-doped g-C3N4,559 Cd-doped g-C3N4,560 S-doped carbon quantum dot loaded hollow tubular g-C3N4,561 single-atom Ni,S-co-coped g-C3N4,562 nitrogen defect/boron dopant engineered tubular g-C3N4,563 Ag–g-C3N4,564 Bi-nanoparticle-decorated g-C3N4 nanosheets (10 wt%),565 Co-doped TiO2rGO,566 rGO-doped ZnAlTi-LDH,567 and graphene oxide/magnetite/cerium-doped titania568 photocatalysts also acted as efficient photocatalysts in the degradation of tetracycline.

3.7.4 Metal oxide composites. Several studies have been reported on the evacuation removal of tetracycline from water using a combination of metal oxides. Wang et al.526 observed 81% (10 min) photocatalytic degradation of tetracycline by irradiating a 5% carbon quantum dots/TiO2 composite prepared by a hydrothermal method with visible light. Such performance of the composite is attributed to the improved separation efficiency of photogenerated electrons and holes. According to Liu et al.,569 excellent catalytic performances is observed for 3%-CuOx/γ-Al2O3 in the simultaneous degradation of tetracycline hydrochloride in a wide pH range of 3.10–9.47 under irradiation by a 300 W xenon lamp (190–1100 nm). In another study, a sol–gel-synthesized calcite/TiO2 photocatalyst accounts for 90% tetracycline removal under UV light in aqueous solution (pH: 7) corresponding to 1.5 g L−1 of catalyst and 50 mg L−1 of tetracycline.570 Hunge et al.571 studied the effect of catalyst loading for the MoS2 (20 wt%)/TiO2 composite and solution pH, in the degradation of tetracycline under UV-vis irradiation of composites and observed its superior performance (95%) compared to TiO2 and MoS2. ZnO/γ-Fe2O3 demonstrated an important role in achieving ∼89% degradation efficiency for tetracycline in water under UV-visible light after 150 min.572

The degradation of tetracycline in water has been investigated on TiO2 decorated on magnetically activated carbon as a function of different parameters under ultraviolet and ultrasound irradiation.573 These findings revealed 93% tetracycline removal at the end of 180 min under optimum conditions corresponding to an optimum intensity of 70 W US power, pH 6.0, catalyst loading of 0.4 g L−1, and initial concentration of tetracycline of 30 mg L−1. ZnO rod-activated carbon fiber,574 Fe3O4/FeP,575 spatially confined Fe2O3 in hierarchical SiO2@TiO2 hollow spheres,576 La-enriched titania–zirconia oxide,577 Ni(OH)2-decorated TiO2,578 IO–TiO2–CdS,579 and WO2.72/ZnIn2S4580 have also been demonstrated as efficient photocatalysts for the removal of tetracycline from water.

Wang et al.581 converted harmful algae into bio-nanohybrid materials by immobilizing Microcystis aeruginosa cells onto PAN–TiO2/Ag hybrid nanofibers. They observed about 96% degradation efficiency for tetracycline hydrochloride under visible light compared to PAN/TiO2/Ag nanofiber (77%) and M. aeruginosa (49%) due to a synergistic effect. It is suggested that enhanced degradation in M. aeruginosa/PAN–TiO2/Ag could be caused by algae facilitating the effective separation of photogenerated electron–holes on TiO2. The presence of ZnO, carbonaceous layers and Ag nanoparticles improved the optical absorption property in the Ag/ZnO/C structure, resulting in 95.8% (35 min) and 90.6% (280 min) degradation of tetracycline hydrochloride under UV- and visible-light irradiation, respectively.582 This is ascribed to efficient photogenerated electron separation and transportation and an increase in the active reaction sites. According to Wei et al.,583 an SiO2–TiO2–C (nC[thin space (1/6-em)]:[thin space (1/6-em)]nTi mol ratio: 3.5) aerogel composite displayed 80.01% degradation efficiency for tetracycline hydrochloride within 180 min under visible light and also retained its high stability and reusability. ·O2 and ·OH were considered as the active species responsible for the photocatalytic degradation of tetracycline. In addition, ternary chitosan comprising chitosan–TiO2–ZnO over graphene,584 palygorskite-supported Cu2O/TiO2,585 CuO/Fe2O3,586 ZnO@zeolitic imidazolate,587 and bimetallic oxide/carbon588 have also been tested for the photocatalytic degradation of tetracycline in water.

3.7.5 Graphitic materials.
3.7.5.1 g-C3N4. Insufficient sunlight usage, low surface area and rapid charge recombination of electron and hole pairs are a major hinderance contributing towards the low photocatalytic performance of g-C3N4.65 As a result, several investigations have been made into the photodegradation of tetracycline using g-C3N4 and its composites. Hernández-Uresti et al.333 prepared a graphite-like C3N4 photocatalyst by the polycondensation of a melamine precursor and observed the following trend for the photodegradation of four different pharmaceuticals in aqueous solution under UV-vis irradiation: tetracycline > ciprofloxacin > salicylic acid > ibuprofen. The active species responsible for the degradation of tetracycline were considered to be photogenerated holes, OH radicals and H2O2. Self-assembly-based g-C3N4 nanoflakes showed up to 70% removal efficiency for tetracycline (20 ppm) within 180 min under light irradiation (420 nm).589

Shi et al.590 studied the degradation performance of tetracycline in real water systems by metal-free g-C3N4 microspheres under various conditions through visible-light catalysis and PMS activation synergy. According to this, the rate constant values for the degradation of tetracycline by photocatalysis, Fenton-like catalysis, and photo-Fenton-like catalysis are 0.013, 0.025, and 0.028 min−1, respectively. The observed superior degradation performance of photo-Fenton-like catalysis is attributed to the synergetic effect between PMS activation and photocatalysis. In another study, Wang and others591 used graphitic carbon nitride microspheres and recorded 80.54% degradation efficiency for the removal of tetracycline hydrochloride under visible-light irradiation for 2 h, corresponding to a photocatalyst dose of 1.0 g L−1, initial concentration of tetracycline hydrochloride solution of 10 mg L−1 and initial pH 7. Porous g-C3N4,592 GQDs/g-C3N4,593 S-doped graphitic carbon nitride,594 hexagonal BN/g-C3N4,595 poly-o-phenylenediamine (POPD)/g-C3N4,596 and N–CNT/mesoporous g-C3N4597 photocatalysts have also been evaluated for the removal of tetracycline.

Jiang et al.598 studied the degradation of tetracycline in aqueous solution using P and S doped g-C3N4 under visible light (λ ≥ 420 nm) and showed higher photocatalytic than bare g-C3N4 or single-doped g-C3N4. According to this, P and S doping in g-C3N4 inhibited the recombination of electron–hole pairs and facilitated the efficient separation of photogenerated charges. The h+ and ·O2 were the dominant active species responsible for the degradation of tetracycline. Porous g-C3N4/TiO2 (g-C3N4[thin space (1/6-em)]:[thin space (1/6-em)]TiO2 mass ratio: 12[thin space (1/6-em)]:[thin space (1/6-em)]1) photocatalysts removed 88.43% of tetracycline from aqueous solution under a xenon lamp for 90 min, which was ascribed to the synergistic effect.599 In another study, a ZrO2-embedded MoS2/g-C3N4 nanocomposite exhibited 94.8% tetracycline degradation in aqueous solution in 90 min under visible light owing to the dual charge-transfer channel between the layers of MoS2/g-C3N4 and ZrO2 nanoparticles.600 Poly-N-isopropylacrylamide (PNIPAM)/Fe3O4/g-C3N4 prepared by a hydrothermal method and thermal photoinitiation under visible-light irradiation decomposed tetracycline into harmless small molecules.601 The catalytic activity remain more or less unaltered even after 5 repeated uses and could be easily separated. In addition, CDs/g-C3N4/BiPO4,602 ZnO/N-doped g-C3N4,603 and g-C3N4/H3PW12O40/TiO2604 exhibited enhanced photocatalytic degradation performance for tetracycline under visible light.


3.7.5.2 Graphene. Binary and ternary graphitic composite materials have been reported as photocatalysts in the removal of tetracycline from aqueous solution.605–621 According to Ren et al.,605 a red mud/graphene oxide (mass ratio: 93[thin space (1/6-em)]:[thin space (1/6-em)]7) composite attained the best degradation rate for tetracycline (79.8%) compared to raw red mud under visible light within 80 min owing to its enhanced specific surface area, light absorption and charge separation. Porous hydroxyapatite (Hap) hollow microspheres as a source of cheap and green photocatalysts have been harnessed in fabricating rGO/Hap composites.606 Investigations revealed significantly enhanced photocatalytic activity of rGO (1.5 wt%)/Hap in tetracycline degradation (92.1%, 30 min) under a xenon lamp (300 W) for full-spectrum irradiation. This is explained on the basis of the photogenerated electrons accumulating at rGO (acting as an electron acceptor) that could interact with O2 to form ·O2. In addition, separated positive holes in the VB of porous hollow Hap (acting as an electron donor) microspheres directly participate in the oxidation of tetracycline.

Heteropoly acid (H3PMo8W4O40)/graphene oxide nanocomposites based on UiO-66 have been synthesized following an in situ growth hydrothermal method and tested as photocatalysts in tetracycline photodegradation under visible-light irradiation.607 The photocatalytic degradation efficiency for tetracycline was found to be significantly higher (95%: 120 min) compared to GO or heteropoly acid. An Fe3O4/GO/ZnO magnetic nanocomposite showed 74% (100 min) degradation of tetracycline hydrochloride under visible-light irradiation.608 This is explained on the basis of ZnO and Fe3O4/GO in Fe3O4/GO/ZnO contributing to the generation of the electron–hole pairs under visible light and promoting the transfer of photogenerated electrons, respectively. In another study, graphene quantum dot decorated ZnO–ZnF2O4 nanocage ternary composites, prepared by a one-step deposition method exhibited superior performance in the degradation of tetracycline hydrochloride under visible light compared to ZnO or ZnO–ZnFe2O4.609 According to Chakraborty et al.,610 an rGO–ZnTe (1[thin space (1/6-em)]:[thin space (1/6-em)]1) photocatalyst facilitated the degradation of tetracycline due to a synergistic effect. It is suggested that the 2D wrinkled surface of rGO contributes in minimizing the recombination probabilities of photoinduced electron–hole pairs. N-doped TiO2 nanoparticles deposited on rGO exhibited more pronounced photodegradation activity for tetracycline hydrochloride than pure TiO2 or N-doped TiO2.611 Subsequent studies on the reusability of N-doped TiO2/rGO also established the stability of the composite photocatalyst.

Kumar et al.612 fabricated ZnO quantum dots (1.5 wt%)/rGO by a hydrothermal method and observed 68% removal of tetracycline from wastewater (pH: 5) after 120 min under visible light. Fe3O4/g-C3N4/rGO exhibited 86.7% degradation rate of tetracycline hydrochloride, following pseudo-second-order kinetics.613 The proposed mechanism suggested ·O2 and ·OH radicals as the most reactive species in the photocatalytic degradation of tetracycline. Ghoreishian et al.614 reported sonophotocatalytic degradation of tetracycline using a flower-like rGO/CdWO4 composite under simulated visible-light irradiation. These findings under optical conditions (pH: 5.7, initial concentration of tetracycline: 13.54 mg L−1, catalyst dosage: 0.216 g L−1, time: 60 min) revealed its photocatalytic catalytic activity to be 1.5 and 3 times higher than that of commercial nano-ZnO and TiO2, respectively.

Interfacial growth of a TiO2–rGO composite by the Pickering emulsion approach showed 94% removal efficiency for tetracycline hydrochloride (10 ppm) after 40 min under the visible light.615 Such significant enhancement in the photocatalytic efficiency of TiO2–rGO is ascribed to its 2D sandwich-like structure. Porous hollow hydroxyapatite microspheres decorated with rGO,616 rGO–CdS,617 rGO–CdS/ZnS,72 Ag/TiO2 nanosheets/rGO,618 Ag/TiO2 nanosheets,619 Ag/TiO2 nanosheets–rGO,620 and TiO2/rGO/activated carbon621 have also been harnessed as photocatalysts in the degradation of tetracycline in aqueous solution.

3.7.6 Heterojunction-based photocatalysts. Heterojunction photocatalysis have attracted attention for the degradation/removal of tetracycline in aqueous solution by various heterojunctions.51 In this regard, a core–shell g-C3N4@Co–TiO2 heterostructured nanofibrous membrane exhibited excellent visible-light-driven degradation of tetracycline hydrochloride.622 Huang et al.623 observed 74.7% degradation efficiency for tetracycline hydrochloride within 30 min by a hierarchical Au (2%)–g-C3N4–ZnO heterostructure under xenon lamp irradiation. Mesoporous TiO2-modified ZnO quantum dots (8%) immobilized on linear low-density polyethylene (LLDPE) under fluorescent light irradiation showed 89.5% removal of tetracycline (initial concentration: 40 mg L−1) from water (pH: 9) within 90 min.624 g-C3N4/CuO (7%)625 and ZnO globular (15 wt%)/g-C3N4626 showed 55% and 78.4% degradation of tetracyclines in 60 and 50 min under simulated solar light (λ > 365 nm) and artificial visible sunlight illumination (λ ≥ 400 nm), respectively.

Ti0.7Sn0.3O2/g-C3N4 (mass ratio: 10 wt%) achieved 83% degradation of tetracycline hydrochloride in 40 min under irradiated visible light.627 This is explained on the basis of an S-scheme between Ti0.7Sn0.3O2 and g-C3N4 to increase and transport photogenerated charges. The ultrasonic-assisted precipitation method has been used to fabricate a ZnO (20 wt%)/GO (2 wt%)/Ag3PO4 heterojunction and it has been used as a photocatalyst in the elimination of tetracycline hydrochloride from wastewater.628 These findings showed 96.32% (75 min) degradation under visible light corresponding to initial concentration of tetracycline of 30 mg L−1 and catalyst dose of 1.0 g L−1.

The degradation rate of tetracycline was found to be about 10 times higher in a g-C3N4/C/Fe3O4 ternary nanocomposite compared to its individual or binary components under simulated solar light.629 The degradation process followed a first-order kinetics model with a much higher apparent rate constant for g-C3N4/C/Fe3O4 (0.0063 min−1) compared to g-C3N4 (0.0029 min−1) or carbon (0.0003 min−1). The photoinduced h+ and ·O2 free radicals are suggested to act as the main active components in the degradation. The enhanced activity of g-C3N4/C/Fe3O4 in tetracycline degradation is attributed to heterojunction formation and is due to the effective separation of the photocarriers. In addition, the introduction of C into g-C3N4/C/Fe3O4 facilitates an enhancement of the optical response range and effective electron transfer.

Liao et al.630 examined the utility of a core–shell BiFeO3/TiO2 heterostructure with a p–n heterojunction as a photocatalyst prepared by forming nanospheres of TiO2 on BiFeO3 (nanocubes) in tetracycline degradation under visible-light irradiation. The findings indicated much higher degradation efficiency of BiFeO3/TiO2 (72.2%) compared to BiFeO3 (64.9%) and TiO2 (38.3%) after 180 min of visible illumination. A BiFeO3/TiO2 p–n heterojunction photocatalyst showed superior degradation efficiency for tetracycline due to its enlarged specific surface area and higher sensitivity to visible light, improved separation and transfer efficiency of photoelectron–hole pairs and a synergistic effect. Fiber-shaped Ag2O/Ta3N5 p–n heterojunctions designed as efficient photocatalysts showed enhanced photocatalytic activity with good stability in photocatalytic activity for tetracycline under visible light (λ > 400 nm) due to the synergistic effect.631 It is anticipated that photogenerated holes and superoxide radicals played prominent roles in the photocatalytic process.

Chen and others632 fabricated an α-Bi2O3/g-C3N4 heterostructure modified by plasmonic metallic Bi and oxygen vacancies and observed a remarkably high degradation rate for tetracycline (90.2%) under visible light after 180 min. Such enhancement is attributed to the formation of a p–n junction arising from a combination of n-type (g-C3N4) and p-type (α-Bi2O3) semiconductors, which is beneficial in a ternary photocatalyst. It is suggested that Bi nanoparticles and the presence of oxygen vacancies favor the consumption and separation of the photogenerated electrons and holes in the ternary heterojunction photocatalyst. Several other heterojunction photocatalysts, such as C3N4@MnFe2O4–rGO,491 BiVO4/TiO2/rGO,633 porous g-C3N4/AgBr/rGO,634 C3N4-supported WO3/BiOCl,635 BiOI/exfoliated C3N4,636 CuO@ZnO,637 ZnO/SnO2,638 Cu2O–TiO2,639 MoS2/Ag/g-C3N4,640 g-C3N4/ZrO2−x,641 and needle SnO2 nanoparticles anchored on exfoliated g-C3N4642 have also shown enhancement and stability in the degradation of tetracycline.

N-doped ZnO–MoS2 binary heterojunctions have been fabricated by a hydrothermal method and used to study its photocatalytic activity for the degradation of tetracycline under visible-light irradiation.643Fig. 19(a)–(d) show corresponding findings based on variations in the degradation of TC with time, corresponding ln(C/C0) vs. time plots, a histogram showing a comparative degradation rate (%) of TC under visible light illumination and a bar graph showing the values of rate constants for all the photocatalysts. It should be noted that photocatalytic degradation of tetracycline followed pseudo-first-order kinetics. In addition, fabricated semiconductor heterojunctions demonstrated enhanced performance for the degradation of tetracycline due to the synergistic effect. Furthermore, the enhanced photostability of the photocatalyst over three cycles for a period of 360 min is ascribed to the transfer of holes from the valence band of N-doped ZnO to the valence band of MoS2.


image file: d3lf00142c-f19.tif
Fig. 19 (a) Kinetic curves for the degradation of TC, (b) ln(C/C0) vs. time curve for the degradation of TC, (c) a histogram showing a comparative degradation rate (%) of TC under visible light illumination and (d) a bar graph showing the values of rate constants for all the photocatalysts (N-doped ZnO nanorods loaded 0.2, 0.5, 1, 2 and 3 wt% of with MoS2 nanoflowers (MNF) are referred to as NZM0.2, NZM0.5, NZM1, NZM2, and NZM3, respectively). Reproduced from ref. 643 with permission from RSC (2017).

A novel type-II Bi2W2O9/g-C3N4 heterojunction has been fabricated and studied for its photocatalytic performance in the removal of tetracycline under simulated solar irradiation and it was compared with Bi2W2O9 and g-C3N4, as displayed in Fig. 20(a).72 It is inferred that Bi2W2O9/g-C3N4 yields high photodegradation (∼95%) compared to the degradation observed for pristine g-C3N4 (75%) or Bi2W2O9 (∼60%). This is attributed to the Bi2W2O9 semiconductor acting as a trap for photogenerated holes and electrons. A photocatalytic mechanism has also been proposed for the Bi2W2O9/g-C3N4 system in Fig. 20(b).


image file: d3lf00142c-f20.tif
Fig. 20 (a) Photocatalytic degradation of tetracycline antibiotic (C0 = 10 mg L−1, pH = 4.89) as a function of irradiation time over Bi2W2O9, g-C3N4 and Bi2W2O9/g-C3N4 samples. (b) Proposed photocatalytic mechanism for the Bi2W2O9/g-C3N4 system under solar-like irradiation. Reproduced from ref. 72 with permission of Elsevier (2020).

Z-scheme WO3/g-C3N4 composite hollow microspheres fabricated by an in situ hydrolysis and polymerization process showed an enhanced degradation rate towards tetracycline hydrochloride (82% in 120 min) under visible-light irradiation.644 The enhanced separation of photoinduced electrons and holes and the synergistic effect of g-C3N4 and WO3 are considered to be a few reasons for this. In addition, the presence of hollow cavities could enable trapping of the incident photons and facilitate availability of more electrons and holes in the photocatalytic process. In another study, a Z-scheme mesoporous Sn3O4/g-C3N4 heterostructure exhibited superior photocatalytic performance in degrading tetracycline hydrochloride present in water.645 A possible photocatalytic reaction mechanism has also been examined in detail for this. In another study, BiOI/g-C3N4/CeO2 (3 wt%) photocatalyst possessed the best photocatalytic activity for degradation of tetracycline (91.6%) under visible-light irradiation.646 It is anticipated that CeO2/g-C3N4 and BiOI/g-C3N4 catalysts block the recombination of photoinduced electron–hole pairs through the formation of a heterojunction.

Dai et al.73in situ prepared 3D-20% polyaniline/perylene diimide (PANI/PDI) and found the degradation rate for tetracycline under visible-light irradiation in a static system, by 15.3 times and 17.0 times those of pure PDI and PANI, respectively. The main reactive species in the degradation of tetracycline comprised superoxide radicals, hydrogen peroxide and holes. Fig. 21(a) and (b) schematically show the electron–hole pair separation process and TC degradation mechanism of a 3D 20%-PANI/PDI heterojunction under visible-light irradiation. Scanning electron microscopy images of 3D PANI/PDI in Fig. 21(c and d) indicate a significant decrease in size after the dissolution/assembly process and the PDI are uniformly/orderly dispersed in the 3D network structure of PANI.


image file: d3lf00142c-f21.tif
Fig. 21 (a) Morphological structure of PANI/PDI. (b) Photocatalytic mechanism of PANI/PDI heterojunction photocatalysts under visible-light irradiation: direct Z-scheme heterojunction mechanism. (c and d) Scanning electron microscopy images of 3D PANI/PDI. (Modified) Reproduced from ref. 73 with permission of Elsevier (2020).

In addition, TiO2−x/ultra thin g-C3N4/TiO2−x,647 K-doped g-C3N4/TiO2/CdS,648 γ-Fe2O3 nanospheres anchored on g-C3N4,649 CQDs/g-C3N4,650 Ag3PO4/MIL-88A(Fe),651 BiOBr/MoS2/GO,652 g-C3N4/MnO2/GO,653 BiVO4@polypyrrole/g-C3N4,654 AgI/BiOBr/rGO,655 graphene-bridged Ag3PO4/Ag/BiVO4,656 g-C3N4 nanoparticles/WO3 hollow microspheres,657 CuIn2S2/g-C3N4,658 Ag3PO4/g-C3N4/ZnO,659 g-C3N4 nanosheet/Ag3PO4/α-Bi2O3,660 LaNiO3-modified C3N4661 and ultrafine TiO2 nanoparticle modified g-C3N4662 heterojunction photocatalysts have also been harnessed in the removal of tetracycline in water.

Table 8 records the performance data of different photocatalysts used in the removal of tetracycline from water.

Table 8 Performance data of tetracycline on its removal in water using different photocatalysts
Photocatalyst Preparative method CIPa/CIP·HClb Catalyst dose pH Light source Degradation and time Rate constant
TiO2526 Solvothermal 50 mg L−1[thin space (1/6-em)]a (100 mL) 30 mg 6.0 Visible light LED 50 W ∼25% (12 min) ∼0.03 min−1
TiO2 (P25 Degussa)527 Commercial 20 mg L−1[thin space (1/6-em)]a (100 L) 1.5 g L−1 8.7 Xenon lamp, 250 Wm−2, 300–800 nm 100% (15 min)
ZnO (Sigma Aldrich)527 Comme rcial 20 mg L−1[thin space (1/6-em)]a (100 mL) 1.0 g L−1 11 Xenon lamp, 250 Wm−2, 300–800 nm 100% (15 min)
Nanosized TiO2 (P25) with 70% anatase and 30% rutile528 Commercial 40 mg L−1[thin space (1/6-em)]b (40 ml) 1000 mg L−1 9 Medium-pressure mercury lamp (UV), λ < 290 nm 525 μW cm−2 95% (60 min)
Nanosized TiO2 with 80% anatase and 20% rutile in presence of H2O2 (100 mg L−1)529 Commercial 55 mg L−1[thin space (1/6-em)]b (250 mL) 1 g L−1 5 UV lamp: 18 W, λ: 254 nm, 2500 μW cm−2 100% (30 min) 7.25 × 10−3 min−1
TiO2–P25 (80% anatase and 20% rutile) in presence of H2O2 (100 mg L−1)530 Commercial 10 mg L−1[thin space (1/6-em)]a (100 mL) 0.2 g L−1 Xenon lamp 300 W, 20 mW cm−2, λ: 350 nm 94.8% (120 min) ∼3.8 × 10−2 min−1
TiO2 (P25) immobilized in chitosan531 Dispersion method 30 mg L−1[thin space (1/6-em)]b 0.12 g 4 UV lamp (30 W), λmax: 360 nm 87% (360 min) 0.025 min−1
Nanometric and immobilized TiO2532 Modified sol–gel method (product calcined at 400 °C) 35 ppmb 300 mg Jelosil HG500 UV lamp, 30 mW cm−2 90% (35 min) 56 ± 2 × 10−3
ZnO nanoparticles (peroxy monosulfate:2 mM)533 Biosynthesis 10 mg L−1[thin space (1/6-em)]b 2 g L−1 7.0 Low-pressure UV lamp (6 W), λ: 254–258 nm 95.6% (90 min) 0.018 min−1
Iron oxide nanoparticles534 Co-precipitation 83 μMb (10 mL) 10 mg 7 Hg quartz lamp (280 W), λ: 180 nm to 623 nm 40% (60 min) 0.0092 min−1
Nanospherical α-Fe2O3 supported on 12-tungstosilicic acid (H2O2: 0.1 ppm/250 ml)535 Solid state dispersion 30 ppma 150 ppm 8 Hg lamp (15 W), λ: 254 nm 97.39% (50 min) 0.0098 min−1
SnO2 hollow microspheres (SnCl2·2H2O[thin space (1/6-em)]:[thin space (1/6-em)]Na3C6H5O7·2H2O mole ratio = 1[thin space (1/6-em)]:[thin space (1/6-em)]4)536 Hydrothermal method 50 mg L−1[thin space (1/6-em)]b (40 ml) 50 mg Hg lamp, λ: 365 nm 76% (140 min) 0.00861 min−1
BiFeO3 (in presence of H2O2: 9.8 mM)538 Calcination of gel formed from bi and Fe nitrates at 600 °C 40 mg L−1[thin space (1/6-em)]a 2 g L−1 4 Hg lamp (300 W), λ = 365 nm 100% (210 min) 0.02650 min−1
Nb doped TiO2 (Nb[thin space (1/6-em)]:[thin space (1/6-em)]Zn molar ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]1)525 Green synthesis 150 mg L−1[thin space (1/6-em)]a (100 mL) 0.25 g L−1 7 250 W xenon arc lamp with a 420 nm cut-off filter 91. 5% (180 min) 7.3 × 10−3 min−1
Au–TiO2 (0.3 g)/PVDF539 Three-step synthesis strategy 20 mla solution 0.1 g Xenon lamp (300 W), λ < 420 nm 75% (120 min) 0.01212 min−1
C-doped TiO2 (in PMSQ)544 Multiple steps 10 mg L−1[thin space (1/6-em)]b (50 mL) 0.5 g 7 W halogen lamp (100 W) with filter (λ > 420 nm) 98% (180 min)
TiO2/acetylene black with PS: 3 mM L−1546 Mixing method 30 mg L−1[thin space (1/6-em)]b (100 mL) 0.5 g L−1 4.1 LED lamp (30 W), λ: 400–780 nm, 50 W m−2 93.3% (120 min) 2.2 × 10−2 min−1
N doped TiO2 diatomite547 Mixing followed by calcination 20 mg L−1[thin space (1/6-em)]b 5 g L−1 6 Xenon lamp (150 W), λ < 400 nm 91% (300 min)
P doped carbon nitride tube (peroxydisulfate: 1.0 g L−1)548 Hydrothermal calcination 20 mg L−1[thin space (1/6-em)]a (100 mL) 0.3 g L−1 4.59 Xenon lamp (300 W) with a cut-off filter of λ: 400 nm, 180 mW cm−2 96.4% (60 min) 0.0492 min−1
Chitosan modified N,S-doped TiO2549 Sol–gel-hydrothermal method 10 mg L−1[thin space (1/6-em)]a (100 mL) 0.6 g L−1 8.2 LED lamp: 18 W 91% (20 min) 0.048 min−1
C,N,S-tri-doped TiO2550 Sol–gel method (thiourea-to-Ti molar ratio of 0.05[thin space (1/6-em)]:[thin space (1/6-em)]1 and calcined at 450) 5.0 mg L−1[thin space (1/6-em)]a 0.5 g L−1 9 Solar stimulator equipped with xenon arc lamp (150 W), λ < 420 nm 98% (180 min) 24.6 × 10−3 min−1
Ag-doped TiO2 (Ag+ to Ti4+molar ratio: 3.0%)551 Template-free route (hydrothermal) 30 mg L−1[thin space (1/6-em)]b 100 mg Xenon lamp (300 W) and (λ > 420 nm) ∼88% (30 min) 6.77 × 10−2 min−1
Ce (2%)–TiO2/halloysite nanotubes552 Modified sol–gel method 20 mg L−1[thin space (1/6-em)]a (100 mL) 50 mg Xenon lamp (300 W), λ > 420 nm 78% (60 min)
TiO2 composite nanofibers doped with CuO553 Electrospinning technique 100 ppma 1.0 g L−1 Neutral Xenon lamp (400 W) 71% (60 min)
N self-doped g-C3N4556 Combination of N self-doping and thermal exfoliation process 10 mg L−1[thin space (1/6-em)]a (100 mL) 0.5 g L−1 Xenon lamp (300 W), λ > 350 nm 89.14% (60 min)
S–g-C3N4/PTFE membrane557 Ultrasonic device method 10 mg L−1[thin space (1/6-em)]b 50 mg 5 300 W xenon light irradiation with a 420 nm cut-off filter 98.1% (120 min) 0.03348 min−1
Ba (2%)-doped g-C3N4558 Facial thermal condensation method 20 mg L−1[thin space (1/6-em)]a (50 mL) 50 mg 10 Xenon lamp (150 W) with 400 nm cut-off filter 91.94% (120 min) 0.0175 min−1
Er (0.0035 g)-doped g-C3N4559 Calcination 25 mg L−1[thin space (1/6-em)]a (50 ml) 25 mg 4 Xenon lamp (35 W) ∼90% (90 min) 0.0204 min−1
Cd (4.6 wt%) doped g-C3N4560 Thermal polymerization method 10 mg L−1[thin space (1/6-em)]a 0.8 g L−1 5 Xenon lamp (300 W) with λ > 420 nm 98.1% (60 min)
S-doped CQDs loaded hollow tubular g-C3N4561 Ultrasonic assisted synthesis strategy 20 mg L−1[thin space (1/6-em)]a 1 g L−1 Xenon lamp (300 W), 0.33100 W cm−2 82.67% (60 min) 0.0293 min−1
Ni–S co-coped g-C3N4562 Thermal polymerization followed by calcination 10 ppmb (30 mL) 5 mg 5 Xenon lamp (300 W), 100 W cm−2 91.77% (60 min) 0.031 min−1
Ag doped g-C3N4564 Heating melamine and urea mixture 20 mg L−1[thin space (1/6-em)]a (50 mL) 0.1 g 7 Solar light 96.8% (120 min)
Bi nanoparticle-decorated g-C3N4 nanosheet (10 wt%)565 Ultrasound-assisted electrostatic self-assembly method 10 mg L−1[thin space (1/6-em)]b 40 mg 7 Lamp (300 W), λ: 420–780 nm 90.7% (70 min)
Co (0.20 wt%) doped TiO2/rGO566 One-pot hydrothermal method 30 mg L−1[thin space (1/6-em)]b (100 mL) 100 mg Halogen lamp: 500 W (400 nm cut-off filter) 60% (180 min)
Magnetic graphene oxide–Ce (10% mass ratio) doped titania568 Dispersion method 25 mg L−1[thin space (1/6-em)]a (100 mL) 50 mg Xenon lamp: 300 W, 400 nm cut-off filter 82.92% (180 min) 0.03005 min−1
MoS2 (20 wt%)/TiO2571 Hydrothermal route 10 mg L−1[thin space (1/6-em)]a 100 mg 5.5 Metal halide lamp: 400 W (UV-vis light source) 95% (100 min) 0.0276 min−1
ZnO/γ-Fe2O3572 Microwave-assisted solution method 30 mg L−1[thin space (1/6-em)]a (20 mL) 0.5 mg L−1 6.7 Halogen lamp, 100 mW cm−2 88.52% (150 min) 0.01321 min−1
Magnetic activated C@TiO2573 Impregnation method 10 mg L−1[thin space (1/6-em)]a 0.4 g L−1 6 UVC lamp (40 W), (λ: 254 nm) ∼100% 180 min 0.19 min−1
ZnO rod-activated carbon fiber (ACF)574 Microwave 40 mg L−1[thin space (1/6-em)]a (150 mL) One piece (5.5 cm of ZnO rod-ACF) 8 UV lamp (20 W), λ: 8365 nm >99% (60 min)
Fe3O4/FeP (molar ratios of Fe[thin space (1/6-em)]:[thin space (1/6-em)]P at 1[thin space (1/6-em)]:[thin space (1/6-em)]6)575 Hydrothermal synthesis and partial phosphating annealing method 50 mg L−1[thin space (1/6-em)]b (40 mL) 20 mg Xenon lamp (1000 W) 88% (180 min) 0.00984 min−1
Hierarchical hollow SiO2–Fe2O3@TiO2576 Dispersion/in situ polymerization/sol–gel approach 10 mg L−1[thin space (1/6-em)]a (50 mL) 0.2 mg mL−1 3–7 Simulated solar-light irradiation 100% (140 min)
Fe2O3 in hierarchical SiO2@TiO2 hollow sphere576 Dispersion/in situ polymerization/sol–gel 10 mg L−1[thin space (1/6-em)]a (50 mL) 0.2 mg mL−1 Natural sunlight irradiation 100% (80 min)
La–TiO2–ZrO2577 Sol–gel process 10 mg L−1[thin space (1/6-em)]a (100 mL) 0.35 mg L−1 5 UV lamp 100% (120 min) 0.0359 min−1
Ni(OH)2 decorated rutile TiO2578 Deposition of Ni(OH)2 on hydrothermally prepared TiO2 nanorods using 0.2 M TiCl4 100 mg L−1[thin space (1/6-em)]a (20 mL) 200 mg 200 W Hg xenon lamp, λ < 420 nm (cut-off filter) 76% (150 min) 0.0090 min−1
3D IO–TiO2–CdS579 Hydrothermal synthesis 30 mg L−1[thin space (1/6-em)]b 30 mg Xenon lamp (λ < 420 nm) >99% (20 min)
1D/2D WO2.72/ZnIn2S4580 Hydrothermal reaction 50 mg L−1[thin space (1/6-em)]b (300 mL) 30 mg Xenon lamp: 300 W (λ > 420 nm) 97.3% (60 min)
PAN/TiO2/Ag nanofiber581 Immobilizing M. aeruginosa cells onto PAN/TiO2/Ag 20 mg L−1[thin space (1/6-em)]b (500 mL) 1 g L−1 6 Halogen lamp: 500 W (λ < 420 nm) 96% (240 min) 5.62 × 10−3 min−1
Ag/ZnO/C582 Calcination and photodeposition route 20 mg L−1[thin space (1/6-em)]b (100 mL) 100 mg Xenon lamp: 500 W, λ > 400 nm 81% (280 min)
Ag/ZnO/C582 Calcination and photodeposition route 20 mg L−1[thin space (1/6-em)]b (100 mL) 100 mg UV lamp: 250 W, λmax: 365 nm 95.8% (35 min)
SiO2–TiO2–C (nc[thin space (1/6-em)]:[thin space (1/6-em)]nTi: 3.5)583 Sol–gel method 10 mg L−1[thin space (1/6-em)]b (50 mL) Visible light (λ > 420 nm) 80.31% (180 min) 0.00831 min−1
Chitosan–TiO2–ZnO584 Sol–gel and ultrasound-assisted method 20 mg L−1[thin space (1/6-em)]a 0.5 g L−1 4 UV 97.2% (180 min)
Palygorskite-supported Cu2O/TiO2585 Liquid phase reduction method 30 mg L−1[thin space (1/6-em)]b (50 mL) 1.0 mg 8.7 Xenon lamp: 500 W 88.81% (240 min) 0.0129 min−1
CuO/Fe2O3586 Green synthesis 20 mg L−1[thin space (1/6-em)]a 40 mg 7 UV irradiation 88% (80 min) 0.048 min−1
Zr0.3Ti/C588 Calcination 10 mg L−1[thin space (1/6-em)]a Xenon lamp (300 W) 98% (30 min) 0.84 L Mol−1 min−1
Polymeric g-C3N4333 Polycondensation 20 mg L−1[thin space (1/6-em)]a (200 mL) 200 mg 5.5 Xenon lamp (35 W) 86% (240 min)
g-C3N4 nanoflakes589 Thermal condensation followed by heat treatment 20 ppma LED (6 W), λ: 365 nm 70% (180 min)
Self-assembled g-C3N4 microsphere591 Supramolecular self-assembly with post-heating treatment 10 mg L−1[thin space (1/6-em)]b 1.0 g L−1 7 Xenon lamp: 500 W (λ > 420 nm) 80.54% (120 min)
Porous g-C3N4592 Calcination of bulk g-C3N4 20 mg L−1[thin space (1/6-em)]b (50 mL) 30 mg 9 Xenon lamp (300 W) with UV cut-off filter 420 nm 91.8% (60 min)
0.5 wt% GQDs/g-C3N4593 Electrostatic interaction method 20 mg L−1[thin space (1/6-em)]b 25 mg 300 W xenon arc lamp, λ > 400 nm ∼67% (120 min)
S-doped graphitic carbon nitride594 Thermal induction copolymerization 30 mg L−1[thin space (1/6-em)]a 0.01 g L−1 4 Solar light 93.8% (60 min)
h-BN (2.0 mg)/g-C3N4595 In situ method 10 mg L−1[thin space (1/6-em)]a (100 mL) 1.0 g L−1 Xenon lamp (300 W, λ > 400 nm) 79.7% (60 min) 0.02775 min−1
POPD/g-C3N4596 Suspension polymerization 10 mg L−1[thin space (1/6-em)]b (50 mL) 0.5 g L−1 Xenon lamp (300 W) 86.0% (120 min)
N doped CNT/mpg-C3N4597 Thermal polycondensation 20 mg L−1[thin space (1/6-em)]b 1.0 g L−1 Xenon lamp (300 W) 67.1% (240 min)
P,S-doped g-C3N4 (hexachloro triphosphazene: 50 mg)598 In situ thermal copolymerization 10 mg L−1[thin space (1/6-em)]a 1.0 g L−1 300 W xenon lamp, λ > 420 nm 85.85% (60 min) 0.03823 min−1
Porous g-C3N4/TiO2 nanoparticles599 Reaction carried out under autoclave 10 mg L−1[thin space (1/6-em)]a (70 mL) 70 mg 5 Xenon lamp irradiation 88.43% (90 min)
ZrO2 nanoparticles@MoS2/g-C3N4600 Multiple steps 20 mg L−1[thin space (1/6-em)]a (100 mL) 50 mg 3 Xenon lamp (300 W), λ > 420 nm 94.8% (90 min) 0.0230 min−1
PNIPAM/Fe3O4/g-C3N4601 Thermal photoinitiation technology 20 mg L−1[thin space (1/6-em)]a (100 mL) 0. 1 g Xenon lamp (300 W): visible light ∼78% (120 min)
CDs doped g-C3N4//BiPO4602 Hydrothermal method 10 mg L−1[thin space (1/6-em)]b 1 g L−1 4 Xenon lamp (500 W), under visible light 75.50% (220 min) 0.0005 min−1
20% ZnO/N doped g-C3N4603 Self-assembled method through electrostatic attraction 20 mgL−1[thin space (1/6-em)]b (200 mL) 0.1 mg L−1 Xenon lamp (300 W) under visible light 81.3% (15 min) 0.1016 min−1
Red mud modified with graphene oxide (mass ratio: 93[thin space (1/6-em)]:[thin space (1/6-em)]7)605 Ultrasonic mixing 10 mg L−1[thin space (1/6-em)]a 50 mg 6.9 Xenon lamp (300 W), λ > 420 nm 79.8% (80 min) 0.02011 min−1
rGO (1.5 wt%) hydroxyapatite microsphere606 Hydrothermal method 60 mg L−1[thin space (1/6-em)]a 1.0 g L−1 5 Xenon lamp (300 W) with full spectrum irradiation 92.1% (30 min) 0.1816 min−1
Heteropoly acid/GO/UiO-66607 In situ growth hydrothermal method 20 ppma (50 mL) 0.02 g 7 Hg lamp 500 W (λ > 400 nm) 95% (120 min)
Fe3O4/GO/ZnO608 Dispersion, followed by hydrothermal treatment 50 mg L−1[thin space (1/6-em)]b 1 mg L−1 Under simulated light irradiation (intensity: 1 kW m−2) 74% (100 min) 14 × 10−3 min−1
GQDs (1 mL)/ZnO–ZrFe2O4609 One-step deposition process 20 mg L−1[thin space (1/6-em)]b 20 mg (50 mL) Xenon lamp (500 W) coupled with 420 nm cut-off filter ∼90% (27 min) 0.08809 min
rGO/ZnTe (1[thin space (1/6-em)]:[thin space (1/6-em)]1)610 Single-pot one-step solvothermal process 10 mg L−1[thin space (1/6-em)]a 100 mg (50 mL) Solar simulator (AM 1.5, 100 mW cm−2) ∼70% (40 min) 0.033 min−1
N doped–TiO2/rGO611 Photoreduction method 10 mg L−1[thin space (1/6-em)]b 50 mg Xenon arc lamp (300 W) with 400 nm cut-off filter 98% (60 min) 0.05655 min−1
1.5 w% ZnO quantum dots/rGO612 Precipitation and hydrothermal methods 20 ppma (50 ml) 50 mg L−1 5 Non halogen lamps (24 V, 250 W) 68% (120 min) 0.00961 min−1
Fe3O4/g-C3N4/rGO613 Ultrasonic dispersion 20 mg L−1[thin space (1/6-em)]b (100 ml L−1) 0.1 g 7 Xenon lamp (500 W) 86.7% (60 min) 0.0306 min−1
rGO/CdWO4614 Heating method 13.54 mg L−1[thin space (1/6-em)]a 0.216 g L−1 5.7 Simulated solar light 100% (60 min) 0.0693 min−1
rGO/CdS617 Solvothermal 0.08 mmolea (40 ml) 40 mg Solar light 83.25% (16 min) 0.13 min−1
Ag/TiO2/rGO618 Ultrasonic impregnation assisted photoreduction strategy 20 mg L−1[thin space (1/6-em)]a (50 mL) 1 g L−1 7 Hg lamp (300 W), λ < 400 nm ∼100% (60 min) 0.1578 min−1
TiO2/rGO/activated carbon621 Hydrothermal method 5 × 10−4 Mb 2.0 g L−1 Xenon lamp (solar simulator) ∼95% (100 min) 0.0286 min−1
Core–shell g-C3N4@Co–TiO2622 Electrospinning approach/thermal polymerization 20 mg L−1[thin space (1/6-em)]b (10 mL) 2 × 2 cm2 membrane 7 Xenon lamp (300 W), λ > 420 nm, 50 mW cm−2 90.8% (60 min) 0.038 min−1
Hierarchical 2% Au–g-C3N4–ZnO623 In situ preparation of g-C3N4 ZnO nanorods on g-C3N4 nanosheets and the deposition of Au nanoparticles 50 mg L−1[thin space (1/6-em)]a (50 mL) 10 mg 9.3 Xenon lamp 74.7% (30 min) 3.998 × 10−2 min−1
Mesoporous TiO2-modified ZnO QDs immobilized on LLDPE624 Casting method 40 mg L−1[thin space (1/6-em)]a (100 mL) 9 Fluorescent lamp: 48 W 89.5% (90 min) 0.01312 min−1
7% CuO/g-C3N4625 Dispersion method 50 mgb (1000 mL) 0.2 g Xenon lamp (500 W), λ > 365 nm 55% (60 min) 0.014 min−1
ZnO globular (15 wt%)/g-C3N4626 In situ growth 20 mg L−1[thin space (1/6-em)]a (100 mL) 20 mg PLS-SXE300 (300 W), L.I: 9.6 W m−2, 400–780 nm 78.4% (50 min)
ZnO (20 wt%)/GO (2 wt%)/Ag3PO4628 Ultrasonic-assisted precipitation method 30 mg L−1[thin space (1/6-em)]b (50 mL) 1.0 g L−1 6 Visible lamp: 65 W 96.32%
g-C3N4/C/Fe3O4629 Sonication and in situ precipitation technique 10 mg L−1[thin space (1/6-em)]a (40 mL) 10 mg Xenon lamp (500 W) 96.4% (120 min) 0.0292 min−1
Core–shell BiFeO3/TiO2630 Hydrolysis and precipitation method 20 mg L−1[thin space (1/6-em)]a (300 mL) 1 g L−1 5 UV light 67.9% (180 min)
Core–shell BiFeO3/TiO2630 Hydrolysis and precipitation method 20 mg L−1[thin space (1/6-em)]a (300 mL) 1 g L−1 5 Visible light 72.2% (180 min)
Fiber-shaped Ag2O/Ta3N5 (molar ratios: 0.3/1)631 Electrospinning–calcination–nitridation method, followed by in situ anchoring of Ag2O deposition 10 mg L−1[thin space (1/6-em)]a (80 mL) Xenon lamp (300 W), λ > 400 nm 78.3% (60 min) 0.0079 min−1
BiVO4/TiO2/rGO633 Reaction under Teflon reactor 10 μg L−1[thin space (1/6-em)]a Xenon lamp (1000 W), λ > 420 nm 96.2% (120 min) 0.02613 min−1
g-C3N4/AgBr/rGO634 Mixing followed by heating 20 mg L−1[thin space (1/6-em)]a 0.05 g (100 mL) Xenon lamp (250 W) 78.4% (90 min)
C3N4@MnFe2O4–rGO491 Impregnation approach 20 mg L−1[thin space (1/6-em)]a (50 mL) + PS 50 mg Xenon lamp (300 W) with 400 nm cut-off filter 94.5% (60 min) 0.0337 min−1
BiOl/exfoliated C3N4 (mass ratio: 0.4)636 Combination of thermal exfoliation and chemical precipitation 20 mg L−1[thin space (1/6-em)]a (50 mL) 1.0 g L−1 6 Xenon lamp (500 W) with 420 nm cut-off filter 86% (30 min) 0.0705 min−1
3.0 wt% CuO@ZnO637 One-pot method 20 ppma 1.5 g L−1 Xenon lamp (300 W), λcutoff: 420 nm, 45.2 mW cm−2 100% (45 min) 113.50 × 10−3 min−1
ZnO/5 wt% SnO2638 Solvothermal process 1 g L−1[thin space (1/6-em)]b (100 mL) 60 mg Xenon lamp (300 W), λ: 420–780 nm ∼90% (60 min) 0.0385 min−1
Cu2O–TiO2639 Surfactant-free preparation method (TiO2[thin space (1/6-em)]:[thin space (1/6-em)]Cu2O = 0.1; 0.2; 0.3) 50 mga (100 mL) 30 mg Xenon lamp (300 W) 91% (60 min) 0.0432 min−1
10 wt% MoS2/Ag/g-C3N4640 Ag deposition and MoS2 coupling is applied co-modify g-C3N4 nanosheets 20 mg L−1[thin space (1/6-em)]a (50 mL) 10 mg 5.5 Xenon lamp (300 W), λ > 420 nm 90.1% (30 min) 0.0507 min−1
g-C3N4 (7.1 wt%)/ZrO2−x641 Anodic oxidation and thermal deposition method (0.06 g melamine taken) 10 ppmb (5 mL) 2 mg Xenon lamp (300 W), λ > 420 nm 90.6% (60 min) 0.0474 min−1
3 wt% needle SnO2 needle nanoparticles anchored on exfoliated g-C3N4642 Hydrothermal method 30 mg L−1[thin space (1/6-em)]a (100 ml) 50 mg Xenon lamp (250 W) with a cut-of filter of 420 nm 95.90% (120 min) 0.0205 min−1
N-doped ZnO nanorods–MoS2 nanoflowers (1 wt% MoS2 loaded in N–ZnO)643 Hydrothermal strategy 0.01 g L−1[thin space (1/6-em)]a 25 mg (50 mL) CFL lamp (45 W), λ ≥ 420 nm 84% (120 min) 14.43 × 10−3 min−1
WO3/g-C3N4644 In situ hydrolysis and polymerization process 25 mg L−1[thin space (1/6-em)]b (100 mL) 50 mg Xenon lamp (300 W) with a 420 nm cut-off filter 82% (120 min) 0.0164 min−1
Sn3O4/g-C3N4 (with load ratio of 3%)645 Two-step hydrothermal process 10 mg L−1[thin space (1/6-em)]b (100 mL) 50 mg Xenon lamp (500 W) 72.2% (120 min) 0.0108 min−1
BiOI/g-C3N4/CeO2 (3 wt%)646 Calcination and hydrothermal treatment 20 mg L−1[thin space (1/6-em)]a (30 mL) 50 mg Xenon lamp (300 W), λ > 420 nm 91.6% (120 min) 0.0205 min−1
TiO2−x/g-C3N4647 Grinding and in situ reduction 10 mg L−1[thin space (1/6-em)]b (50 mL) 50 mg 9 Xeon-lamp (300 W), λ > 420 nm 87.7% (90 min) 31.7 × 10−3 min−1
K doped g-C3N4/TiO2/CdS648 Hydrothermal method 20 mg L−1[thin space (1/6-em)]a (50 mL) 50 mg Xenon lamp (300 W), λ > 420 nm 94.2% (30 min) 0.08554 min−1
γ-Fe2O3 nanospheres (5%) anchored on g-C3N4649 Anchoring mesoporous γ-Fe2O3 nanospheres on g-C3N4 nanosheet surface 10 mg L−1[thin space (1/6-em)]b (100 mL) 50 mg Xenon light source (500 W) with 420 nm cut-off filter 73.8% (120 min) 0.0134 min−1
0.50 wt% CQDs/g-C3N4650 Low-temperature process 10 mg L−1[thin space (1/6-em)]b (100 mL) 50 mg Xenon lamp (250 W) with 420 nm UV-cut-off filter 78.6% (210 min) ∼0.0065 min−1
BiOBr/MoS2/graphene oxide652 Hydrolysis method 10 mg L−1[thin space (1/6-em)]b 25 mg Xenon lamp (300 W) with 380-nm cut-off filter >98% (40 min) 0.04277 min−1
g-C3N4/MnO2/GO653 Wet-chemical method 10 mg L−1[thin space (1/6-em)]b (100 mL) 0.5 g L−1 6 Xenon lamp (300 W) with a 420 nm filter 91.4% (90 min)
BiVO4@Polypyrrole/g-C3N4654 Dispersion method 30 mg L−1[thin space (1/6-em)]a (50 mL) 30 mg Xenon lamp (300 W), λ > 420 nm 90% (120 min)
AgI/BiOBr/rGO655 Solvothermal method followed by in situ precipitation 20 mg L−1[thin space (1/6-em)]a (100 mL) 50 mg Xenon lamp (5.00 W), simulated sunlight 94.2% (80 min) 0.018 min−1
Ag/Ag3PO4/BiVO4/rGO656 In situ deposition method followed by photo-reduction 10 mg L−1[thin space (1/6-em)]a 0.5 g L−1 6.75 Xenon lamp (300 W) 94.96% (60 min)
g-C3N4/WO3657 Dispersion method 10 mg L−1[thin space (1/6-em)]b (100 ml) 40 mg Xenon lamp (500 W), 320–780 nm, 100 mW cm−2 79.8% (180 min)
(50%)CuIn2S4/g-C3N4658 Synthesis under autoclave 20 mg L−1[thin space (1/6-em)]a (100 mL) 0.5 g L−1 Xenon lamp (300 W), λ > 420 nm cut-off filter 83.7% (60 min) 0.02583 min−1
TiO2/g-C3N4662 Co-annealing process 20 mg L−1[thin space (1/6-em)]b (40 mL) 250 mg 7 Xenon lamp (150 W) 99.40% (120 min) 3.70 × 10−4 min−1


3.8 Diclofenac

Diclofenac (DCF), an important non-steroidal anti-inflammatory drug, finds multifaceted applications as a painkiller primarily for dysmenorrhea, rheumatoid arthritis and inflammation.663,664 The intake of diclofenac even at low levels by humans and other living organisms is reported to have an adverse biochemical effect. The solubility and high polarity of diclofenac in water and lower degradability account for its water pollution. Further, it can accumulate in food chains owing to its migration through the aquatic medium (surface water, drinking water, underground water) in food chains. In view of this, the following photocatalytic methods have been used in the removal of diclofenac from water.665–748
3.8.1 Metal oxides. Rizza et al.667 studied the degradation of diclofenac sodium by UV/TiO2 for a wide range of initial DCF concentrations (5–80 mg L−1) and photocatalyst loadings (0.2–1.6 g L−1) in a batch reactor system. These results showed 100% removal of DCF compared to ∼3% and 14% for TiO2 (dark conditions) and photolysis (UV) corresponding to the initial concentration of 5 mg L−1 and catalyst dosage of 0.2 g L−1. The photocatalytic degradation of real pharmaceutical wastewater (pH: 9) including diclofenac and other drugs by TiO2/H2O2 was found to be 45.11% under UV-mediated irradiation within 120 minutes.668 TiO2 nanofilm membranes fixed on glass panels have also been explored in the removal of diclofenac sodium from wastewater under UV irradiation.669 Schulze-Hennings et al.670 studied the durability of the coating containing TiO2 on glass for the photocatalytic degradation of diclofenac sodium in water using UVA irradiation. The effectiveness of ZnO and V2O5 has also been tested in the photocatalytic degradation of diclofenac sodium in water under solar and UV irradiation.671 The emerging findings indicated 100% photodegradation efficiency for V2O5 compared to ZnO under UV and solar irradiation corresponding to the initial DCF concentration of 300 mg L−1, catalyst dosage of 1.0 g L−1 and pH 4. The relatively higher rate constant values of V2O5 under UV (k: 0.0196 min−1) and solar (k: 0.0141 min−1) irradiation compared to the corresponding values for ZnO in the photodegradation of DCF also supported this. In another report, investigations were made to study the factors affecting diclofenac decomposition in water by UVA/TiO2 photocatalysis.672 According to Bagal et al.,673 UV/TiO2/H2O2 fabricated by a hydrodynamic cavitation approach showed 95% degradation of diclofenac sodium under the optimized operating conditions.

ZnO showed highly active photodegradation of diclofenac sodium in aqueous solution under UV lamp irradiation compared to solar radiation.674 Mimouni et al.675 investigated the effect of heat treatment on the photocatalytic activity of α-Fe2O3 nanoparticles towards diclofenac elimination. The findings in Fig. 22(a) and (b) show the highest degradation for α-Fe2O3 (calcinated at 300 °C) and the value of the degradation rate constant corresponds to 0.060 min−1. The generation of extremely active OH· radicals is responsible for the total photodegradation of DCF, as schematically described in Fig. 22(c). Meroni et al.676 achieved 70% degradation of diclofenac (25 ppm) by a piezo-enhanced sonophotocatalytic approach based on ZnO (0.1 g L−1) subjected to UV-light irradiation for 360 min. In addition, ZnO modified with rare earth elements (Ce, Yb) and Fe,677 NixZn1−x Fe2O4 (x = 0, 0.3, 0.7),678 cobalt ferrite,679 MgO,680 and WO3681 photocatalysts have also been investigated for the removal of diclofenac from aqueous solution.


image file: d3lf00142c-f22.tif
Fig. 22 (a) Conversion plots for photodegradation of DCF in the presence of α-Fe2O3 calcinated at different temperatures. (b) The degradation rate of different samples at 120 min. (c) Schematic presentation on the generation of OH· radicals in α-Fe2O3. Reproduced from ref. 675 with permission from Springer (2022).
3.8.2 Metal–metal oxides. Chakhtouna and coworkers682 reviewed the role of Ag nanoparticles in enhancing the photocatalytic activity of Ag/TiO2 in the removal of pharmaceutical pollutants from aqueous solutions under UV and visible light. Espino-Estévez et al.683 synthesized Ag and Pd nanocomposites of TiO2 (TiO2–Ag and TiO2–Pd) by a sol–gel method and observed almost 100% (120 min) photocatalytic degradation of diclofenac sodium salt in water under a UV light source. It was also noted that photocatalytic degradation of DCF follows first-order kinetics. In another study, Ag@Ag2O/WO3 and Ag@Ag2S/WO3 were prepared by following a deposition hydrothermal route and used as photocatalysts.684 Subsequent studies have shown high degradation of DCF (60 mg L−1, pH: 12) in the presence of H2O2 (1 × 10−4 M) under visible light (λ > 420 nm, 160 W) in the presence of Ag@Ag2O/WO3 (k = 32.0 × 10−3 min−1) and Ag@Ag2S/WO3 (k = 7.3 × 10−3 min−1) catalysts. Further investigations have also revealed that ·O2 plays an important role in the degradation of DCF.
3.8.3 Doped metal oxides. Nguyen et al.685 removed diclofenac from wastewater using a submerged photocatalytic membrane reactor comprising immobilized N–TiO2 under visible irradiation. It was also noted that DCF removal efficiency is enhanced under visible irradiation by coupling H2O2 with the photocatalytic process. C-doped TiO2 synthesized by a microwave digestion method showed almost complete removal of diclofenac after about 160 min under visible light corresponding to diclofenac concentration of 50 mg L−1, catalyst concentration of 250 mg L−1 and light intensity of 8000 lx.686 The doping of titania with 25 wt% Mg resulted in 55% and 48% degradation of diclofenac sodium under UV and visible irradiation, respectively.95 An Mn (0.6 mol%) and Ag (0.5 mol%) co-doped TiO2 aerogel exhibited 86% removal of diclofenac under UVA-light irradiation after 4 h.687 The photodegradation rates followed first-order kinetics with a highest apparent rate constant of 0.0064 min−1.

The photocatalytic performance of a sodium diclofenac solution (pH: 6.5) in F-doped (20 wt%) ZnO under simulated solar radiation indicated the complete degradation of diclofenac sodium of concentration: 10 mg L−1 under the optimized experimental conditions (ZnO–F concentration: 1 g L−1).688 The enhanced photocatalytic activity of F-doped TiO2 is ascribed to the reduction in the recombination rate of electron–hole pairs. In another similar study, fluorine (0.25, 0.5 and 1 at%)-doped ZnO nano- and meso-crystalline ZnO showed high rates of diclofenac degradation in water compared to bare ZnO.689 Chaudhari and others690 used a sol–gel method to prepare Mn/CeO2, Cu/CeO2 Ag/CeO2 (metal semiconductors) and Agl/CeO2 (an n–p semiconductor–semiconductor) by doping with Mn, Cu, Ag and AgI, respectively. Further investigations have been made to compare their photocatalytic degradation for diclofenac sodium in water under the same optimal conditions (pH: 7, diclofenac concentration: 10 ppm) within 90 min exposure to UV light. It is noted that AgI-doped CeO2 (1 g L−1) exhibited higher degradation of diclofenac sodium solution (95%) compared to Mn/CeO2, Cu/CeO2 or Ag/CeO2, such enhancement in the photocatalytic activity of AgI/CeO2 is attributed to its larger surface area and charge separation efficiency.

In addition, Ce@TiO2,691 granular activated carbon modified with N-doped TiO2,692 C,N-co-doped TiO2,693 Ce,Mn-co-doped TiO2,694 N,S-co-doped carbon quantum dots/TiO2,695 TiO2 doped with B, F, N, P,696 and S,N,C-tri-doped TiO2697 photocatalysts have been investigated for the removal of diclofenac from aqueous solution.

3.8.4 Metal oxide composites. Alalm et al.182 investigated the solar photocatalytic degradation of pharmaceuticals, namely amoxicillin, diclofenac, and paracetamol, using TiO2 immobilized on powdered activated carbon (TiO2/AC). According to this, degradation corresponding to the initial concentration of pharmaceuticals of 50 mg L−1 and TiO2/AC dosage of 1.2 g L−1 followed the order: amoxicillin (100%: in 120 min) > diclofenac (83% beyond 180 min) > paracetamol (70% in 180 min). TiO2–WO3 (molar ratio: 10[thin space (1/6-em)]:[thin space (1/6-em)]1) synthesized by a hydrothermal method was the most effective catalyst in the photocatalytic removal of diclofenac under visible-light irradiation compared to pure TiO2.698 The composite catalyst successfully degraded diclofenac almost completely in 270 min corresponding to pH 5, initial diclofenac concentration of 25 mg L−1 and catalyst concentration of 0.6 g L−1. Subsequent studies showed the catalyst retained 80% catalyst efficiency after four consecutive reaction cycles. N-doped WO3/TiO2 synthesized by a sol–gel method enhanced the degradation of diclofenac sodium using simulated solar light owing to the synergistic effect and narrowing of the bandgap.699 The visible-light-irradiated photocatalytic degradation of diclofenac sodium using ZnO–WO3 has shown better catalytic activity than bare ZnO.700 These studies revealed ZnO–WO3 (Zn[thin space (1/6-em)]:[thin space (1/6-em)]W mole ratio: ≈10[thin space (1/6-em)]:[thin space (1/6-em)]1) exhibiting ∼76% degradation efficiency at a given pH (6), DCF diclofenac concentration (20 mg L−1) and catalyst loading (0.8 g L−1).

Cordero-García et al.701 studied the effect of carbon doping on WO3/TiO2 on the photocatalytic degradation of diclofenac sodium and observed its higher photocatalytic activity compared to WO3/TiO2 and TiO2. Hydroxyapatite/TiO2 (dose: 4 g L−1) in water degraded DCF (initial concentration: 5 ppm) by 95% in 24 h on irradiating it with simulated solar light.702 According to Sun et al.,703 the intensity of UV irradiation plays a more significant role in the significant removal of diclofenac by a nano-TiO2/diatomite composite in a photocatalytic reactor. According to this, diclofenac degraded completely at 30 min under higher UV irradiation intensity at a flux of 3.0 L h−1. A visible-light-responsive TiO2/Ag3PO4 (10[thin space (1/6-em)]:[thin space (1/6-em)]1) nanocomposite immobilized in a spherical polymeric matrix showed almost complete removal of diclofenac (k: 0.018 min−1) in 120 min corresponding to initial drug concentration of 20 mg L−1 bead loading of 10 g L−1, and reaction volume of 0.8 L.704 The ·OH radical and h+ are reported to be the primary reactive oxygen species in the photodegradation of diclofenac.

An Ag–Ag2O/reduced TiO2 nanophotocatalyst demonstrated 99.8% degradation of diclofenac after 50 min of visible irradiation.705 This is attributed to the effective charge separation, enhanced visible light absorbance and localized SPR of nanocrystalline Ag0. Silvestri et al.706 synthesized PPy–ZnO (25[thin space (1/6-em)]:[thin space (1/6-em)]1) via a polymerization method and studied the degradation of DCF under simulated solar light. In this regard, the composite catalyst (1 g L−1) facilitated 81% (60 min) degradation of diclofenac (10 mg L−1) with h+ the main reactive species involved in the reaction. This performance is ascribed to the mesoporous structure, superior surface area and reduced band gap of PPy–ZnO. According to Das et al.,707 a titania–zirconia (Zr/Ti mass ratio of 11.8 wt%) composite catalyst exhibited a reasonably higher removal of DCF (∼92.41%) compared to the anatase form of titania without zirconia.

Attempts have been made to eliminate diclofenac sodium from wastewater through the photocatalytic degradation of hydrothermally prepared TiO2–SnO2 (Ti–Sn molar ratio: 1[thin space (1/6-em)]:[thin space (1/6-em)]1, 5[thin space (1/6-em)]:[thin space (1/6-em)]1, 10[thin space (1/6-em)]:[thin space (1/6-em)]1, 20[thin space (1/6-em)]:[thin space (1/6-em)]1 and 30[thin space (1/6-em)]:[thin space (1/6-em)]1) under various operating conditions.708 The results indicated the TiO2–SnO2 catalyst with a molar ratio of 20[thin space (1/6-em)]:[thin space (1/6-em)]1 to be the most effective photocatalyst compared to the other binary composites. The catalyst achieved complete degradation of diclofenac under optimum conditions comprising initial drug concentration of 20 mg L−1, catalyst loading of 0.8 g L−1 and pH 5. The photocatalyst also displayed excellent repeatability and better stability over repeated reaction cycles. Fe3O4/TixOy/activated carbon,709 Fe3O4 (nanosphere)/Bi2S3 (nanorod)/BiOBr (nanosheet)710 TiO2@ZnFe2O4/Pd,711 nanotubular titanium dioxide–polyethersulfone (PES) membrane,712 Al2O3–Nd2O3,713 and TiO2–zeolite714 based photocatalysts have also been evaluated for the photocatalytic degradation of diclofenac.

3.8.5 Graphitic materials.
3.8.5.1 g-C3N4 and its composites. Carbon quantum dot (CQD)-modified porous g-C3N4 (dose: 200 mg L−1) synthesized using 20 mL of CQD stock solution showed almost complete degradation of diclofenac solution (pH: 9) of an initial concentration of 10 mg L−1 in 12 min under visible light.715 This is attributed to the tuning of the band structure and enhanced separation of charge carriers. The studies also suggested DCF degradation to be dominated by a photosensitization-like mechanism. The CQD/g-C3N4 photocatalyst also exhibited excellent reusability, as evident from studies in the 5th cycle (>90%). Pd quantum dots (1 wt%) deposited on g-C3N4 (dose: 0.5 g L−1) achieved 100% removal of diclofenac solution (initial concentration: 1 mg L−1, pH: 7) within 15 min under solar light.716 The rate constant (0.72 min−1) was found to be 8 times higher than that of g-C3N4. Such enhanced photocatalytic activity has been explained based on its narrowed bandgap, reduction in the recombination of photogenerated charge carriers and availability of a photosensitization-like electron transfer pathway.

Graphite-like C3N4-modified Ag3PO4 nanoparticles exhibited highly enhanced photocatalytic activity under visible-light irradiation owing to the synergistic effect.717 This is mainly ascribed to the matching band potentials between Ag3PO4 and g-C3N4, effectively suppressing recombination of electron–hole pairs and promoting their separation efficiency. Diclofenac sodium and ibuprofen (5 mg L−1) achieved complete degradation (180 min) in the presence of carbon microspheres (dia: 0.9–1.9 μm) supported on an anatase phase of TiO2 (mass ratio TiO2 to C microspheres: 2) heterostructure photocatalyst under solar light.718 Further studies revealed the high performance of the photocatalyst even after five successive cycles (80%) as evident from the findings in the first cycle (94%).

Hu et al.719 fabricated eco-friendly 2D heterojunction photocatalyst composites (BCCNT) comprising C-doped supramolecule based g-C3N4 (BCCN) layers and TiO2 nanoparticles and corresponding findings are displayed in Fig. 23(a). It should be noted that degradation of diclofenac solution (10 mg L−1, initial pH: 5.05) by 1 g L−1 of 30% C-doped supramolecular based g-C3N4 (BCCNT) reached 98.92% within 30 min under LED lamp illumination owing to ·O2 and h+ as the main active species. Further investigations established that the degradation kinetics of DCF fitted the pseudo-first-order equation (Fig. 23(b)) with an apparent reaction rate constant (kapp: 0.1796 min−1) about 29.4 times higher than BCCN (0.0061 min−1). A possible mechanism for the photodegradation of DCF under LED lamp irradiation is also displayed in Fig. 23(c).


image file: d3lf00142c-f23.tif
Fig. 23 (a) Ct/C0versus time plots of different photocatalysts. (b) Respective kinetic curves (inset) and apparent reaction rate constants of diclofenac (conditions: [DCF]0 = 10 mg L−1, [Catal.] =1 g L−1, no pH adjustment and pHinitial = 5.05) and (c) possible mechanism for the photodegradation of DCF and CBZ under LED lamp irradiation over 30% BCCNT composites. Reproduced from ref. 719 with permission from Elsevier (2019).

An AgI/gC3N4 (AgI molar mass ratio: 45%) composite photocatalyst exhibited almost complete degradation of diclofenac sodium in 6 min under visible-light irradiation compared to AgI and g-C3N4.720 The reaction rate constant value of AgI/gC3N4 (k: 0.561 min−1) was found to be ∼12.5 and 43.2 times higher than those achieved by AgI (0.045 min−1) and g-C3N4 (0.013 min−1). The photocatalytic degradation of diclofenac was guided by photogenerated holes and superoxide anion radicals as the main reactive species. Such enhanced photocatalytic activity of AgI/g-C3N4 is ascribed to the heterojunction between g-C3N4 and AgI that facilitated interfacial charge transfer and prevented the recombination of electron–hole pairs. Ag/g-C3N4 (mass ratio of Ag: 54%) heterostructure photocatalysts prepared by photodeposition under ambient conditions showed complete degradation of DCF compared to g-C3N4 under visible-light irradiation and followed pseudo-first-order kinetics. The rate constant was k = 0.0429 min−1.721 The rate constant of diclofenac degradation over Ag/g-C3N4 was almost 3.1 times higher than that of pure g-C3N4. Further investigations also revealed generated holes as the main reactive species in diclofenac degradation and also established the excellent stability of Ag/g-C3N4. CNT–Ni@TiO2:W nanoparticles722 and C3N4/NH2-MIL-125 (ref. 723) have also shown remarkable performance in the removal of diclofenac present in water.


3.8.5.2 Graphene composites. The removal of diclofenac (and amoxicillin) has been reported by maltodextrin/reduced graphene and maltodextrin/reduced graphene/copper oxide nanocomposites.724 Kovacic et al.725 fabricated S-doped TiO2/rGO by a one-pot solvothermal method to study the removal of diclofenac sodium in aqueous medium (pH 4) under simulated solar irradiation. These findings revealed strong dependence on rGO loading of the photocatalytic performance of S–TiO2/rGO in the degradation of DCF. Accordingly, 5 wt% rGO in TiO2 showed improved diclofenac photocatalytic activity compared to bare TiO2 owing to the effective photogenerated charge separation, as inferred from a photoluminescence study. John et al.726 investigated sunlight-mediated removal of diclofenac sodium from water (25 mg L−1) using TiO2–reduced graphene oxide (75 mg L−1) and persulfate (20 mg L−1). They achieved an efficiency of more than 98% within 30 min under sunlight illumination. The diclofenac degradation followed the Langmuir–Hinshelwood mechanism and pseudo-first-order kinetics with a pseudo-first-order rate constant (99.4 × 103 min−1) about twice that of TiO2–rGO (50.9 × 10−3 min−1). A hydrothermally synthesized BiOCl–GO composite showed 100% and 47.88% removal of DCF from solution (25 mg L−1) under UV light and visible spectrum solar light, respectively.727 Li et al.728 also used a hydrothermal method to synthesize an Ag–BiOI–rGO nanocomposite. They observed the complete removal of diclofenac (10.0 mg mL−1) by 5 mol% Ag–BiOI–rGO (5 wt%) in 80 min under visible-light irradiation compared to pure BiOI, Ag–BiOI or BiOI–rGO photocatalysts (50 mg in 50 mL). This is attributed to the enhanced charge separation and reduced recombination of photogenerated charge carriers due to Ag and rGO in BiOCl. Other studies reported ∼93% decomposition of diclofenac sodium (25 mg L−1) solution (pH: 6) within 6 min by cubic Ag/AgBr/GO (0.030 g) on illumination with sunlight.729 It is suggested that the large surface area of the catalyst as well as the superior charge separation and transfer efficiency accounted for this. UV-light-assisted activation of persulfate by rGO–Cu3BiS3 (30 mg) reportedly achieved 81% degradation of DCF in 60 min.730 An AgFeO2–graphene/Cu2(BTC)3 MOF heterojunction has also been studied under sunlight for the degradation of diclofenac in aqueous solution.209
3.8.6 Heterojunctions, S- and Z-scheme-based composites. Co3O4/WO3 nanocomposites were fabricated by dispersing WO3 in a solution of cobalt acetate (pH: 7) followed by heating at 90 °C.731 It showed 90.8% degradation of diclofenac sodium salt (15 ppm) solution (pH: 10.7) under visible-light irradiation. According to this, the formation of a monoclinic phase of WO3 and a p–n heterojunction maximizing the generation of non-selective OH radicals and reducing electron–hole pair combination and the strong absorption of visible light account for such a performance. A solar-active Fe3O4@SrTiO3/Bi4O5I2 heterojunction photocatalyst imparted 98.4% diclofenac removal in 90 min under simulated solar-light irradiation.732 A vis–NIR-driven S-scheme, an WO3−x/S-doped g-C3N4 nanocomposite, exhibited ∼99.5% degradation rate for diclofenac.733 g-C3N4/BaBiO3 heterojunctions contributed enhanced photocatalysis of diclofenac sodium under visible light through interfacial charge transfer.734 The photocatalytic activity of g-C3N4/BaBiO3 is reported to be 6.5 and 5 times higher than BaBiO3 and g-C3N4, respectively. Visible-light-responsive N,S-co-doped TiO2@MoS2,735 S,B-co-doped g-C3N4 nanotube@MnO2,736 oxygen-doped-g-C3N4/ZnO/TiO2@halloysite nanotubes,737 and Pt–TiO2–Nb2O5738 also displayed enhanced photocatalytic degradation of diclofenac.

The optimal BiOCl/CuBi2O4 exhibited a 90% degradation rate for aqueous DCF in 60 min under visible-light irradiation.739 The degradation followed pseudo-first-order kinetics (k: 0.03539 min−1), much higher than CuBi2O4 (k: 0.00139 min−1) or BiOCl (k: 0.00319 min−1). Such enhanced photocatalytic performance of BiOCl/CuBi2O4 is most likely to be due to the upgraded charge separation and transfer caused by the formation of an S-scheme heterojunction and the presence of oxygen vacancies. Chen et al.740 investigated the photocatalytic performance and mechanism of a Z-scheme CuBi2O4/Ag3PO4 photocatalyst in the degradation of diclofenac sodium under visible-light irradiation. Studies have also been reported on Z-scheme CuBi2O4/Ag3PO4 to study the effects of pH, H2O2, and S2O82− on the visible-light-driven degradation of diclofenac sodium.741

Visible-light-driven TiO2/g-C3N4 achieved maximum degradation efficiency (93.49%) for the removal of diclofenac sodium from aqueous solution (5 ppm) and the process followed pseudo-first-order kinetics.742 Such a Z-scheme photocatalyst successfully prevents the fast recombination of electron–hole pairs. Elangovan and others743 prepared a TiO2–CdS heterojunction following a two-step hydrothermal treatment. Subsequent use of this as a photocatalyst achieved 86% diclofenac degradation within 4 h under visible-light irradiation. It was suggested that the direct Z-scheme heterojunction structure accounts for the direct charge transfer between heterojunction catalysts. Investigations of a TiO2–CdS photocatalyst in five successive reaction cycles established its appreciable photochemical stability and reusability. ZnSnO3/Bi2WO6,744 Ag3PO4/g-C3N4,745 V2O5-B-doped g-C3N4,746 MoS2/Cd0.9Zn0.1S747 and MoO3@ZrO2748 photocatalysts have also shown enhanced degradation of diclofenac and diclofenac sodium.

Table 9 records the data on the performance of metal oxides and carbonaceous materials based photocatalyst in the removal of diclofenac from water under optimum conditions.

Table 9 Performance data on removal of diclofenac in water presence of various photocatalysts
Photocatalysts Preparative method DCF Catalyst dose pH Light source Degradation and time Rate constant
TiO2[thin space (1/6-em)]665 Sol–gel method 5 ppm (100 mL) 50 mg 6 Xenon arc lamp, 300 W, 70 mW cm−2, λcut-off: 420 nm ∼80% (120 min)
TiO2[thin space (1/6-em)]665 Sol–gel method 5 ppm (100 mL) 50 mg 6 Natural sunlight ∼72% (120 min)
TiO2P25666 Commercial 2 mg L−1 200 mg L−1 Blacklight Philips TLK05 (40 W), 290–400 nm 100% (60 min) ∼0.09 min−1
TiO2SG666 Commercial 2 mg L−1 200 mg L−1 Blacklight Philips TLK05 (40 W), 290–400 nm 100% (30 min) ∼0.13 min−1
TiO2 aerogel P25 (Degussa)667 Commercial 5 mg L−1 (100 mL) 0.2 g L−1 125 W black light fluorescent lamp: 300–420 nm 100% (80 min) 4.24 × 10−2 min−1
TiO2 nano thin film on glass slide669 Chemical bath deposition 10 ppm 25 × 75 mm deposited film 2 UV lamp 26% (12 min)
TiO2 immobilized on glass670 Solution method 0.5 mg L−1 Film of area 144 cm2 6.2–7.2 UVA lamp: 15 W (300–400 nm) ∼100% (26 h) 0.15 h−1
ZnO (Merck)671 Commercial 300 mg L−1 1.0 g L−1 4 UV 90.7% (180 min) 0.0144 min−1
ZnO (Merck)671 Commercial 300 mg L−1 1.0 g L−1 4 Solar 56.5% (190 min) 0.0044 min−1
V2O5 (Merck)671 Commercial 300 mg L−1 1.0 g L−1 4 UV ∼100% (180 min) 0.0196 min−1
V2O5 (Merck)671 Commercial 300 mg L−1 1.0 g L−1 4 Solar ∼100% (180 min) 0.0141 min−1
TiO2 immobilized on activated carbon182 Temperature impregnation method 50 mg L−1 (4 L) 1.2 g L−1 10 Solar irradiation ∼85% (180 min) 0.010 min−1
Degussa P25 TiO2 (75% A:25% R)/H2O2: 1.4 mM672 Commercial 5 mg L−1 250 mg L−1 UVA lamp (9 W lamp) ∼99.5% (60 min)
TiO2 (anatase and rutile)673 Commercial 20 ppm (5 L) 0.3 g L−1 4 UV lamp: 250 W 80.25% (120 min) 0.0152 min−1
TiO2 (anatase and rutile)/H2O2: 0.3 g L−1[thin space (1/6-em)]673 Commercial 20 ppm (5 L) 0.3 g L−1 4 UV lamp: 250 W 95.7% (120 min) 0.0273 min−1
ZnO674 Commercial 30 μM 0.25 g L−1 3 UV lamp: 40 W, 254 nm 95% (5 min) 0.403 min−1
α-Fe2O3 nanoparticles (calcinated at 300 °C)675 Drying followed by heat treatment 15 mg L−1 (100 mL) 1 g L−1 UVC lamp: 15 W, 254 nm 96% (120 min) 0.04 min−1
MgO nanoparticles680 Direct precipitation method 10 mg L−1 0.1 g 6.5 UV light source (254 nm) 100% (60 min) 0.1191 min−1
TiO2–Pd683 Sol–gel method 50 g L−1 (0.20 L) 1 g L−1 5 UV light source (15 W), 300–400 nm 100% (120 min) ∼0.05 min−1
TiO2–Ag683 Sol–gel method 50 mg L−1 (0.20 L) 1 g L−1 5 UV light source (15 W), 300–400 nm 100% (120 min) ∼0.04 min−1
Ag/Ag2O/WO3 (H2O2: 1 × 10−4 mM)684 Deposition/hydrothermal 0.006 g (100 mL) 0.1 g 12 Mercury lamp (160 W), λ ≥ 400 nm 85% (60 min) 32.0 × 10−3 min−1
C-doped TiO2 (anatase phase)686 Microwave digestion method 50 μg L−1 250 mg L−1 7.5 High-pressure W visible lamp (150 W), λ > 400 nm, 8000 lx ∼100% (150 min) 0.0334 min−1
Mg (25 wt%)-doped SiO2687 Mixing of Mg/SiO2 with MgCl2 20 mg L−1 (25 mL) 0.7 g L−1 4.3 UV light 55% (60 min)
Mg (25 wt%)-doped SiO2687 Mixing of Mg/SiO2 with MgCl2 20 mg L−1 (25 mL) 0.7 g L−1 4.3 Visible light 48% (60 min)
F (0.25)-doped ZnO nano689 Hydrothermal approach 10 mg L−1 (100 mL) 1.0 g L−1 UV-LEDs strip: 10 W, 365 nm 85% (30 min) and ∼99% (180 min) 0.06 min−1
Mn doped CeO2690 Sol–gel 10 ppm 1.0 g L−1 7 Mercury vapour lamp (125 W) with cut-off wavelength of 455 nm 48% (60 min)
Cu doped CeO2690 Sol–gel 10 ppm 1.0 g L−1 7 Mercury vapour lamp (125 W) with cut-off wavelength of 470 nm 50% (60 min)
Ag doped CeO2690 Sol–gel 10 ppm 1.0 g L−1 7 Mercury vapour lamp (125 W) with cut-off wavelength of 510 nm 57% (60 min)
AgI/CeO2690 Sol–gel 10 ppm 1.0 g L−1 Mercury vapour lamp (125 W) with cut-off wavelength of 460 nm 88% (60 min) 1.758 × 104 L Mol−1 min−1
Ce@TiO2691 Precipitation method 5 μM (100 mL) 75 mg UV light ∼100% (80 min)
1% Ce–0.6% Mn/TiO2694 Sol–gel method 10 mg L−1 50 mg L−1 6 UV lamp: 30 W, λ: 254 nm 94% (240 min) 0.012 min−1
N,S co-doped-CQDs/TiO2695 Via in situ phase inversion method 10 ppm (200 mL) 1.5 g (25 cm2 membrane area) Visible-light irradiation (λ > 400 nm) 62.3% (150 min)
UV light (λ < 380 nm) ∼55% (150 min)
B (5 wt%) doped TiO2696 Sol–gel method 15 mg dm−3 250 mg dm−3 UV lamp ∼30% (120 min) 0.0035 min−1
P (5 wt%) doped TiO2696 Sol–gel method 15 mg dm−3 250 mg dm−3 UV lamp ∼24% (120 min) 0.0019 min−1
F (5 wt%) doped TiO2696 Sol–gel method 15 mg dm−3 250 mg dm−3 UV lamp ∼27% (120 min) 0.0021 min−1
C–S–N-tri-doped TiO2 (thiourea/Ti molar ratio: 0.2[thin space (1/6-em)]:[thin space (1/6-em)]1)697 Sonochemical method 25 mg L−1 (50 mL) 0.05 g L−1 Neutral pH Sunlight 76.48% (90 min) 0.0632 min−1
TiO2–WO3 (10[thin space (1/6-em)]:[thin space (1/6-em)]1 molar ratio)698 Hydrothermal method 25 mg L−1 (100 mL) 0.6 g L−1 5 Metal halide lamp, 400 W, visible light 100% (210 min)
Hydroxyapatite–TiO2702 Annealing of Ti salt and hydroxyapatite 5 mg L−1 (50 mL) 4 g L−1 UV lamp, λ: 365 nm, 1.80 mW cm−2 95% (24 h)
Nano TiO2/diatomite703 Hydrolysis, precipitation and roasting of diatomite and TiCl4 400 μg L−1 0.5 g L−1 UV lamps: 16 W, 254 nm, 1.17 mW cm−2 100% (30 min)
Immobilized (12 wt% TiO2)/Ag3PO4 (10[thin space (1/6-em)]:[thin space (1/6-em)]1)704 Sol–gel method 20 mg L−1 10 g L−1 (beads), 0.8 L Visible light source ∼90% (120 min) 0.018 min−1
4.25-Ag–Ag2O/r-TiO2-0.130705 One-step solution reduction strategy 5 mg·L−1 (100 mL) 30 mg Visible light 100% (50 min) 0.04767 min−1
PPy[thin space (1/6-em)]:[thin space (1/6-em)]ZnO (25[thin space (1/6-em)]:[thin space (1/6-em)]1)706 Via polymerization method 10 mg L−1 (100 mL) 1 g L−1 6 Xenon lamp (250–800 nm) 81% (60 min) 0.986 min−1
TiO2–SnO2 (molar ratio: 20 to 1)708 Hydrothermal method 20 mg L−1 0.8 g L−1 5 UV lamp 100% (300 min) 0.0147 min−1
Fe3O4/Bi2S3/BiOBr (with Bi2S3 mass ratio of 4%)710 One-pot solvothermal 10 mg L−1 (50 mL) 0.03 g L−1 5 LED lamp (50 W), 475 nm 93.81% (40 min) 0.0527 min−1
TiO2@ZnFe2O4/Pd711 Photodeposition technique 10 mg L−1 0.03 g L−1 4 Solar light 84.87% (120 min) 0.0172 min−1
Nanotubular TiO2-PES712 Via anodization of TiO2 nanotubes on polyethersulfone membrane 5 mg L−1 Circular membranes (Dia: 47 mm) UVA sunlamp (7.6 mW cm−2) ∼94% (240 min) 9.96 × 10−3 min−1
Al2O3-(15%) Nd2O3713 Sol–gel method 80 ppm 200 mg (200 mL) UV lamp, 254 nm, 4400 μW cm−2 >92.0% (40 min) 9.5 × 10−2 min−1
CQDs (50 mL) modified g-C3N4715 Mixing method 10 mg L−1 (50 mL) 200 mg L−1 9 Xenon arc lamp (300 W) with UV cut-off filter (λ ≥ 400 nm), 150 ± 5 mW cm−2 100% (12 min) 0.47 min−1
TiO2–carbon microspheres (CMS) with Ti[thin space (1/6-em)]:[thin space (1/6-em)]CMS molar ratio = 2718 Solvothermal treatment 5 mg L−1 (50 mL) 250 mg L−1 6.0 Xenon lamp (500 W m−2). With light correction filter (λ ≤ 350 nm) 100% (180 min)
30% TiO2-hybridize C-doped based g-C3N4719 In situ method 10 mg L−1 (100 mL) 1 g L−1 5.05 LED lamp: 50 W, 380–780 nm 98.92% (30 min) 0.1796 min−1
AgI/g-C3N4 (molar ratio of AgI: 45%)720 Deposition–precipitation method 1 mg L−1 (100 mL) 10 mg Xenon lamp (300 W), λ ≥ 400 nm, 100 mW cm−2 100% (6 min) 0.561 min−1
Ag modified g-C3N4 (mass ratio of Ag: 54%)721 Photodeposition 100 mg L−1 (100 mL) 10 mg Xenon lamp: 300 W with cut-off filter (λ ≥ 400 nm), 100 mW cm−2 ∼100% (120 min) 0.0429 min−1
TiO2–rGO in presence of persulfate726 Solvothermal treatment (using 5 wt% GO) 25 mg L−1 (50 mL), (persulfate:20 mg L−1) 75 mg L−1 4 Sunlight (1.25 × 106 lx) >98% (30 min) 99.4 × 10−3 min−1
BiOCl–GO727 One-pot hydrothermal method 25 mg L−1 (100 mL) 1 g L−1 5 Visible spectrum solar light (17.38 mW cm−2) 47.88% (180 min)
5 Mol% Ag–BiOI–rGO 5 wt%728 Hydrothermal strategy 10.0 μg mL−1 (50 mL) 50 mg Halogen lamp: 300 W 100% (80 min) 0.026 min−1
Ag/AgBr/GO729 Sonochemical route 25 mg L−1 (25 mL) 0.030 g 6.2 Sunlight irradiation ∼93% (6 min)
rGO–Cu3BiS3 (15%)/PS (5 mM)730 Solvothermal process 10 mg L−1 (50 mL) 30 mg UV LED light (15 W) 85% (60 min) 3.8 × 10−2 min−1
Co3O4/WO3 (annealed)731 Dispersion method 15 ppm (50 ml) 30 mg 6.8 Mercury lamp (80 W) with cut-off of 420 nm 90.8% (180 min) 0.1412 min−1
Fe3O4@SrTiO3/Bi4O5I2732 In situ hydrothermal route 10 mg L−1 0.3 mg mL−1 6 Xenon lamp (300 W) 98.4% (90 min) 0.06214 min−1
N,S co-doped TiO2@MoS2735 Hydrothermal method 0.15 mg L−1 0.98 g L−1 5.5 Visible LED light irradiation 98% (150 min) 0.002 min−1
S–B-co-doped g-C3N4 nanotubes–MnO2 (PMS: 0.06 mM)736 Hydrothermal 20 mg L−1 0.5 g L−1 7 Visible light (8 × 8 W), 460 nm 99% (10 min)
Pt–TiO2–Nb2O5738 Multiple steps 12.5 mg L−1 (100 mL) 0.5 g L−1 UV-LED 100% (20 min) 0.446 min−1
BiOCl/CuBi2O4 (mass ratio: 40%)739 Solvothermal process 50 mg L−1 (40 mL) 1 mg mL−1 Xenon lamp (300 W), λ > 420 nm ∼90% (60 min) 0.03539 min−1
CuBi2O4/Ag3PO4 (1[thin space (1/6-em)]:[thin space (1/6-em)]1)740 Combination of hydrothermal and in situ deposition 10 mg L−1 (50 mL) 0.025 g Xenon lamp (300 W) with cut-off filter at λ ≥ 400 nm ∼90% (120 min) 0.0143 min−1
CuBi2O4/Ag3PO4 (mass ratio of 3[thin space (1/6-em)]:[thin space (1/6-em)]7)741 Hydrothermal synthesis and in situ deposition method 10 mg L−1 25 mg (50 mL) 4.42 Xenon lamp (300 W), λ > 400 nm 82% (60 min) 0.0072 minc
CuBi2O4/Ag3PO4 (mass ratio of 3[thin space (1/6-em)]:[thin space (1/6-em)]7)/S2O82−: 1–06 mM741 Hydrothermal synthesis and in situ deposition method 10 mg L−1 25 mg (50 mL) 4.42 Xenon lamp (300 W), λ > 400 nm 100% (60 min) 0.0272 min−1
CuBi2O4/Ag3PO4 (mass ratio of 3[thin space (1/6-em)]:[thin space (1/6-em)]7)/H2O2: 1 mM741 Hydrothermal synthesis and in situ deposition method 10 mg L−1 25 mg (50 mL) 4.42 Xenon lamp (300 W), λ > 400 nm 98.40% (60 min) 0.0162 min−1
TiO2/g-C3N4742 Wet impregnation method 5 ppm 0.3 g 5 W halogen lamp (1000 W) 93.49% (90 min) 0.0324 min−1
Ag3PO4/g-C3N4 (30%)745 Deposition–precipitation method 1 mg L−1 (100 mL) 0.1 g L−1 Xenon lamp (300 W) with filter (λ ≥ 400 nm) ∼100% (12 min) 0.453 min−1
50% V2O5–g-C3N4 (molar ratio: 30%)746 Mixing method 10 mg L−1 0.2 mg mL−1 >7 Monochromatic blue lamps (8 W), 465 ± 40 nm 100% (<105 min) ∼0.53 min−1
MoS2/Cd0.9Zn0.1S747 One-step hydrothermal method 20 μM (50 mL) 25 mg Xenon lamp (300 W) with 420 nm cut-off filter 86% (30 min)


3.9 Atenolol

Atenolol (ATL) belongs to the group of β-blockers and is extensively used in the treatment of cardiovascular diseases, such as hypertension, coronary arterial disease and cardiac arrhythmia.749 As a result, it has been widely detected in sewage effluent, surface water and wastewater treatment plants on its release into the environment through urban discharges. Atenolol can prevent the growth of human embryonic cells and is toxic to water species. Therefore, it is essential to develop simple and cost-effective technologies for the effective removal of ATL in wastewater before release into natural water.750–780
3.9.1 Metal oxides. Several studies have been done into carrying out the degradation of atenolol using commercial as well as synthetic TiO2 compared to ZnO.750–752 Hapeshi et al.753 used a variety of commercially available TiO2 as photocatalysts and found the following relative catalytic activity for the conversion of atenolol: Degussa P25 (67%) > Hombicat UV 100 (39%) > Tronox A-K-1 (30%) > Aldrich (15%) > Tronox TRHP-2 (10%) > Tronox TR (9%) In another study, nano-TiO2 crystal phase (anatase TiO2, rutile TiO2, and mixed phase) coupled with UV-LED was used to study the influence of several parameters on atenolol photodegradation.754 It was noted that the mixed phase completely degraded atenolol in 60 min under UV-LED (365 nm) corresponding to the ATL concentration of 18.77 μM, catalyst dosage of 2.0 g L−1, light intensity of 774 μW cm−2 and pH 7.6. This is in all likelihood due to several contributions originating from the large specific surface area of the catalyst, excellent charge separation efficiency, and the influence of light absorption. The photodegradation of atenolol followed pseudo-first-order kinetics (k: 0.064 min−1).

Among the different commercial TiO2 catalysts, TiO2 (Degussa P25) aqueous suspensions (250 mg L−1) delivered 80% photocatalytic conversion of atenolol (10 mg L−1) under irradiation by a 1 kW Xe-OP lamp in 120 min.755 TiO2 (Degussa P25) has been tested for the removal by degradation of atenolol, acetaminophen, sulfamethoxazole in hospital wastewater.756 Rimoldi et al.757 evaluated the degradation of tetracycline hydrochloride, paracetamol, caffeine and atenolol, both as individual pollutants and in mixtures, using UV and simulated-solar-mediated TiO2. According to Ponkshe and Thakur,758 degradation of atenolol (2 × 10−4 M) using different commercially available TiO2 (0.03 g L−1) as photocatalysts in a 100 mL reaction solution (natural pH) under UV light for 120 min followed the order: Aeroxide TiO2 P25 (94%) > TiO2 Hombikat UV 100 (68%) > Merck TiO2 (60%) > TiO2 Kronoclean 7000 (45%). Rogé et al.759 prepared ZnO nanowires by metal organic chemical vapor deposition and investigated their photocatalytic activity in a solution containing atenolol and sulfadimidine under low-power 365 nm UV light (2.28 mW cm−2). The corresponding pseudo-first-order rate constants in these pollutants were found to be 6.5 × 10−3 and 2.3 × 10−3 min−1. Several other studies also reported the photocatalytic degradation of atenolol in aqueous solution using Degussa TiO2 P25 suspension,760 TiO2,761 TiO2/salicylaldehyde–NH2-MIL-101(Cr)762 and ZnO.763

3.9.2 Metal-doped and metal–metal oxides. Ramasamy et al.104 fabricated an Ag-doped ZnO photocatalyst to study its performance as a photocatalyst in the visible-light region for the photocatalytic degradation of atenolol (and acetaminophen) in a water medium. The corresponding removal efficiencies were found to be about 70 and 91% for [ATL]int = [ACT]int: 5 mg L−1, pH: 8.5, time: 120 min, and Ag–ZnO: 1 g L−1. These findings also confirmed that the removal process takes place through the OH· pathway in the removal of the pollutants. Fe–TiO2 and Ag–TiO2 mediated visible-light photocatalysis removed atenolol from aqueous solution under optimum conditions by 75.5% (98 min) and 68.3% (120 min), respectively.764

Atenolol has been removed from domestic wastewater effluent using green-synthesized Fe (0–5%)-doped TiO2 (Fe–TiO2) under visible-light irradiation.765 These findings showed 85% removal of atenolol in the presence of Fe (2 wt%)–TiO2 after 105 min at solution pH 9, initial atenolol concentration of 10 mg L−1 and catalyst dose of 1.25 g L−1. The degradation of atenolol by visible-light-activated Fe–TiO2 was attributed to the cleavage of the ether bond, hydroxylation of the aromatic ring and oxidation of amine moieties. Alternatively, the enhanced photocatalytic activity for atenolol by Fe-doped TiO2 due to the reduced band gap of TiO2 cannot be ruled out.

Ag–TiO2 (Ag/Ti molar ratio: 2%) microtubes showed enhanced degradation of atenolol under UV-light irradiation (λ: 365 nm, power: 0.111 mW cm−2).766 Further investigations revealed Ag acting as a good photogenerated electron acceptor for photocatalysis. Cobalt-doped TiO2 nanoparticles (dose: 2.0 g L−1) exhibited about 90% photodegradation (ATL: 15 mg L−1, H2O2: 2.0 mL, pH: 2) of atenolol in 40 min under UV irradiation.767 The photodegradation of atenolol followed first-order kinetics, and the process involved the formation of hydroxyl free radicals and superoxide oxygen anions as active species.

3.9.3 Metal oxide composites. A Bi2O3/TiO2 composite was successfully synthesized by a solvothermal method and its photocatalytic performance was tested for the removal of atenolol removal from aqueous solution under UVC and visible-light irradiation.768 The investigations revealed the decomposition of atenolol to be better for Bi2O3/TiO2 (68.92%) than Bi2O3 (22.58%) after 60 minutes under optimum conditions (pH: 7, catalyst dosage: 400 mg L−1 and initial concentration of atenolol: 10 mg L−1). Stojanović et al.769 fabricated a TiO2/zeolite composite by a solid-state dispersion method and investigated the photocatalytic degradation of atenolol from an aqueous solution (pH ∼ 6.5) under simulated solar light. These findings indicated ∼94% and 88% degradation of atenolol after 70 min for ZSM-5 combined with P25 TiO2 and ZSM-5/TiO2 nanocrystals, respectively. Corchero et al.770 prepared Fe3O4@AgCl and Fe3O4@TiO2 nanocatalysts using an ionic liquid. Subsequent evaluation of their effectiveness as photocatalysts under UV light (30 min) showed the degradation of atenolol by 66.0% and 43.7%, respectively. The photocatalytic degradation of atenolol has also been reported using BiOCl@Fe3O4771 and immobilized titania/silica on glass slides.772
3.9.4 Graphitic material composite. A hydrothermally prepared graphene oxide–TiO2 (1.5 g L−1) composite showed 72% degradation of atenolol (25 ppm) solution (pH: 6) under visible-light irradiation after 1 h.773 The inclusion of graphene oxide in the composite facilitated enhanced electron–hole pair separation. The photocatalytic activities of immobilized graphene–TiO2774 and graphene oxide/ZnO composite775 have also been examined for the photocatalytic degradation of atenolol under UV and solar irradiation, respectively. A metal-free exfoliated g-C3N4 photocatalyst showed the efficient removal of atenolol from urban wastewater under visible light.776 In another study, carbon nitride modified by graphene quantum dots exhibited 86% photocatalytic degradation efficiency for atenolol, which still remained above 83% after five cycles.777
3.9.5 Heterojunctions and Z-scheme-based photocatalysts. Kumar et al.778 used recycled LiFePO4 from batteries in combination with B@C3N4 and CuFe2O4, which were harnessed as sustainable nanojunctions to study xenon-lamp-mediated atenolol degradation and showed 99.5% and 85.3% (60 min) degradation efficiency by B@C3N4/LiFePO4/CuFe2O4 and B@C3N4/LiFePO4/CuFe2O4 (30%) photocatalysts. Z-Scheme Y–Ag3PO4/CQDs/BiVO4 exhibited 90.9% degradation efficiency for atenolol under visible light (6 h) compared to Y–Ag3PO4 and BiVO4 (Fig. 24(a)).779 This could be attributed to an increase in the visible-light absorption and electron mediators as a result of the synergistic effect. The kinetic constant in the photocatalytic degradation of atenolol was found to be ∼2.8 times that of pristine Ag3PO4 in the presence of Y–Ag3PO4/CQDs/BiVO4 and a possible mechanism has also been proposed, as shown in Fig. 24(b). Both Y–Ag3PO4 and BiVO4 generated photogenerated carriers under visible-light illumination. CQDs not only increase the visible-light absorption of Y–Ag3PO4/CQDs/BiVO4 but also act as electron mediators. Simultaneously, oxygen defects caused by the doping of Y3+ into Ag3PO4 are a capture centre for photogenerated electrons to generate ·O2, inhibiting the recombination of photogenerated electron–hole pairs. In another study, a double Z-scheme rGO/CuFe2O4/CdS/Bi2S3 QDs nanoheterojunction exhibited ∼76.5% degradation of atenolol photo-Fenton-assisted photocatalytic degradation of atenolol in 360 min under visible-light irradiation.780 The degradation of atenolol was attributed to enhanced surface oxygen vacancies, the formation of OH· and h+ and the photo-Fenton reaction.
image file: d3lf00142c-f24.tif
Fig. 24 (a) Photocatalytic activities of photocatalysts for atenolol degradation of Ag3PO4, Y–Ag3PO4, BiVO4, Y–Ag3PO4/BiVO4, Y–Ag3PO4/CQDs and Y–Ag3PO4/CQDs/BiVO4 for the degradation of atenolol and (b) Z-scheme photocatalysis mechanism for atenolol degradation by Y–Ag3PO4/CQDs/BiVO4. Reproduced from ref. 779 with permission from Elsevier (2020).

Table 10 records data on the performance of different photocatalysts on removal of diclofenac from water under optimum conditions.

Table 10 Performance data on removal of atenolol in water in the presence of different photocatalysts
Photocatalyst Preparation method ATL Catalyst dose pH Light source Degradation (time) Rate constant
TiO2: mixed phase (source: Shandong Xiya Chemical Co)754 Commercial 18.77 μM 2 g L−1 7.6 UV-lamp (365 nm) and I0: 774 mW cm−2 100% (60 min) 0.064 min−1
TiO2 (75% A + 25% R) Degussa P25755 Commercial 10 mg L−1 250 mg L−1 8 Xenon-OP lamp (1 kW), I0: 272.3 W m−2 80% (120 min)
Degussa P25756 Commercial 10 mg L−1 1.0 g L−1 Natural solar irradiation 100% (400 kJ m−2)
Degussa TiO2 P25758 Commercial 37.6 mM 2.0 g L−1 (25 mL) 6.8 High-pressure Hg lamp (125 W), 365 nm, 31.3 mW m−2 ∼100% (60 min) 0.0570 min−1
Degussa P25 TiO2[thin space (1/6-em)]760 Commercial 37.6 μM 2.0 g L−1 7 UV light 100% (60 min)
TiO2 immobilized on the clinoptilolite nano particles support762 Dispersion method 10 mg L−1 (25 mL) 1.5 g L−1 UV lamp (80 W) 75% (60 min)
TiO2 immobilized on Salicylaldehyde-NH2-MIL 101 (Cr) support762 Dispersion method 10 mg L−1 (25 mL) 1.5 g L−1 Xenon lamp (100 W) 82% (60 min)
ZnO nanoparticles763 Synthetic method 20 mg L−1 10 mg L−1 7 9 W UVC lamp 100% (120 min)
Fe–TiO2[thin space (1/6-em)]764 Green method 5 mg L−1 (100 mL) 1005 mg L−1 8 Xenon arc lamp, 300 W, λ: 650 nm 71.2% (98 min)
Ag–TiO2[thin space (1/6-em)]764 Green method 5 mg L−1 (100 mL) 1065 mg L−1 8 Xenon arc lamp, 300 W, λ: 650 nm 65.7% (120 min)
Ag–ZnO microtubes104 Solution method 5 mg L−1 1 g L−1 8.5 W halogen lamp (300 W) 70.2% (120 min) 0.01 min−1
Fe–TiO2[thin space (1/6-em)]765 Green synthesis 10 mg L−1 1.25 g L−1 9 300 W halogen lamp 85% (105 min) 0.013 min−1
Ag–TiO2 microtubes (Ag/Ti molar ratio: 2%)/O3[thin space (1/6-em)]766 Calcination 20 mg L−1 0.2 g 9.11 Medium-pressure Hg lamp: 365 nm and 0.111 mW cm−2 92.23% (9 min) 0.3275 min−1
Co doped-TiO2 (H2O2: 2.0 mL L−1)767 Mixing followed by calcination 15 mg L−1 2.0 g L−1 2 UV (200 nm) 90% (40 min) 0.059 min−1, 1.75 × 10−4 g mg−1 min−1
Bi2O3/TiO2[thin space (1/6-em)]768 Solvothermal method 10 mg L−1 400 mg L−1 7 UVC (visible-light irradiation) 68.92% (60 min)
TiO2/zeolites769 Solid-state dispersion method 50 mg L−1 1 g L−1 (40 mL) 6.5 Lamp (Osram Vitalux (300 W)) ∼94% (70 min) 0.132 ± 0.001 min−1
TiO2@Fe3O4[thin space (1/6-em)]770 Mixing method 10 ppm 0.75 g L−1 5.5 Low-pressure Hg vapour lamp (UVC. l: 280 nm) 43.7% (30 min)
Fe3O4@AgCl770 Mixing method 10 ppm 0.75 g L−1 5.5 Low-pressure Hg vapour lamp (UVC. l: 280 nm) 66% (30 min)
Fe3O4@TiO2[thin space (1/6-em)]770 Mixing method 10 ppm 0.75 g L−1 5.5 Low-pressure Hg vapour lamp (UVC. l: 280 nm) 66% (30 min)
BiOCl@Fe3O4 with [PS]: 1.0 mM771 Precipitation process 2.5 mg L−1 0.1 g L−1 6.5 Xenon lamp (simulated sunlight): 500 W ∼99% (60 min) (5.34–6.04) × 10−2 min−1
Graphene oxide–TiO2[thin space (1/6-em)]773 Hydrothermal 25 ppm 1.5 g L−1 (150 mL) 6 1000 W xenon arc lamp, 750 mW cm−2 72% (60 min)
Y–Ag3PO4/CQDs/BiVO4[thin space (1/6-em)]779 Mixing method 10 mM (50 mL) 5 mg photocatalyst 250 W xenon lamp with UV cut-off filter, λ > 420 nm 90.9% (6 h) 0.50 h−1


4 Future scope and perspectives

Pharmaceutical pollutants found in water supplies through human and animal consumption of antibiotics, antipyretics, analgesics, etc. are considered potential hazards to the environment, humans and aquatic life.781 However, conventional wastewater treatment methods are ineffective in eliminating them completely. In view of this, the photocatalytic degradation of these pharmaceutical pollutants using semiconducting materials is considered an effective method.

An efficient semiconducting material acting as an efficient photocatalyst is guided by enhanced visible-light absorption, facilitating charge carrier migration and a reduced recombination rate. In view of this, TiO2, WO3, ZnO, Fe2O3, CdS, MoS2etc. are widely used photocatalysts for the photodegradation of pharmaceutical pollutants in water.23–39 However, the large band gaps of photogenerated charge carriers, i.e. rapid recombination rate (i.e., short lifetimes) of photogenerated charge carriers, instability in an aqueous medium, reusability of the photocatalyst and poor absorption ability for visible light, are a few drawbacks that limit the practical applications of metal oxide as photocatalysts. Therefore, increasing attention has been focused on achieving the effective separation of photogenerated charge carriers, improvements in the visible-light response and other factors782 through designing and constructing advanced light energy harvesting assemblies for environmental remediation.783 This problem has been overcome by modifying semiconducting metal oxides through doping, composite formation, immobilizing semiconducting materials on supports and heterojunction formation for the removal of drugs from contaminated water. In addition, the combination of these semiconducting metal oxides with carbon-based materials, such as activated carbon, biochar, carbon nanotubes, carbon dots, g-C3N4 and graphene, has also attracted a lot of attention in the removal of pharmaceutical pollutants present in wastewater. However, there are still several research gaps in the removal of antibiotics by photocatalysts. These future challenges are described below.

The expensive precursors used in the synthesis of metal oxides limit their large-scale application. Therefore, it is desirable to realize the simple, facile, affordable, low-cost synthesis of photocatalysts. The specific surface area,782 crystallite size,784 size, shape and overall structure785 of photocatalysts play important roles in the photocatalytic activity of emerging pollutants. This needs to be correlated with light trapping, charge separation and pollutant adsorption ability parameters under optimized operational conditions.

Carbon-based materials have also attracted significant interest in recent years due to their unique physicochemical, optical and electrical properties following band-gap tuning, composite formation and heterojunction construction, etc.40–50 The enhanced photo-efficiency of the corresponding nanocomposites is ascribed to improvement in visible-range absorption, fast charge carrier migration and reduced recombination rate. However, their choices are limited to batch experiments at the laboratory scale rather than the pilot scale. As a result, there is a gap between on-going research and its application.

The literature revealed considerable interest in investigating the photocatalytic degradation of individual pharmaceutical pollutants in water. However, wastewater could contain complex pollutant mixtures, including other organic and inorganic species originating from heavy metals, dyes, personal care products, pesticides and other sources.757,786–794 This can affect the degradation process for pharmaceutical pollutants through interference and matrix effects. Therefore, attention also needs to be focused on developing photocatalysts capable of simultaneously removing pharmaceuticals even in the presence of other pollutants/interfering substances in the wastewater. Recovery, reusability, and stability remain other issues in the development of high-performing photocatalysts in wastewater treatment. Toxicity assessment is considered to be one of important parameters in the treatment of wastewater by photocatalysis.795 This could be ascribed to the formation of carcinogenic secondary metabolites due to the incomplete mineralization of targeted contaminants.

Nanomaterial-based photocatalysts have shown great promise due to their superior adsorptive and photocatalytic properties in the removal of pharmaceutical pollutants.51,57 In this regard, leaching of toxic components could adversely affect the quality of the water environment. This aspect remains a matter of great concern and as a consequence, extensive investigations are needed to fully understand the role of various photocatalyst nanoparticles and their toxicity risks in aquatic environments.796,797 Therefore, it remains challenging to recover and separate the nanoparticle-based photocatalysts invariably used in water treatment. Recently, this difficulty has been overcome by immobilizing the photocatalysts on various support materials. Therefore, in the future innovations will be needed for effective, eco-friendly, and sustainable immobilization techniques for the separation/recovery and reuse of photocatalytic materials. Existing research has also invariably focused on laboratory-scale photocatalysis in the degradation of emerging pharmaceutical pollutants without much implementation in real water systems. More studies need to be focused at the pilot and industrial scale levels for its commercialization. The fabrication of economical, environmentally friendly and effective photocatalysts taking into account many of these aspects remains a major challenge in this field.

5 Conclusions

Antibiotics have been invariably used in different fields, such as the medical field, agriculture, and veterinary medicine for the purpose of killing or preventing bacterial growth. However, the presence of these pharmaceutical pollutants on entering surface water and groundwater are a potential threat to human and marine lives and need to be eliminated. Considering this, various conventional processes have been developed for the removal of these pharmaceutical pollutants. However, their choice is limited due to their high cost as well as incomplete elimination of contaminants from the contaminated water.

In view of this, the current review highlights recent advances in the applications of different photocatalysts to the removal of emerging pharmaceutical pollutants in wastewater. As a result, the performance of several metal oxides, carbonaceous materials, composites including surface modification, doping with metals/nonmetals, heterojunction formation, and immobilization using support materials, homo- or hetero-materials composed of two or more inorganic phases, inorganic semiconductors coupled with carbon-based materials, inorganic semiconductors hybridized with 2D materials as excellent photocatalysts have been reviewed to find out the optimum removal efficiency for the pollutants (acetaminophen, amoxicillin, sulfamethoxazole, acetaminophen, norfloxacin, ciprofloxacin, tetracycline, diclofenac and atenolol) in water. However, secondary pollution produced by the formation of by-products during the photocatalytic process, leaching of dopants/active components of the photocatalysts, and the generation of excess CO2 during the photocatalysis process are additional challenges that need to be addressed in future. Further, most of these findings are reported on the laboratory scale, and real-world and industrial-scale applications have yet to be fully realized. The further development of low-cost, robust photocatalysts utilizing semiconductors and renewable visible/solar light to solve both the world crises of energy supply and environmental pollution remains a pressing demand for industrial application.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

The author began working on this review article while still a Professor in the Department of Chemistry, Indian Institute of Technology, Kharagpur and thus, remains very appreciative for making this effort possible. Author also expresses his heartfelt thanks to Professor Ashok Kumar Gupta, Department of Civil Engineering, Indian Institute of Technology, Kharagpur and his research scholars Brahma Gupta, Vishal Parida, Adarsh Singh and Akash Rawat for the valuable interactions. The table of content figure was drawn with the help of Biorender and acknowledged. Author also thanks Dr. Kunal Manna and Dr. Ayon Karmakar for their help in arranging copyright permissions and the Figures, respectively.

References

  1. https://www.unesco.org/en/articles/imminent-risk-global-water-crisis-warns-un-world-water-development-report-2023 .
  2. I. Bashir, F. A. Lone, R. A. Bhat, S. A. Mir, Z. A. Dar and S. A. Dar, in Bioremediation and biotechnology sustainable approaches to pollution degradation, ed. K. R. Hakeem, R. A. Bhat and H. Qadri, Springer, 2020, pp. 1–26,  DOI:10.1007/978-3-030-35691-0_1.
  3. https://reliefweb.int/report/world/wastewater-resource-may-2022 .
  4. S. Kundu and A. K. Gupta, Chem. Eng. J., 2006, 122, 93–106 CrossRef CAS.
  5. S. Dutta, B. Gupta, S. K. Srivastava and A. K. Gupta, Mater. Adv., 2021, 2, 4497–4531 RSC.
  6. S. Dutta, S. K. Srivastava and A. K. Gupta, Mater. Adv., 2021, 2, 2431 RSC.
  7. S. Senapati, S. K. Srivastava and S. B. Singh, Nanoscale, 2012, 4, 6604–6612 RSC.
  8. P. A. Kobielska, A. J. Howarth, O. K. Farha and S. Nayak, Coord. Chem. Rev., 2018, 358, 92–102 CrossRef CAS.
  9. S. Ayoob and A. K. Gupta, Crit. Rev. Environ. Sci. Technol., 2006, 36, 433–487 CrossRef CAS.
  10. A. Singh, D. Saidulu, A. K. Gupta and V. Kubsad, J. Environ. Chem. Eng., 2022, 10, 109012 CrossRef CAS.
  11. B. Gupta, R. S. Ambekar, R. M. Tromer, P. S. Ghosal, R. Sinha, A. Majumder, P. Kumbhakar, P. M. Ajayan, D. S. Galvao, A. K. Gupta and C. S. Tiwary, RSC Adv., 2021, 11, 19788–19796 RSC.
  12. B. S. Rathi, P. S. Kumar and D.-V. N. Vo, Sci. Total Environ., 2021, 797, 149134 CrossRef CAS PubMed.
  13. M. Hejna, D. Kapuścińska and A. Aksmann, Int. J. Environ. Res. Public Health, 2022, 19, 7717 CrossRef CAS PubMed.
  14. K. Samal, S. Mahapatra and M. H. Ali, Energy Nexus, 2022, 6, 100076 CrossRef CAS.
  15. A. C. Duarte, S. Rodrigues, A. Afonso, A. Nogueira and P. Coutinho, Pharmaceuticals, 2022, 15, 393 CrossRef CAS PubMed.
  16. T. Zhang, K. Lv, Q. Lu, L. Wang and X. Liu, J. Environ. Sci., 2021, 104, 415–429 CrossRef CAS PubMed.
  17. J. O'Neill, Tackling Drug-Resistant Infections Globally: Final Report and Recommendations, 2016, Available online: https://amr-review.org/sites/default/files/160518_Final%20paper_with%20cover.pdf.
  18. S. I. Polianciuc, A. E. Gurzău, B. Kiss, M. G. Ştefan and F. Loghin, Med. Pharm. Rep., 2020, 93, 231–240 Search PubMed.
  19. V. Vinayagam, S. Murugan, R. Kumaresan, M. Narayanan, M. Sillanpää, D. V. N. Vo, O. S. Kushwaha, P. Jenis, P. Potdar and S. Gadiya, Chemosphere, 2022, 300, 134597 CrossRef CAS PubMed.
  20. V. K. Parida, S. K. Srivastava, S. Chowdhury and A. K. Gupta, Chem. Eng. J., 2023, 472, 144969 CrossRef CAS.
  21. A. A. Aryee, R. Han and L. Ou, J. Cleaner Prod., 2022, 368, 133140 CrossRef CAS.
  22. B. Sruthi, S. Janani and S. S. Khan, Sep. Purif. Technol., 2021, 279, 119709 CrossRef.
  23. D. Kanakaraju, B. D. Glass and M. Oelgemoller, Environ. Chem. Lett., 2014, 12, 27–47 CrossRef CAS.
  24. M. Minella, D. Fabbri, P. Calza and C. Minero, Curr. Opin. Green Sustainable Chem., 2017, 6, 11–17 CrossRef.
  25. R. Karuppannan, S. Mohan and T.-O. Do, in Nanostructured materials for environmental applications, ed. S. Balakumar, V. Keller and M. V. Shankar, Springer Nature Switzerland AG, 2021, pp. 85–112 Search PubMed.
  26. A. Krishnan, A. Swarnalal, D. Das, M. Krishnan, V. S. Saji and S. M. A. Shibli, J. Environ. Sci., 2024, 139, 389–417 CrossRef PubMed.
  27. N. Roy, S. A. Alex, N. Chandrasekaran, A. Mukherjee and K. Kannabiran, J. Environ. Chem. Eng., 2021, 9, 104796 CrossRef CAS.
  28. R. Gusain, K. Gupta, P. Joshi and O. P. Khatri, Adv. Colloid Interface Sci., 2019, 272, 102009 CrossRef CAS PubMed.
  29. S. Gautam, H. Agrawal, M. Thakur, A. Akbari, H. Sharda, R. Kaur and M. Amini, J. Environ. Chem. Eng., 2020, 8, 103726 CrossRef CAS.
  30. T. Velempini, E. Prabakaran and K. Pillay, Mater. Today Chem., 2021, 19, 100380 CrossRef CAS.
  31. R. Krakowiak, J. Musial, P. Bakun, M. Spychała, B. Czarczynska-Goslinska, D. T. Mlynarczyk, T. Koczorowski, L. Sobotta, B. Stanisz and T. Goslinski, Appl. Sci., 2021, 11, 8674 CrossRef CAS.
  32. K. P. Gopinath, N. V. Madhav, A. Krishnan, R. Malolan and G. Rangarajan, J. Environ. Manage., 2020, 270, 110906 CrossRef CAS PubMed.
  33. D. Chen, Y. Cheng, N. Zhou, P. Chen, Y. Wang, K. Li, S. Huo, P. Cheng, P. Peng, R. Zhang, L. Wang, H. Liu, Y. Liu and R. Ruan, J. Cleaner Prod., 2020, 268, 121725 CrossRef CAS.
  34. A. Mirzaei, Z. Chen, F. Haghighat and L. Yerushalmi, Sustain. Cities Soc., 2016, 27, 407–418 CrossRef.
  35. P. Kar, K. Shukla, P. Jain, G. Sathiyan and R. K. Gupta, Nano Mater. Sci., 2021, 3, 25–46 CrossRef CAS.
  36. K. S. Varma, R. J. Tayade, K. J. Shah, P. A. Joshi, A. D. Shukla and V. G. Gandhi, Water-Energy Nexus, 2020, 3, 46–61 CrossRef.
  37. G. Ren, H. Han, Y. Wang, S. Liu, J. Zhao, X. Meng and Z. Li, Nanomaterials, 2021, 11, 1804 CrossRef CAS PubMed.
  38. R. M. S. Sendão, J. C. G. Esteves da Silva and L. P. da Silva, Chemosphere, 2022, 301, 134731 CrossRef PubMed.
  39. L. G. Devi and R. Kavitha, Appl. Surf. Sci., 2016, 360, 601–622 CrossRef.
  40. J. Ge, Y. Zhang and S.-J. Park, Materials, 2019, 12, 1916 CrossRef CAS PubMed.
  41. R. Ma, Y. Xue, Q. Ma, Y. Chen, S. Yuan and J. Fan, Nanomaterials, 2022, 12, 4045 CrossRef CAS PubMed.
  42. T. H. Pham, N. M. Viet, P. T. T. Hoai, S. H. Jung and T. Y. Kim, Environ. Res., 2023, 231, 116246 CrossRef CAS PubMed.
  43. K. K. Gangu, S. Maddila and S. B. Jonnalagadda, Sci. Total Environ., 2019, 646, 1398–1412 CrossRef CAS PubMed.
  44. R. B. González-González, A. Sharma, R. Parra-Saldívar, R. A. Ramirez-Mendoza, M. Bilal and H. M. N. Iqbal, J. Hazard. Mater., 2022, 423, 127145 CrossRef PubMed.
  45. M. Abdullah, J. Iqbal, M. S. U. Rehman, U. Khalid, F. Mateen, S. N. Arshad, A. G. Al-Sehemi, H. Algarni, O. A. Al-Hartomy and T. Fazal, Chemosphere, 2023, 317, 137834 CrossRef CAS PubMed.
  46. M.-F. Li, Y.-G. Liu, G.-M. Zeng, N. Liu and S.-B. Liu, Chemosphere, 2019, 226, 360–380 CrossRef CAS PubMed.
  47. M. Minale, Z. Gu, A. Guadie, D. M. Kabtamu, Y. Li and X. Wang, J. Environ. Manage., 2020, 276, 111310 CrossRef CAS PubMed.
  48. O. C. Olatunde and D. C. Onwudiwe, Int. J. Environ. Res. Public Health, 2021, 18, 1529 CrossRef CAS PubMed.
  49. B. A. Bhanvase, T. P. Shende and S. H. Sonawane, Environ. Technol. Rev., 2017, 6, 1–14 CrossRef CAS.
  50. W. W. Anku, E. M. Kiarii, R. Sharma, G. M. Joshi, S. K. Shukla and P. P. Govender, in A New Generation Material Graphene: Applications in Water Technology, ed. M. Naushad, Springer International Publishing AG, part of Springer Nature, 2019, pp. 187–208 Search PubMed.
  51. V. K. Parida, S. K. Srivastava, A. K. Gupta and A. Rawat, Mater. Express, 2023, 13, 1–38 CrossRef CAS.
  52. A. Abbasnia, A. Zarei, M. Yeganeh, H. R. Sobhi, M. Cholami and A. Esrafili, Inorg. Chem. Commun., 2022, 145, 109959 CrossRef CAS.
  53. S. S. Imam, R. Adnan and N. H. M. Kaus, Toxicol. Environ. Chem., 2018, 100, 518–539 CrossRef.
  54. S. Shurbaje, P. T. Huong and T. M. Altahtamounti, Catalysts, 2021, 11, 437 CrossRef.
  55. https://pubchem.ncbi.nlm.nih.gov/ .
  56. Z. L. Wang, J. Phys.: Condens.Matter, 2004, 16, R829–R858 CrossRef CAS.
  57. A. Pattnaik, J. N. Sahu, A. K. Poonia and P. Ghosh, Chem. Eng. Res. Des., 2023, 190, 667–686 CrossRef CAS.
  58. Z. Wu, Z. Xiong and B. Lai, Environmental Functional Materials, 2022, 1, 298–315 CrossRef.
  59. Y. Wang, Z. Lu, Z. Zhu, X. Zhao, N. Gao, D. Wang, Z. Hua, Y. Yan, P. Huoa and M. Song, RSC Adv., 2016, 6, 51877–51887 RSC.
  60. S. Heidari, M. Haghighi and M. Shabani, J. Cleaner Prod., 2020, 259, 120679 CrossRef CAS.
  61. P. Wang, B. Huang, Y. Dai and M.-H. Whangbo, Phys. Chem. Chem. Phys., 2012, 14, 9813–9825 RSC.
  62. A. Kumar, P. Choudhary, A. Kumar, P. H. C. Camargo and V. Krishnan, Small, 2022, 18, 2101638 CrossRef CAS PubMed.
  63. E. S. Araújo, M. F. G. Pereira, G. M. G. da Silva, G. F. Tavares, C. Y. B. Oliveira and P. M. Faia, Toxics, 2023, 11, 658 CrossRef PubMed.
  64. A. Moharana, A. Kumar, A. Thakur, D.-V. N. Vo, A. Sharma and D. Kumar, in Nanostructured photocatalysts from Fundamental to practical applications, 2021, pp. 145–167,  DOI:10.1016/B978-0-12-823007-7.00010-9.
  65. S. Biswas and A. Pal, Catalysts, 2023, 13, 925 CrossRef CAS.
  66. B. Samir, N. Bouazizi, P. N. Fotsing, J. Cosme, V. Marquis, G. L. Dotto, F. L. Derf and J. Vieillard, Appl. Sci., 2023, 13, 8074 CrossRef CAS.
  67. S. Kurwadkar, T. V. Hoang, K. Malwade, S. R. Kanel, W. F. Harper Jr and G. Struckhof, Nanotechnol. Environ. Eng., 2019, 4, 12 CrossRef CAS.
  68. Y. Tang, X. Liu, C. Ma, M. Zhou, P. Huo, L. Yu, J. Pan, W. Shia and Y. Yan, New J. Chem., 2015, 39, 5150–5160 RSC.
  69. B. Partoens and F. M. Peeters, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 74, 075404 CrossRef.
  70. Y. Ren, D. Zeng and W.-J. Ong, Chin. J. Catal., 2019, 40, 289–319 CrossRef CAS.
  71. X. He, T. Kai and P. Ding, Environ. Chem. Lett., 2021, 19, 4563–4601 CrossRef CAS PubMed.
  72. S. Obregón, M. A. Ruíz-Gómez, V. Rodríguez-González, A. Vázquez and D. B. Hernández-Uresti, Mater. Sci. Semicond. Process., 2020, 113, 105056 CrossRef.
  73. W. Dai, L. Jiang, J. Wang, Y. Pu, Y. Zhu, Y. Wang and B. Xiao, Chem. Eng. J., 2020, 397, 125476 CrossRef CAS.
  74. M. Miceli, P. Frontera, A. Macario and A. Malara, Catalysts, 2021, 11, 591 CrossRef CAS.
  75. V. Wongso, H. K. Chung, N. S. Sambudi, S. Sufian, B. Abdullah, M. D. H. Wirzal and W. L. Ang, J. Photochem. Photobiol., A, 2020, 394, 112436 CrossRef CAS.
  76. A. H. Zyoud, A. Zubi, S. Hejjawi, S. H. Zyoud, M. H. Helal, S. H. Zyoud, N. Qamhieh, A. R. Hajamohideen and H. S. Hilal, J. Environ. Chem. Eng., 2020, 8, 104038 CrossRef CAS.
  77. C. A. Aguilar, C. Montalvo, J. G. Ceron and E. Moctezuma, Int. J. Environ. Res., 2011, 5, 1071–1078 CAS.
  78. C. A. Aguilar, C. Montalvo, B. B. Zermeño, R. M. Cerón, J. G. Cerón, F. Anguebes and M. A. Ramírez, Int. J. Environ. Sci. Technol., 2019, 16, 843–852 CrossRef CAS.
  79. R. Trujillano, V. Rives and I. García, Molecules, 2022, 27, 2904 CrossRef CAS PubMed.
  80. A. Marizcal-Barba, J. A. Sanchez-Burgos, V. Zamora-Gasga and A. Perez Larios, Photochem, 2022, 2, 225–236 CrossRef CAS.
  81. C. J. Lin, W. T. Yang, C.-Y. Chou and S. Y. H. Liou, Chemosphere, 2016, 152, 490–495 CrossRef CAS PubMed.
  82. X. Zhang, F. Wu, X. W. Wu, P. Chen and N. Deng, Materials, 2008, 157, 300–307 CAS.
  83. S. A. Lozano-Morales, G. Morales, M. Á. L. Zavala, A. Arce-Sarria and F. Machuca-Martínez, Processes, 2019, 7, 319 CrossRef CAS.
  84. N. Kaneva, A. Bojinova, K. Papazova, D. Dimitrov, K. Zaharieva, Z. Cherkezova-Zheleva and A. Eliyas, Arch. Pharmacal Res., 2016, 39, 1418–1425 CrossRef CAS PubMed.
  85. R. Katal, M. H. D. A. Farahani and H. Jiangyong, Sep. Purif. Technol., 2020, 230, 115859 CrossRef CAS.
  86. M. Cantarella, A. D. Mauro, A. Gulino, L. Spitaleri, G. Nicotra, V. Privitera and G. Impellizzeri, Appl. Catal., B, 2018, 238, 509–517 CrossRef CAS.
  87. M. Jiménez-Salcedo, M. Monge and M. T. Tena, Photochem. Photobiol. Sci., 2022, 21, 337–347 CrossRef PubMed.
  88. O. Nasr, O. Mohamed, A.-S. Al-Shirbini and A.-M. Abdel-Wahab, J. Photochem. Photobiol., A, 2019, 374, 185–193 CrossRef CAS.
  89. R. Katal, S. M. Panah, M. Saeedikhani, M. Kosari, C. C. Sheng, O. S. Leong, G. Xiao and H. Jiangyong, Adv. Mater. Interfaces, 2018, 5, 1801440 CrossRef.
  90. T. H. Yu, W. Y. Cheng, K.-J. Chao and S. Y. Lu, Nanoscale, 2013, 5, 7356–7360 RSC.
  91. C. A. Huerta-Aguilar, A. A. Ramírez-Alejandre, P. Thangarasu, J. A. Arenas-Alatorre, I. A. Reyes-Dominguez and M. de la L. Corea, Catal. Sci. Technol., 2019, 9, 3066 RSC.
  92. S. Sayegh, F. Tanos, A. Nada, G. Lesage, F. Zaviska, E. Petit, V. Rouessac, I. Iatsunskyi, E. Coy, R. Viter, D. Damberga, M. Weber, A. Razzouk, J. Stephan and M. Bechelany, Dalton Trans., 2022, 51, 2674–2695 RSC.
  93. V. Y. Safitri, A. Santoni, D. V. Wellia, K. Khoiriah and S. Safni, Mol. Ther., 2017, 12, 189–195 CAS.
  94. Y. A. Shaban and H. M. Fallata, Res. Chem. Intermed., 2019, 45, 2529–2547 CrossRef CAS.
  95. W. L. da Silva, M. A. Lansarin, J. H. Z. dos Santos and F. Silveira, Water Sci. Technol., 2016, 74, 2370–2383 CrossRef PubMed.
  96. L. Rimoldi, D. Meroni, E. Falletta, A. M. Ferretti, A. Gervasini, G. Cappelletti and S. Ardizzone, Appl. Surf. Sci., 2017, 424, 198–205 CrossRef CAS.
  97. Z. Khani, D. Schieppati, C. L. Bianchi and D. C. Boffito, Catalysts, 2019, 9, 949 CrossRef CAS.
  98. A. M. S. Asadi and M. Malakootian, J. Mater. Sci.: Mater. Electron., 2019, 30, 14878–14889 CrossRef.
  99. M. L. P. Dalida, K. M. S. Amer, C.-C. Su and M.-C. Lu, Environ. Sci. Pollut. Res., 2014, 21, 1208–1216 CrossRef CAS PubMed.
  100. M. Abid, E. Makhoul, F. Tanos, I. F. Iatsunskyi, E. Coy, G. Lesage, M. Cretin, D. Cornu, A. B. H. Amara and M. Bechelany, Membranes, 2023, 13, 204 CrossRef CAS PubMed.
  101. M. D. G. de Luna, J. C.-T. Lin, M. J. N. Gotostos and M.-C. Lu, Sustainable Environ. Res., 2016, 26, 161–167 CrossRef CAS.
  102. A. Alzamly, F. Hamed, T. Ramachandran, M. Bakiro, M. Bakiro, S. H. Ahmed, S. Mansour, S. Salem, K. Abdul al, N. S. Al Kaabi, M. Meetani and A. Khalee, J. Water Reuse Desalin., 2019, 9, 31–46 CrossRef CAS.
  103. H. A. Abbas, R. A. Nasr, R. N. Vannier and T. S. Jamil, J. Environ. Sci., 2022, 112, 331–342 CrossRef CAS PubMed.
  104. B. Ramasamy, J. Jeyadharmarajan and P. Chinnaiyan, Environ. Sci. Pollut. Res., 2021, 28, 39637–39647 CrossRef CAS PubMed.
  105. V. H.-T. Thi and B.-K. Lee, Mater. Res. Bull., 2017, 96, 171–182 CrossRef CAS.
  106. R. A. Abri, F. A. Marzouqi, A. T. Kuvarega, M. A. Meetani, S. M. Z. Al Kindy, S. Karthikeyan, Y. Kim and R. Selvaraj, J. Photochem. Photobiol., A, 2019, 384, 112065 CrossRef.
  107. D. R. Kumar, K. S. Ranjith, Y. Haldorai, A. Kandasami and R. T. R. Kumar, ACS Omega, 2019, 4, 11973–11979 CrossRef PubMed.
  108. O. F. S. Khasawneh, P. Palaniandy, M. Ahmadipour, H. Mohammadi, M. Razak and B. Hamdan, J. Environ. Chem. Eng., 2021, 9, 104921 CrossRef CAS.
  109. O. F. S. Khasawneh, P. Palaniandy and L. P. Teng, MethodsX, 2019, 6, 2735–2743 CrossRef PubMed.
  110. H. Shahabi, A. Allahrasani and A. Naghizadeh, Desalin. Water Treat., 2019, 149, 164–170 CrossRef CAS.
  111. P. M. Álvareza, J. Jaramillo, F. López-Piñero and P. K. Plucinski, Appl. Catal., B, 2010, 100, 338–345 CrossRef.
  112. B. Czech and K. Tyszczuk-Rotko, Sep. Purif. Technol., 2018, 206, 343–355 CrossRef CAS.
  113. T. A. Fernandes, S. G. Mendo, L. P. Ferreira, N. R. Neng, M. C. Oliveira, A. Gil, M. D. Carvalho, O. C. Monteiro, J. M. F. Nogueira and M. J. Calhorda, Environ. Sci. Pollut. Res., 2021, 28, 17228–17243 CrossRef CAS PubMed.
  114. R. Mu, Y. Ao, T. Wu, C. Wang and P. Wang, J. Hazard. Mater., 2020, 382, 121083 CrossRef CAS PubMed.
  115. J. H. F. Chau, C. W. Lai, B. F. Leo, J. C. Juan and M. R. Johan, Catal. Commun., 2022, 163, 106396 CrossRef CAS.
  116. S. Basha, D. Keane, K. Nolan, M. Oelgemöller, J. Lawler, J. M. Tobin and A. Morrissey, Environ. Sci. Pollut. Res., 2015, 22, 2219–2230 CrossRef CAS PubMed.
  117. A.-M. Abdel-Wahab, A.-S. Al-Shirbini, O. Mohamed and O. Nasr, J. Photochem. Photobiol., A, 2017, 347, 186–198 CrossRef CAS.
  118. N. Jallouli, K. Elghniji, H. Trabelsi and M. Ksibi, Arabian J. Chem., 2017, 10, S3640–S3645 CrossRef CAS.
  119. L. Yanyan, T. A. Kurniawan, Z. Ying, A. B. Albadarin and G. Walker, J. Mol. Liq., 2017, 243, 761–770 CrossRef.
  120. C.-T. Chang, J.-J. Wang, T. Ouyang, Q. Zhang and Y.-H. Jing, Mater. Sci. Eng., B, 2015, 196, 53–60 CrossRef CAS.
  121. N. Miranda-García, S. Suárez, M. Ignacio Maldonado, S. Malato and B. Sánchez, Catal. Today, 2014, 230, 27–34 CrossRef.
  122. V. Vaiano, M. Matarangolo and O. Sacco, Chem. Eng. J., 2018, 350, 703–713 CrossRef CAS.
  123. S. Behravesh, N. Mirghaffari, A. A. Alemrajabi, F. Davar and M. Soleimani, Environ. Sci. Pollut. Res., 2020, 27, 26929–26942 CrossRef CAS PubMed.
  124. G. Fan, H. Peng, J. Zhang, X. Zheng, G. Zhu, S. Wang and L. Hong, Catal. Sci. Technol., 2018, 8, 5906–5919 RSC.
  125. T. A. Kurniawan, L. Yanyan, T. Ouyang, A. B. Albadarin and G. Walker, Mater. Sci. Semicond. Process., 2018, 73, 42–50 CrossRef CAS.
  126. G. Fan, X. Zheng, J. Luo, H. Peng, H. Lin, M. Bao, L. Hong and J. Zhou, Chem. Eng. J., 2018, 351, 782–790 CrossRef CAS.
  127. F. Al Marzouqi, R. Selvaraj and Y. Kim, Mater. Res. Express, 2019, 6, 125538 CrossRef CAS.
  128. A. Smýkalová, B. Sokolová, K. Foniok, V. Matejka and P. Praus, Nanomaterials, 2019, 9, 1194 CrossRef PubMed.
  129. M. Z. A. Warshagha and M. Muneer, Int. J. Environ. Anal. Chem., 2022, 102, 6339–6358 CrossRef.
  130. M. Kohantorabi, G. Moussavi, P. Oulego and S. Giannakis, Sep. Purif. Technol., 2022, 299, 121723 CrossRef CAS.
  131. R. R. Solís, M. A. Quintana, M. Á. Martín-Lara, A. Pérez, M. Calero and M. J. Muñoz-Batista, Int. J. Mol. Sci., 2022, 23, 12871 CrossRef PubMed.
  132. S. Gupta, J. Gandhi, S. Kokate, L. G. Raikar, V. G. Kopuri and H. Prakash, Heliyon, 2023, 9, e16450 CrossRef CAS PubMed.
  133. A. H. C. Khavar, G. Moussavi and A. R. Mahjoub, Appl. Surf. Sci., 2018, 440, 963–973 CrossRef.
  134. B. Palas, G. Ersöz and S. Atalay, Chem. Eng. Sci., 2021, 242, 116593 CrossRef CAS.
  135. J. Zhu, Z. Zhu, H. Zhang, H. Lu, W. Zhang, Y. Qiu, L. Zhuc and S. Küppers, Appl. Catal., B, 2018, 225, 550–562 CrossRef CAS.
  136. H. Tao, X. Liang, Q. Zhang and C.-T. Chang, Appl. Surf. Sci., 2015, 324, 258–264 CrossRef CAS.
  137. E. C. Umejuru, E. Prabakaran and K. Pillay, ACS Omega, 2021, 6, 11155–11172 CrossRef CAS PubMed.
  138. C. Gomez-Solis, R. Mendoza, J. F. Rios-Orihuela, G. Robledo-Trujillo, L. A. Diaz-Torres, J. Oliva and V. Rodriguez-Gonzalez, J. Environ. Manage., 2021, 290, 112665 CrossRef CAS PubMed.
  139. L. Allagui, B. Chouchene, T. Gries, G. Medjahdi, E. Girot, X. Framboisier, A. B. H. Amara, K. Balan and R. Schneider, Appl. Surf. Sci., 2019, 490, 580–591 CrossRef CAS.
  140. S. Sravya, D. R. Devi, N. Belachew, K. Eswara Rao and K. Basavaiah, RSC Adv., 2021, 11, 12030 RSC.
  141. Y. L. Wang, M. Penas-Garzon, J. J. Rodriguez, J. Bedia and C. Belver, Chem. Eng. J., 2022, 446, 137229 CrossRef CAS.
  142. L. K. B. Paragas, V. D. Dang, R. S. Sahu, S. Garcia-Segura, M. D. G. de Luna, J. A. I. Pimentel and R.-A. Doong, Sep. Purif. Technol., 2021, 272, 117567 CrossRef CAS.
  143. X. Du, X. Bai, L. Xu, L. Yang and P. Jin, Chem. Eng. J., 2020, 384, 123245 CrossRef CAS.
  144. M. Z. A. Warshagha and M. Muneer, Environ. Nanotechnol., Monit. Manage., 2022, 18, 100728 CAS.
  145. Z. Ren, Z. Zhao, Z. Yang, B. Cheng and X. Yang, Nano, 2021, 16, 2150100 CrossRef CAS.
  146. M. Noorisepehr, K. Ghadirinejad, B. Kakavandi, A. R. A. Esfahani and A. Asadi, Chemosphere, 2019, 232, 140–151 CrossRef CAS PubMed.
  147. S. Moradi, A. A. Isari, F. Hayati, R. Rezaei, R. R. Kalantary and B. Kakavandi, Chem. Eng. J., 2021, 414, 128618 CrossRef CAS.
  148. I. Gozlan, A. Rotstein and D. Avisar, Chemosphere, 2013, 91, 985–992 CrossRef CAS PubMed.
  149. K. D. Radosavljević, A. V. Golubović, M. M. Radišić, A. R. Mladenović, D. Ž. Mijin and S. D. Petrović, Chem. Ind. Chem. Eng. Q., 2017, 23, 187–195 CrossRef.
  150. M. Fazilati, Desalin. Water Treat., 2019, 169, 222–231 CrossRef CAS.
  151. J. H. O. S. Pereira, A. C. Reis, O. C. Nunes, M. T. Borges, V. J. P. Vilar and R. A. R. Boaventura, Environ. Sci. Pollut. Res., 2014, 21, 1292–1303 CrossRef CAS PubMed.
  152. B. Ambrosetti, L. Campanella and R. Palmisano, J. Environ. Sci. Eng. A, 2015, 4, 273–281 Search PubMed.
  153. E. S. Elmolla and M. Chaudhuri, Desalination, 2010, 252, 46–52 CrossRef CAS.
  154. R. Palmisano, L. Campanella and B. Ambrosetti, J. Environ. Anal. Chem., 2015, 2, 1000143 Search PubMed.
  155. D. Klauson, J. Babkina, K. Stepanova, M. Krichevskaya and S. Preis, Catal. Today, 2010, 151, 39–45 CrossRef CAS.
  156. F. S. Moosavi and T. Tavakoli, Environ. Sci. Pollut. Res., 2016, 23, 23262–23270 CrossRef CAS PubMed.
  157. D. Dimitrakopoulou, I. Rethemiotaki, Z. Frontistis, N. P. Xekoukoulotakis, D. Venieri and D. Mantzavinos, Environ. Manage., 2012, 98, 168–174 CAS.
  158. N. Ellepola and G. Rubasinghege, Environments, 2022, 9, 77 CrossRef PubMed.
  159. A. F. Alkaim and A. M. Aljobree, Int. J. Adv. Sci. Technol., 2020, 29, 5480–5487 Search PubMed.
  160. F. Moosavi, C. Cheng, T. T. Gheinani, M. Traore, A. Kanaev and M. Nikravech, Chem. Eng. Trans., 2019, 73, 175–180 Search PubMed.
  161. E. S. Elmolla and M. Chaudhuri, J. Hazard. Mater., 2010, 173, 445–449 CrossRef CAS PubMed.
  162. S. Pourmoslemi, A. Mohammadi, F. Kobarfard and N. Assi, Desalin. Water Treat., 2016, 57, 14774–14784 CrossRef CAS.
  163. K. M. M. Al-zobai, A. A. Hassan and N. O. Kariem, Solid State Technol., 2020, 63, 1–9 Search PubMed.
  164. B. A. Utami, H. Sutanto, I. Alkian, F. Sa'Adah and E. Hidayanto, Cogent Eng., 2022, 9, 2119534 CrossRef.
  165. T. T. Nguyen, S. N. Nam, J. Son and J. Oh, Sci. Rep., 2019, 9, 9349 CrossRef PubMed.
  166. D. Balarak and F. K. Mostafapour, Indones. J. Chem., 2019, 19, 211–218 CrossRef CAS.
  167. R. Mohammadi, B. Massoumi and M. Rabani, Int. J. Photoenergy, 2012, 514856,  DOI:10.1155/2012/514856.
  168. R. Mohammadi, B. Massoumi and H. Eskandarloo, Desalin. Water Treat., 2015, 53, 1995–2004 CrossRef CAS.
  169. E. T. Wahyuni, P. Y. Yulikayani and N. H. Aprilita, J. Mater. Environ. Sci., 2020, 11, 670–683 CAS.
  170. N. Olama, M. Dehghani and M. Malakootian, Appl. Water Sci., 2018, 8, 97 CrossRef.
  171. Lalliansanga, D. Tiwari, S.-M. Lee and D.-J. Kim, Environ. Res., 2022, 210, 112914 CrossRef CAS PubMed.
  172. Y. T. Gaim, S. M. Yimanuh and Z. G. Kidanu, J. Compos. Sci., 2022, 6, 317 CrossRef CAS.
  173. M. F. R. Samsudin, F. Maeght, N. Kamarudin and S. Sufian, Malays. J. Microsc., 2020, 16, 37–43 Search PubMed.
  174. T. C. M. V. Do, D. Q. Nguyen, K. T. Nguyen and P. H. Le, Materials, 2019, 12, 2434 CrossRef CAS PubMed.
  175. M. R. Rezaei, M. H. Sayadi and N. Ravankhah, Journal of Natural Environment, 2021, 74, 331–344 Search PubMed.
  176. L. Bergamonti, C. Bergonzi, C. Graiff, P. P. Lottici, R. Bettini and L. Elviri, Water Res., 2019, 163, 114841 CrossRef CAS PubMed.
  177. S. Moles, J. Berges, M. P. Ormad, M. J. Nieto-Monge, J. Gómez and R. Mosteo, Environ. Sci. Pollut. Res., 2021, 28, 24167–24179 CrossRef CAS PubMed.
  178. D. Kanakaraju, J. Kockler, C. A. Motti, B. D. Glass and M. Oelgemöller, Appl. Catal., B, 2015, 166–167, 45–55 CrossRef CAS.
  179. L. M. Pastrana-Martínez, S. Morales-Torres, S. A. C. Carabineiro, J. G. Buijnsters, J. L. Figueiredo, A. M. T. Silva and J. L. Faria, Appl. Surf. Sci., 2018, 458, 839–848 CrossRef.
  180. Q. Li, H. Kong, R. Jia, J. Shao and Y. He, RSC Adv., 2019, 9, 12538 RSC.
  181. F. M. dela Rosa, J. Papac, S. Garcia-Ballesteros, M. Kovačić, Z. Katančić, H. Kušić and A. L. Božić, Adv. Sustainable Syst., 2021, 5, 2100119 CrossRef CAS.
  182. M. G. Alalm, A. Tawfik and A. Ookawara, J. Environ. Chem. Eng., 2016, 4, 1929–1937 CrossRef.
  183. Q. Li, R. Jia, J. Shao and Y. He, J. Cleaner Prod., 2019, 209, 755–761 CrossRef CAS.
  184. F. A. Sulaiman and A. I. Alwared, J. Ecol. Eng., 2022, 23, 293–304 CrossRef.
  185. A. Aissani, M. Kameche and K. Benabbou, Inorg. Nano-Met. Chem., 2022, 52, 1197–1207 CrossRef CAS.
  186. N. Torkian, A. Bahrami, A. Hosseini-Abari, M. M. Momeni, M. Abdolkarimi Mahabadi, A. Bayat, P. Hajipour, H. A. Rourani, M. S. Abbasi, S. Torkian, Y. Wen, M. Y. Mehr and A. Hojjati-Najafabadi, Environ. Res., 2022, 207, 112157 CrossRef CAS PubMed.
  187. Y. Chen, X. Zhang, L. Wang, X. Cheng and Q. Shang, Chem. Eng. J., 2020, 402, 126260 CrossRef CAS.
  188. S. Rani, A. Garg and N. Singh, Toxicol. Environ. Chem., 2021, 103, 137–153 CrossRef CAS.
  189. T. D. N. Thi, L. H. Nguyen, X. H. Nguyen, H. V. Phung, T. H. T. Vinh, P. V. Viet, N. V. Thai, H. N. Le, D. T. Pham, H. T. Van, L. H. T. Thi, T. D. P. Thi, T. L. Minh, H. H. P. Quang, H. P. N. Vu, T. T. Duc and H. M. Nguyen, Mater. Sci. Semicond. Process., 2022, 142, 106456 CrossRef.
  190. A. Fawzy, H. Mahanna and M. Mossad, Environ. Sci. Pollut. Res., 2022, 29, 68532–68546 CrossRef CAS PubMed.
  191. H. Liu, L. Wang, S. Wei, Y. Wu, Y. Zheng, F. Yuan and J. Hou, Opt. Mater., 2022, 123, 111835 CrossRef CAS.
  192. D. Huang, X. Sun, Y. Liu, H. Ji, W. Liu, C.-C. Wang, W. Ma and Z. Cai, Chin. Chem. Lett., 2021, 32, 2787–2791 CrossRef CAS.
  193. I. F. Silva, I. F. Teixeira, R. D. F. Rios, G. M. do Nascimento, I. Binatti, H. F. V. Victória, K. Krambrock, L. A. Cury, A. P. C. Teixeira and H. O. Stumpf, J. Hazard. Mater., 2021, 401, 123713 CrossRef CAS PubMed.
  194. N. Q. Thang, A. Sabbah, L.-C. Chen, K.-H. Chen, C. M. Thi and P. V. Viet, Chemosphere, 2021, 282, 130971 CrossRef CAS PubMed.
  195. S. Le, C. Zhu, Y. Cao, P. Wang, Q. Liu, H. Zhou, C. Chen, S. Wang and X. Duan, Appl. Catal., B, 2022, 303, 120903 CrossRef CAS.
  196. M. Dou, J. Wang, B. Gao, Z. Ma and X. Huang, Chem. Eng. J., 2020, 394, 124899 CrossRef CAS.
  197. E. G. Shankar, S. Billa, A. B. V. Kiran Kumar and J. S. Yu, Ceram. Int., 2021, 47, 30572–30583 CrossRef CAS.
  198. V. V. Pham, T. K. Truong, L. V. Hai, H. P. P. La, H. T. Nguyen, V. Q. Lam, H. D. Tong, Q. Nguyen, A. Sabbah, K.-H. Chen, S.-J. You and T. M. Cao, ACS Appl. Nano Mater., 2022, 5, 4506–4514 CrossRef CAS.
  199. M. Dou, J. Wang, B. Gao, C. Xu and F. Yang, Chem. Eng. J., 2020, 383, 123134 CrossRef CAS.
  200. Q. Chen, L. Chen, J. Qi, Y. Tong, Y. Lv, C. Xu, J. Ni and W. Liu, Chin. Chem. Lett., 2019, 30, 1214–1218 CrossRef CAS.
  201. R. Changotra, Q. He and A. Dhir, Chem. Eng. J., 2022, 442, 136201 CrossRef CAS.
  202. D. Balarak, N. Mengelizadeh, P. Rajiv and K. Chandrika, Environ. Sci. Pollut. Res., 2021, 28, 49743–49754 CrossRef CAS PubMed.
  203. J. Song, Z. Xu, W. Liu and C.-T. Chang, Mater. Sci. Semicond. Process., 2016, 52, 32–37 CrossRef CAS.
  204. C. Yang, X. You, J. Cheng, H. Zheng and Y. Chen, Appl. Catal., B, 2017, 200, 673–680 CrossRef CAS.
  205. M. F. R. Samsudin and S. Sufian, J. Mol. Liq., 2020, 314, 113530 CrossRef CAS.
  206. D. V. Thuan, T. L. Nguyen, H. H. P. Thi, N. T. Thanh, S. Ghotekar, A. K. Sharma, M. T. Binh, T. T. Nga, T.-D. Pham and D. P. Cam, Opt. Mater., 2022, 123, 111885 CrossRef.
  207. F. Beshkar, A. Al-Nayili, O. Amiri, M. Salavati-Niasari and M. Mousavi-Kamazani, J. Alloys Compd., 2022, 892, 162176 CrossRef CAS.
  208. M. R. S. Nivetha, J. V. Kumar, J. S. Ajarem, A. A. Allam, V. Manikandan, R. Arulmozhi and N. Abirami, Environ. Res., 2022, 209, 112809 CrossRef CAS PubMed.
  209. E. M. El-Fawal, S. A. Younis and T. Zaki, J. Photochem. Photobiol., A, 2020, 401, 112746 CrossRef CAS.
  210. Y. Belaissa, D. Nibou, A. A. Assadi, B. Bellal and M. Trari, J. Taiwan Inst. Chem. Eng., 2016, 68, 254–265 CrossRef CAS.
  211. B. Gao, J. Wang, M. Dou, C. Xu and X. Huang, Environ. Sci. Pollut. Res., 2020, 27, 7025–7039 CrossRef CAS PubMed.
  212. M. H. T. Tung, T. T. T. Phuong, D. M. N. Tram, D. M. The, N. V. N. Mai, T. T. T. Hien, L. T. C. Nhung, N. T. T. Binh, C. V. Hoang, D. N. Nhiem, T.-D. Pham and N. T. D. Cam, Diamond Relat. Mater., 2022, 121, 108788 CrossRef CAS.
  213. Q. Li, H. Kong, P. Li, J. Shao and Y. He, J. Hazard. Mater., 2019, 373, 437–446 CrossRef CAS PubMed.
  214. K. H. Leong, H. Y. Chu, S. Ibrahi and P. Saravanan, Beilstein J. Nanotechnol., 2015, 6, 428–437 CrossRef PubMed.
  215. P. Hajipour, A. Eslami, A. Bahrami, A. Hosseini-Abari, F. Y. Saber, R. Mohammadi and M. Y. Mehr, Ceram. Int., 2021, 47, 33875–33885 CrossRef CAS.
  216. M. Moradi, F. Hasanvandian, A. A. Isari, F. Hayati, B. Kakavandi and S. R. Setayesh, Appl. Catal., B, 2021, 285, 119838 CrossRef CAS.
  217. G. Prasannamedha and P. S. Kumar, J. Cleaner Prod., 2020, 250, 119553 CrossRef CAS.
  218. J. Musial, D. T. Mlynarczyk and B. J. Stanisz, Sci. Total Environ., 2023, 856, 159122 CrossRef CAS PubMed.
  219. C. H. Wu, C. Y. Kuo, C. D. Dong, C. W. Chen and Y. L. Lin, Water Sci. Technol., 2019, 79, 349–355 CrossRef CAS PubMed.
  220. L. Prieto-Rodriguez, S. Miralles-Cuevas, I. Oller, P. Fernández-Ibánez, A. Agüera, J. Blanco and S. Malato, Appl. Catal., B, 2012, 128, 119–125 CrossRef CAS.
  221. A. Hu, X. Zhang, D. Luong, K. D. Oakes, M. R. Servos, R. Liang, S. Kurdi, P. Peng and Y. Zhou, Waste Biomass Valorization, 2012, 3, 443–449 CrossRef CAS.
  222. M. R. Eskandarian, H. Choi, M. Fazli and M. H. Rasoulifard, Chem. Eng. J., 2016, 300, 414–422 CrossRef CAS.
  223. M. N. Abellán, B. Bayarri, J. Giménez and J. Costa, Appl. Catal., B, 2007, 74, 233–241 CrossRef.
  224. D. Balarak and H. Azarpira, Int. J. ChemTech Res., 2016, 9, 731–738 CAS.
  225. J. Carbajo, M. Jiménez, S. Miralles, S. Malato, M. Faraldos and A. Bahamonde, Chem. Eng. J., 2016, 291, 64–73 CrossRef CAS.
  226. A. Kutuzova, T. Dontsova and W. Kwapinski, Catalysts, 2021, 11, 728 CrossRef CAS.
  227. O. Porcar-Santos, A. Cruz-Alcalde, N. López-Vinent, D. Zanganas and C. Sans, Sci. Total Environ., 2020, 736, 139605 CrossRef CAS PubMed.
  228. J. R. Kim and E. Kan, J. Environ. Manage., 2016, 180, 94–101 CrossRef CAS PubMed.
  229. N. Miranda-Garcia, S. Suarez, B. Sanchez, J. M. Coronado, S. Alato and M. I. Maldonsado, Appl. Catal., B, 2011, 103, 294–301 CrossRef CAS.
  230. S. Ramasundaram, M. G. Seid, J. W. Choe, E.-J. Kim, Y. C. Chung, K. Cho, C. Lee and S. W. Hong, Chem. Eng. J., 2016, 306, 344–351 CrossRef CAS.
  231. M. J. Arlos, M. M. Hatat-Fraile, R. Liang, L. M. Bragg, N. Y. Zhou, S. A. Andrews and M. R. Servos, Water Res., 2016, 101, 351–361 CrossRef CAS PubMed.
  232. A. Mirzaei, L. Yerushalmi, Z. Chen, F. Haghighat and J. Guo, Water Res., 2018, 132, 241–251 CrossRef CAS PubMed.
  233. T. Makropoulou, I. Kortidis, K. Davididou, D. E. Motaung and E. Chatzisymeon, J. Water Process. Eng., 2020, 36, 101299 CrossRef.
  234. F. Beheshti, R. M. A. Tehrani and A. Khadir, Int. J. Environ. Sci. Technol., 2019, 16, 7987–7996 CrossRef CAS.
  235. A. Tiwari, A. Shukla, Lalliansanga, D. Tiwari and S. M. Lee, J. Environ. Manage., 2018, 220, 96–108 CrossRef CAS PubMed.
  236. E. Borowska, J. F. Gomes, R. C. Martins, R. M. Quinta-Ferreira, H. Horn and M. Gmurek, Catalysts, 2019, 9, 500 CrossRef CAS.
  237. L.-F. Chiang and R. A. Doong, Sep. Purif. Technol., 2015, 156, 1003–1010 CrossRef CAS.
  238. R. Zanella, E. Avella, R. M. Ramírez-Zamora, F. Castillón-Barraza and J. C. D. Álvarez, Environ. Technol., 2018, 39, 2353–2364 CrossRef CAS PubMed.
  239. A. Tsiampalis, Z. Frontistis, V. Binas, G. Kiriakidis and D. Mantzavinos, Catalysts, 2019, 9, 612 CrossRef CAS.
  240. M. Jahdi, S. B. Mishra, E. N. Nxumalo, S. D. Mhlanga and A. K. Mishra, Appl. Catal., B, 2020, 267, 118716 CrossRef CAS.
  241. M. Jahdi, S. B. Mishra, E. N. Nxumalo, S. D. Mhlanga and A. K. Mishra, RSC Adv., 2020, 10, 27662 RSC.
  242. L. K. B. Paragas, M. D. G. de Luna and R.-A. Doong, Chemosphere, 2018, 210, 1099–1107 CrossRef CAS PubMed.
  243. A. A. Isari, F. Hayati, B. Kakavandi, M. Rostami, M. Motevassel and E. Dehghanifard, Chem. Eng. J., 2020, 392, 123685 CrossRef CAS.
  244. T.-B. Nguyena, C. P. Huang, R. Doong, C.-W. Chen and C.-D. Dong, Chem. Eng. J., 2020, 384, 123383 CrossRef.
  245. A. Mirzaei, M. Eddah, S. Roualdès, D. Ma and M. Chaker, Chem. Eng. J., 2021, 422, 130507 CrossRef CAS.
  246. I. Zammit, V. Vaiano, A. R. Ribeiro, A. M. T. Silva, C. M. Manaia and L. Rizzo, Catalysts, 2019, 9, 222 CrossRef.
  247. X. Xie, S. Li, H. Zhang, Z. Wang and H. Huang, Sci. Total Environ., 2019, 659, 529–539 CrossRef CAS PubMed.
  248. L. Fernández, M. Gamallo, M. A. González-Gómez, C. Vázquez-Vázquez, J. Rivas, M. Pintado and M. T. Moreira, J. Environ. Manage., 2019, 237, 595–608 CrossRef PubMed.
  249. V. Polliotto, F. R. Pomilla, V. Maurino, G. Marcì, A. B. Prevot, R. Nisticò, G. Magnacca, M. C. Paganini, L. Ponce Robles, L. Perez and S. Malato, Catal. Today, 2019, 328, 164–171 CrossRef CAS.
  250. N. T. T. Nguyen, A. Q. K. Nguyen, M. S. Kim, C. Lee, S. Kim and J. Kim, Sep. Purif. Technol., 2021, 267, 118610 CrossRef CAS.
  251. N. Wang, X. Li, Y. Yang, Z. Zhou, Y. Shang and X. Zhuang, J. Water Process. Eng., 2020, 36, 101335 CrossRef.
  252. N. Rioja, P. Benguria, F. J. Peñas and S. Zorita, Environ. Sci. Pollut. Res., 2014, 21, 11168–11177 CrossRef CAS PubMed.
  253. M. O. Alfred, M. O. Omorogie, O. Bodede, R. Moodley, A. Ogunlaja, O. G. Adeyemi, C. Gunter, A. Trubert, H. Eckert, L. D. A. Silva, A. S. S. de Camargo, A. de J. Motheo, S. M. Clarke and E. I. Unuabonah, Chem. Eng. J., 2020, 398, 125544 CrossRef.
  254. E. H. Mourid, E. M. El Mouchtari, L. El Mersly, L. Benaziz, S. Rafqah and M. Lakraimi, J. Photochem. Photobiol., A, 2020, 396, 112530 CrossRef CAS.
  255. M. Długosz, P. Zmudzki, A. Kwiecień, K. Szczubiałka and J. Krzek, J. Hazard. Mater., 2015, 298, 146–153 CrossRef PubMed.
  256. R. Noroozi, M. Gholami, M. Farzadkia and R. R. Kalantary, Environ. Sci. Pollut. Res., 2022, 29, 56403–56418 CrossRef CAS PubMed.
  257. N. Malesic-Eleftheriadou, E. Evgenidou, G. Z. Kyzas, D. N. Bikiaris and D. A. Lambropoulou, Chemosphere, 2019, 234, 746–755 CrossRef CAS PubMed.
  258. J. Li, F. Wang, L. Meng, M. Han, Y. Guo and C. Sun, J. Colloid Interface Sci., 2017, 485, 116–122 CrossRef CAS PubMed.
  259. K. Kouvelis, A. A. Kampioti, A. Petala and Z. Frontistis, Catalysts, 2022, 12, 882 CrossRef CAS.
  260. O. Mertah, A. Gomez-Aviles, A. Kherbeche, C. Belver and J. Bedia, J. Environ. Chem. Eng., 2022, 10, 108438 CrossRef CAS.
  261. W. Zhu, Z. Li, C. He, S. Faqian and Y. Zhou, J. Alloys Compd., 2018, 754, 153–162 CrossRef CAS.
  262. D. Awfa, M. Ateia, M. Fujii and C. Yoshimura, J. Water Process. Eng., 2019, 31, 100836 CrossRef.
  263. J. Martini, C. A. Orge, J. L. Faria, M. F. R. Pereira and O. S. G. P. Soares, Appl. Sci., 2019, 9, 2652 CrossRef CAS.
  264. M. Chen, C. Guo, S. Hou, L. Wu, J. Lv, C. Hu, Y. Zhang and J. Xu, J. Hazard. Mater., 2019, 366, 219–228 CrossRef CAS PubMed.
  265. A. Mirzaei, L. Yerushalmi, Z. Chen and F. Haghighat, J. Hazard. Mater., 2018, 359, 516–526 CrossRef CAS PubMed.
  266. G. K. Teye, J. Huang, Y. Li, K. Li, L. Chen and W. K. Darkwah, Nanomaterials, 2021, 11, 2609 CrossRef CAS PubMed.
  267. X. Ao, Z. Li and H. Zhang, J. Cleaner Prod., 2022, 356, 131822 CrossRef CAS.
  268. Y. Song, J. Qi, J. Tian, S. Gao and F. Cui, Chem. Eng. J., 2018, 341, 547–555 CrossRef CAS.
  269. W. Zhu, F. Sun, R. Goei and Y. Zhou, Appl. Catal., B, 2017, 207, 93–102 CrossRef CAS.
  270. L. Zhou, G. Zou and G. L. Deng, Catalysts, 2018, 8, 272 CrossRef.
  271. L. Lin, H. Wang and P. Xu, Chem. Eng. J., 2017, 310, 389–398 CrossRef CAS.
  272. S.-H. Liu, Y.-S. Wei and J. S. Lu, Chemosphere, 2016, 154, 118–123 CrossRef CAS PubMed.
  273. M. Nawaz, A. A. Khan, A. Hussain, J. Jang, H.-Y. Jung and D. S. Lee, Chemosphere, 2020, 261, 127702 CrossRef CAS PubMed.
  274. L. Zhou, O. G. Alvarez, C. S. Mazon, L. Chen, H. Deng and M. Sui, Catal. Sci. Technol., 2016, 6, 5972 RSC.
  275. W. Zhu, F. Sun, R. Goei and Y. Zhou, Catal. Sci. Technol., 2017, 7, 2591–2600 RSC.
  276. M. H. de M. Rodrigues, P. A. R. de Sousa, K. C. M. Borges, L. de M. Coelho, R. F. Gonçalves, M. D. Teodoro, F. da V. Motta, R. M. do Nascimento and M. G. Júnior, J. Alloys Compd., 2019, 808, 151711 CrossRef.
  277. A. Kumar, G. Sharma, M. Naushad, Z. A. Alothman and P. Dhiman, Earth Syst. Environ., 2022, 6, 141–156 CrossRef.
  278. J. Qiu, C. Yue, W. Zheng, F. Liu and J. Zhu, Chin. Chem. Lett., 2021, 32, 3431–3434 CrossRef CAS.
  279. A. Kumar, A. Kumar, G. Sharma, A. H. Al-Muhtaseb, M. Naushad, A. A. Ghfar and F. J. Stadler, Chem. Eng. J., 2018, 334, 462–478 CrossRef CAS.
  280. Z. Huang, X. Dai, Z. Huang, T. Wang, L. Cui, J. Ye and P. Wu, Chemosphere, 2019, 221, 824–833 CrossRef CAS PubMed.
  281. L. Zhang, G. Meng, B. Liu and X. Ge, J. Mol. Liq., 2022, 360, 119427 CrossRef CAS.
  282. D. Roy, S. Niyogi and S. De, Process Saf. Environ. Prot., 2022, 161, 723–738 CrossRef CAS.
  283. T. Ke, S. Shen, K. Yang and D. Lin, Appl. Surf. Sci., 2022, 580, 152302 CrossRef CAS.
  284. M. Ren, Y. Ao, P. Wang and C. Wang, Chem. Eng. J., 2019, 378, 122122 CrossRef CAS.
  285. C. Liu, J. Xu, J. Niu, M. Chen and Y. Zhou, Sep. Purif. Technol., 2020, 241, 116622 CrossRef CAS.
  286. C. Liu, J. Xu, X. Du, Q. Li, Y. Fu and M. Chen, Opt. Mater., 2021, 112, 110742 CrossRef CAS.
  287. J. Xu, J. Chen, Y. Ao and P. Wang, Chin. Chem. Lett., 2021, 32, 3226–3230 CrossRef CAS.
  288. L. Zhou, W. Zhang, L. Chen and H. Deng, J. Colloid Interface Sci., 2017, 487, 410–417 CrossRef CAS PubMed.
  289. A. Mirzaei, Z. Chen, F. Haghighat and L. Yerushalm, Chemosphere, 2018, 205, 463–474 CrossRef CAS PubMed.
  290. G. Liu, H. Wang, D. Chen, C. Dai, Z. Zhang and Y. Feng, Sep. Purif. Technol., 2020, 237, 116329 CrossRef CAS.
  291. Y. Zhang, Y. Li and Y. Yuan, J. Colloid Interface Sci., 2023, 645, 860–869 CrossRef CAS PubMed.
  292. E. Brillas, Chemosphere, 2022, 286, 131849 CrossRef CAS PubMed.
  293. J. Jan-Roblero and J. A. Cruz-Maya, Molecules, 2023, 28, 2097 CrossRef CAS PubMed.
  294. N. Jallouli, L. M. Pastrana-Martinez, A. R. Ribeiro, N. F. F. Moreira, J. L. Faria, O. Hentati, A. M. T. Silva and M. Ksibi, Chem. Eng. J., 2018, 334, 976–984 CrossRef CAS.
  295. M. Jiménez-Salcedo, M. Monge and M. T. Tena, Chemosphere, 2019, 215, 605–618 CrossRef PubMed.
  296. M. O. Miranda, W. E. C. Cavalcanti, F. F. Barbosa, J. A. de Sousa, F. I. da Silva, S. B. C. Pergher and T. P. Braga, RSC Adv., 2021, 11, 27720 RSC.
  297. I. Georgaki, E. Vasilaki and N. Katsarakis, Am. J. Anal. Chem., 2014, 5, 518–534 CrossRef.
  298. M. Tanveer, G. T. Guyer and G. Abbas, Water Environ. Res., 2019, 91, 822–829 CrossRef CAS PubMed.
  299. N. Rastkari, A. Eslami, S. Nasseri, E. Piroti and A. Asadi, Pol. J. Environ. Stud., 2017, 26, 785–794 CrossRef CAS PubMed.
  300. A. S. Sá, R. P. Feitosa, L. Honório, R. Peña-Garcia, L. C. Almeida, J. S. Dias, L. P. Brazuna, T. G. Tabuti, E. R. Triboni, J. A. Osajima and E. C. da Silva-Filho, Materials, 2021, 14, 5891 CrossRef PubMed.
  301. M. Ulfa, D. Prasetyoko, H. Bahruji and R. E. Nugraha, Materials, 2021, 14, 6779 CrossRef CAS PubMed.
  302. S. Khalaf, J. H. Shoqeir, F. Lelario, S. A. Bufo, R. Karamanand and L. Scrano, Catalysts, 2020, 10, 560 CrossRef CAS.
  303. F. S. Braz, M. R. A. Silva, F. S. Silva, S. J. Andrade, A. L. Fonseca and M. M. Kondo, J. Environ. Prot., 2014, 5, 620–626 CrossRef.
  304. J. Bohdziewicz, E. Kudlek and M. Dudziak, Desalin. Water Treat., 2016, 57, 1552–1563 CrossRef CAS.
  305. J. Choina, A. Bagabas, C. Fisher, G. U. Flechsig, H. Kosslick, A. Alshammari and A. Schulz, Catal. Today, 2015, 241, 47–54 CrossRef CAS.
  306. N. Rosman, W. N. W. Salleh, J. Jaafar, Z. Harun, F. Aziz and A. F. Ismail, Catalysts, 2022, 12, 209 CrossRef CAS.
  307. A. Eslami, M. M. Amini, A. R. Yazdanbakhsh, A. Mohseni-Bandpei, A. A. Safari and A. Asadi, J. Chem. Technol. Biotechnol., 2016, 91, 2693–2704 CrossRef CAS.
  308. T. M. Khedr, S. M. El-Sheikh, A. Hakki, A. A. Ismail, W. A. Badawy and D. W. Bahnemann, J. Photochem. Photobiol., A, 2017, 346, 530–540 CrossRef CAS.
  309. V. Bhatia and A. Dhir, J. Environ. Chem. Eng., 2016, 4, 1267–1273 CrossRef CAS.
  310. A. Jraba, Z. Anna and E. Elaloui, J. Mater. Sci.: Mater. Electron., 2020, 31, 1072–1083 CrossRef CAS.
  311. R. Pelosato, V. Carrara and I. N. Sora, Chem. Eng. Trans., 2019, 73, 181–186 Search PubMed.
  312. Q. Sun, Y.-P. Peng, H. Chen, K.-L. Chang, Y.-N. Qiu and S.-W. Lai, J. Hazard. Mater., 2016, 319, 121–129 CrossRef CAS PubMed.
  313. S. M. El-Sheikh, T. M. Khedr, A. Hakki, A. A. Ismail, W. A. Badawy and D. W. Bahnemann, Sep. Purif. Technol., 2017, 173, 258–268 CrossRef CAS.
  314. C. Yuan, C.-H. Hung, H.-W. Li and W.-H. Chang, Chemosphere, 2016, 155, 471–478 CrossRef CAS PubMed.
  315. L. Lin, W. Jiang, M. Bechelany, M. Nasr, J. Jarvis, T. Schaub, R. R. Sapkota, P. Miele, H. Wang and P. Xu, Chemosphere, 2019, 220, 921–929 CrossRef CAS PubMed.
  316. Y. Gu, J. Yperman, R. Carleer, J. D'Haen, J. Maggen, S. Vanderheyden, K. Vanreppelen and R. M. Garcia, Chemosphere, 2019, 217, 724–731 CrossRef CAS PubMed.
  317. M. Mohadesi, A. Gouran and K. Seifi, Environ. Sci. Pollut. Res., 2022, 29, 34338–34348 CrossRef CAS PubMed.
  318. N. Liu, J. Wang, J. Wu, Z. Li, W. Huang, Y. Zheng, J. Lei, X. Zhang and L. Tang, Mater. Res. Bull., 2020, 132, 111000 CrossRef CAS.
  319. T. R. Bastami, A. Ahmadpour and F. A. Hekmatikar, J. Ind. Eng. Chem., 2017, 51, 244–254 CrossRef.
  320. C. S. L. Fung, M. Khan, A. Kumar and I. M. C. Lo, Sep. Purif. Technol., 2019, 216, 102–114 CrossRef CAS.
  321. M. P. Villavicencio, A. E. Morales, M. de L. R. Peralta, M. Sánchez-Cantú, L. R. Blanco, E. C. Anota, J. H. C. García and F. Tzompantzi, Catal. Lett., 2020, 150, 2385–2399 CrossRef.
  322. J. Liao, Y. Xu, Y. Zhao, C.-C. Wang and C. Ge, ACS Appl. Nano Mater., 2021, 4, 1898–1905 CrossRef CAS.
  323. G. G. Lenzi, M. F. Lopes, D. I. Andrade, J. S. Napoli, A. Parolin, Y. B. Fávaro, M. E. K. Fuziki, L. N. B. de Almeida, T. G. Josué, D. T. Dias and A. M. Tusset, Water Sci. Technol., 2021, 84, 2158–2179 CrossRef CAS PubMed.
  324. M. Khan, C. S. L. Fung, A. Kumar and I. M. C. Lo, J. Hazard. Mater., 2019, 365, 733–743 CrossRef CAS PubMed.
  325. K. Kang, M. Jang, M. Cui, P. Qiu, B. Park, S. A. Snyder and J. Khim, J. Mol. Catal. A: Chem., 2014, 390, 178–186 CrossRef CAS.
  326. C. B. Anucha, I. Altin, E. Bacaksiz, I. Degirmencioglu, T. Kucukomeroglu, S. Yilmaz and V. N. Stathopoulos, Separations, 2021, 8, 24 CrossRef.
  327. S. Sambaza, A. Maity and K. Pillay, J. Saudi Chem. Soc., 2022, 26, 101563 CrossRef CAS.
  328. A. Mohamed, A. Salama, W. S. Nasser and A. Uheida, Environ. Sci. Eur., 2018, 30, 47 CrossRef PubMed.
  329. E. Yilmaz, S. Salem, G. Sarp, S. Aydin, K. Sahin, I. R. Korkmaz and D. Yuvali, Talanta, 2020, 213, 120813 CrossRef CAS PubMed.
  330. B.-T. Zhang, Q. Wang, Y. Zhang, Y. Teng and M. Fan, Sep. Purif. Technol., 2020, 242, 116820 CrossRef CAS.
  331. N. Rosman, W. N. W. Salleh, N. A. M. Razali, S. Z. Ahmad, N. H. Ismail, F. Aziz, Z. Harun, A. F. Ismail and N. Yuso, Mater. Today: Proc., 2021, 42, 69–74 CAS.
  332. A. Uheida, A. Mohamed, M. Belaqziz and W. S. Nasser, Sep. Purif. Technol., 2019, 212, 110–118 CrossRef CAS.
  333. D. Hernández-Uresti, A. Vázquez, D. Sanchez-Martinez and S. Obregón, J. Photochem. Photobiol., A, 2016, 324, 47–52 CrossRef.
  334. X. Wang, J. Luo, Y. Huang, J. Mei and Y. Chen, Environ. Sci.: Water Res. Technol., 2021, 7, 610 RSC.
  335. A. Raja, P. Rajasekaran, K. Selvakumar, M. Arivanandhan, S. A. Bahadur and M. S. Swaminathan, Appl. Surf. Sci., 2020, 513, 145803 CrossRef CAS.
  336. A. Kumar, G. Sharma, M. Naushad, A. H. Al-Muhtaseb, A. Kumar, I. Hira, T. Ahamad, A. A. Ghfar and F. J. Stadler, J. Environ. Manage., 2019, 231, 1164–1175 CrossRef CAS PubMed.
  337. R. Akbarzadeh, C. S. L. Fung, R. A. Rather and I. M. C. Lo, Chem. Eng. J., 2018, 341, 248–261 CrossRef CAS.
  338. Z. Sun, X. Zhang, X. Dong, X. Liu, Y. Tan, F. Yuan, S. Zheng and C. Li, J. Materiomics, 2020, 6, 582–592 CrossRef.
  339. Z.-D. Lei, J.-J. Wang, L. Wang, X.-Y. Yang, G. Xu and L. Tang, J. Hazard. Mater., 2016, 312, 298–306 CrossRef CAS PubMed.
  340. W. Cao, Y. Yuan, C. Yang, S. Wu and J. Cheng, Chem. Eng. J., 2020, 391, 123608 CrossRef CAS.
  341. Y. Zhang, J. Li and H. Liu, J. Nanomater., 2020, 6094984,  DOI:10.1155/2020/6094984.
  342. L. Wang, Q. Sun, Y. Dou, Z. Zhang, T. Yan and Y. Li, J. Hazard. Mater., 2021, 413, 125288 CrossRef CAS PubMed.
  343. X. Chen, X. Li, J. Yang, Q. Sun, Y. Yang and X. Wu, Int. J. Hydrogen Energy, 2018, 43, 13284–13293 CrossRef CAS.
  344. J. Wang, L. Tang, G. Zeng, Y. Deng, Y. Liu, L. Wang, Y. Zhou, Z. Guo, J. Wang and C. Zhang, Appl. Catal., B, 2017, 209, 285–294 CrossRef CAS.
  345. A. Kumar, M. Khan, X. Zeng and I. M. C. Lo, Chem. Eng. J., 2018, 353, 645–656 CrossRef CAS.
  346. N. Liu, F. Fei, W. Dai, J. Lei, F. Bi, B. Wang, G. Quan, X. Zhang and L. Tang, J. Colloid Interface Sci., 2022, 625, 965–977 CrossRef CAS PubMed.
  347. S. Mao, C. Liu, M. Xia, F. Wang and X. Ju, Catal. Sci. Technol., 2021, 11, 3466–3480 RSC.
  348. S.-H. Liu and W.-T. Tang, Sci. Total Environ., 2020, 731, 139172 CrossRef CAS PubMed.
  349. Y. Cong, Y. Li, X. Wang, X. Wei, L. Che and S.-W. Lv, Sep. Purif. Technol., 2022, 297, 121531 CrossRef CAS.
  350. P. J. Mafa, U. S. Swana, D. Liu, J. Gui, B. B. Mamba and A. T. Kuvarega, Colloids Surf., A, 2021, 612, 126004 CrossRef CAS.
  351. J. Zhang, L. Xin, L. Qianwen, L. Yuqian, G. Liu and X. Shi, Ceram. Int., 2020, 46, 106–113 CrossRef CAS.
  352. A. Kumar, M. Khan, L. Fang and I. M. C. Lo, J. Hazard. Mater., 2019, 370, 108–116 CrossRef CAS PubMed.
  353. M. Liang, Z. Zhang, R. Long, Y. Wang, Y. Yu and Y. Pei, Environ. Pollut., 2020, 259, 113770 CrossRef CAS PubMed.
  354. M. E. Malefane, U. Feleni, P. J. Mafa and A. T. Kuvarega, Appl. Surf. Sci., 2020, 514, 145940 CrossRef CAS.
  355. Q. Wei, S. Xiong, W. Li, C. Jin, Y. Chen, L. Hou, Z. Wu, Z. Pan, Q. He, Y. Wang and D. Tang, J. Alloys Compd., 2021, 885, 160984 CrossRef CAS.
  356. A. Zhou, L. Liao, X. Wu, K. Yang, C. Li, W. Chen and P. Xie, Sep. Purif. Technol., 2020, 250, 117241 CrossRef CAS.
  357. N. Rosman, W. N. W. Salleh, M. A. Mohamed, Z. Harun, A. F. Ismail and F. Aziz, Sep. Purif. Technol., 2020, 251, 117391 CrossRef CAS.
  358. L. Chierentin and H. R. N. Salgado, Crit. Rev. Anal. Chem., 2016, 46, 22–39 CrossRef CAS PubMed.
  359. B. Holmes, R. N. Brogden and D. M. Richards, Drugs, 1985, 30, 482–513 CrossRef CAS PubMed.
  360. A. J. Schaeffer, Eur. Urol., 1990, 17, 19–23 CrossRef PubMed.
  361. H. Yang, L. Mei, P. Wang, J. Genereux, Y. Wang, B. Yi, C. Au, L. Dang and P. Feng, RSC Adv., 2017, 7, 45721 RSC.
  362. M. M. Haque and M. Muneer, J. Hazard. Mater., 2007, 145, 51–57 CrossRef CAS PubMed.
  363. Z. Liu, X. Yu, F. Yang, J. Nie, N. Zhao, J. Li and B. Yao, IOP Conf. Ser. Earth Environ. Sci., 2020, 453, 012084 CrossRef.
  364. M. Chen, Y. Huang and W. Chu, Chin. J. Catal., 2019, 40, 673–680 CrossRef CAS.
  365. M. Sayed, L. A. Shah, J. A. Khan, N. S. Shah, J. Nisar, H. M. Khan, P. Zhang and A. R. Khan, J. Phys. Chem. A, 2016, 120, 9916–9931 CrossRef CAS PubMed.
  366. S.-L. Zhou, S. Zhang, F. Liu, J.-J. Liu, J.-J. Xue, D.-J. Yang and C.-T. Chang, J. Photochem. Photobiol., A, 2016, 328, 97–104 CrossRef CAS.
  367. S. Zhang, S.-L. Zhou, J.-J. Liu, D.-J. Yang, J.-J. Xue and C.-T. Chang, J. Nanosci. Nanotechnol., 2018, 18, 4087–4092 CrossRef CAS PubMed.
  368. R. Xiao, Y. Zhang, S. Wang, H. Zhu, H. Song, G. Chen, H. Lin, J. Zhang and J. Xiang, Environ. Sci. Pollut. Res., 2021, 28, 69301–69313 CrossRef CAS PubMed.
  369. S. S. Mohtar, F. Aziz, A. F. Ismail, N. S. Sambudi, H. Abdullah, A. N. Rosli and B. Ohtani, Catalysts, 2021, 11, 1160 CrossRef CAS.
  370. X. Jin, X. Zhou, P. Sun, S. Lin, W. Cao, Z. Li and W. Liu, Chemosphere, 2019, 237, 124433 CrossRef CAS PubMed.
  371. R. Kaushik, P. K. Samal and A. Halder, ACS Appl. Nano Mater., 2019, 2, 7898–7909 CrossRef CAS.
  372. M. Chen and W. Chu, J. Hazard. Mater., 2012, 219, 183–189 CrossRef PubMed.
  373. M. Manasa, P. R. Chandewar and H. Mahalingam, Catal. Today, 2021, 375, 522–536 CrossRef CAS.
  374. N. S. Shah, J. A. Khan, M. Sayed, Z. U. H. Khan, A. D. Rizwan, N. Muhammad, G. Boczkaj, B. Murtaza, M. Imran, H. M. Khan and G. Zaman, Chem. Eng. J., 2018, 351, 841–855 CrossRef CAS.
  375. Y. Hsu, J. Thomas, C. T. Chang and C. M. Ma, J. Nanosci. Nanotechnol., 2021, 21, 3099–3106 CrossRef CAS PubMed.
  376. B. P. Patil and R. V. Jayaram, Catal. Green Chem. Eng., 2021, 4, 51–63 CrossRef CAS.
  377. P. Garcia-Muñoz, N. P. Zussblatt, G. Pliego, J. A. Zazo, F. Fresno, B. F. Chmelka and J. A. Casas, J. Environ. Manage., 2019, 238, 243–250 CrossRef PubMed.
  378. N. T. T. Trang, T. Q. Vinh, H. V. Giang, N. S. Mai, N. T. Dong, P. T. Linh, N. V. Hoang and N. M. Tan, Vietnam J. Sci. Technol., 2020, 58, 13–19 CrossRef.
  379. X. Wang, Y. Sun, L. Yang, Q. Shang, D. Wang, T. Guo and Y. Guo, Sci. Total Environ., 2019, 656, 1010–1020 CrossRef CAS PubMed.
  380. J. Gou, Q. Ma, X. Deng, Y. Cui, H. Zhang, X. Cheng, X. Li, M. Xie and Q. Cheng, Chem. Eng. J., 2017, 308, 818–826 CrossRef CAS.
  381. J. Li, M. Han, Y. Guo, F. Wang and C. Sun, Chem. Eng. J., 2016, 298, 300–308 CrossRef CAS.
  382. Q. Zhao, C.-C. Wang and P. Wang, Chin. Chem. Lett., 2022, 33, 4828–4833 CrossRef CAS.
  383. X.-J. Wen, C.-G. Niu, D.-W. Huang, L. Zhang, C. Liang and G.-M. Zeng, J. Catal., 2017, 355, 73–86 CrossRef CAS.
  384. W. Liu, T. He, Y. Wang, G. Ning, Z. Xu and X. Chen, Sci. Rep., 2020, 10, 11903 CrossRef CAS PubMed.
  385. L. Bharali, J. Kalita, S. S. Dhar and N. S. Moyon, ChemistrySelect, 2022, 7, e202203487 CrossRef CAS.
  386. C.-H. Zhang, X.-B. Zhao, Y.-J. Li and D.-W. Sun, Yingyong Huaxue, 2021, 38, 99–106 CAS.
  387. Z. Jiao, J. Zhang, Z. Liu and Z. Ma, J. Photochem. Photobiol., A, 2019, 371, 67–75 CrossRef CAS.
  388. W. Y. Fei, W. F. Liang, L. J. Hua, X. Z. Jie, S. Y. Han, Z. Q. Xin, Y. Kun, L. W. Ying, L. Jin and L. G. Guang, Zhongguo Huanjing Kexue, 2018, 38, 1346–1355 Search PubMed.
  389. W. Zhang, Z. Bian, X. Xin, L. Wang, X. Geng and H. Wang, Chemosphere, 2021, 262, 127955 CrossRef CAS PubMed.
  390. Q. Chen, Y. Hao, Z. Song, M. Liu, D. Chen, B. Zhu, J. Chen and Z. Chen, Ecotoxicol. Environ. Saf., 2021, 225, 112742 CrossRef CAS PubMed.
  391. W. Liu, J. Zhou and J. Yao, Ecotoxicol. Environ. Saf., 2020, 190, 110062 CrossRef CAS PubMed.
  392. S. L. Prabavathi and V. Muthuraj, Colloids Surf., A, 2019, 567, 43–54 CrossRef.
  393. L. Ma, J. Duan, B. Ji, Y. Liu, C. Li, C. Li, W. Zhao and Z. Yang, J. Alloys Compd., 2021, 869, 158679 CrossRef CAS.
  394. Y. Zhao, X. Liang, X. Hu and J. Fan, J. Colloid Interface Sci., 2021, 589, 336–346 CrossRef CAS PubMed.
  395. W. Zhao, J. Duan, B. Ji, L. Ma and Z. Yang, J. Environ. Chem. Eng., 2018, 8, 102206 CrossRef.
  396. J. Wu, J. Bai, Z. Wang, Z. Liu, Y. Mao, B. Liu and X. Zhu, Environ. Technol., 2022, 43, 95–106 CrossRef CAS PubMed.
  397. H. Mohan, S. Vadivel, P. M. Sathya, M. Fujii, G. H. Ha, G. Kim and T. Shin, ACS Appl. Energy Mater., 2022, 5, 12851–12859 CrossRef CAS.
  398. B. Ji, W. Zhao, J. Duan, L. Fu, L. Ma and Z. Yang, RSC Adv., 2020, 10, 4427–4435 RSC.
  399. Q. Wu, Y. Liu, H. Jing, H. Yu, Y. Lu, M. Huo and H. Huo, Chem. Eng. J., 2020, 390, 124615 CrossRef CAS.
  400. S. Wen, M. Chen and H. Cao, Int. J. Electrochem. Sci., 2021, 16, 210733,  DOI:10.20964/2021.07.63.
  401. G. Li, S. Huang, N. Zhu, H. Yuan, D. Ge and Y. Wei, Chem. Eng. J., 2021, 421, 127852 CrossRef CAS.
  402. J. Guo, C.-H. Shen, J. Sun, X.-J. Xu, X.-Y. Li, Z.-H. Fei, Z.-T. Liu and X.-J. Wen, Sep. Purif. Technol., 2021, 259, 118109 CrossRef CAS.
  403. S. Subudhi, L. Paramanik, S. Sultana, S. Mansingh, P. Mohapatra and K. Parida, J. Colloid Interface Sci., 2020, 568, 89–105 CrossRef CAS PubMed.
  404. W. Zhang, Y. Meng, Y. Liu, H. Shen, Z. Ni, S. Xia, W. Han, Y. Li and H. Tang, J. Environ. Chem. Eng., 2022, 10, 107812 CrossRef CAS.
  405. M. Wang, H. Yu, P. Wang, Z. Chi, Z. Zhang, B. Dong, H. Dong, K. Yu and H. Yu, Sep. Purif. Technol., 2021, 274, 118692 CrossRef CAS.
  406. S. L. Prabavathi, K. Govindan, K. Saravanakumar, A. Jang and V. Muthuraj, J. Ind. Eng. Chem., 2019, 80, 558–567 CrossRef CAS.
  407. N. Yin, H. Chen, X. Yuan, Y. Zhang, M. Zhang, J. Guo, Y. Zhang, L. Qiao, M. Liu and K. Song, J. Hazard. Mater., 2022, 436, 129317 CrossRef CAS PubMed.
  408. Z. Zhao, Q. Ling, Z. Li, K. Yan, C. Ding, P. Chen, L. Yang, Z. Sun and M. Zhang, Sep. Purif. Technol., 2023, 308, 122928 CrossRef CAS.
  409. J. Jiang and Y. Li, Crystals, 2021, 11, 1173 CrossRef CAS.
  410. J. Zhang, Z. Zhu, J. Jiang and H. Li, Molecules, 2020, 25, 3706 CrossRef CAS PubMed.
  411. Y. Lv, H. Liu, D. Jin, H. Yang, D. He, Z. Zhang, Y. Zhang, J. Qu and Y.-N. Zhang, Chem. Eng. J., 2022, 429, 132092 CrossRef CAS.
  412. L. Chang, Y. Pu, P. Jing, J. Ji, X. Wei, B. Cao, Y. Yu, S. Xu and H. Xie, Surf. Interfaces, 2022, 31, 102010 CrossRef CAS.
  413. A. Kumar, S. K. Sharma, G. Sharma, A. H. Al-Muhtaseb, M. Naushad, A. A. Ghfar and F. J. Stadler, J. Hazard. Mater., 2019, 364, 429–440 CrossRef CAS PubMed.
  414. J. Li, Z. Xia, D. Ma, G. Liu, N. Song, D. Xiang, Y. Xin, G. Zhang and Q. Chen, J. Colloid Interface Sci., 2021, 586, 243–256 CrossRef CAS PubMed.
  415. C. Zhang, M. Jia, Z. Xu, W. Xiong, Z. Yang, J. Cao, H. Peng, H. Xu, Y. Xiang and Y. Jing, Chem. Eng. J., 2022, 430, 132652 CrossRef CAS.
  416. X. Gu, T. Chen, J. Lei, Y. Yang, X. Zheng, S. Zhang, Q. Zhu, X. Fu, S. Meng and S. Chen, Chin. J. Catal., 2022, 43, 2569–2580 CrossRef CAS.
  417. Y. Liu, R. Chen, F. Wu, W. Gan, L. Chen, M. Zhang and Z. Sun, Sep. Purif. Technol., 2023, 324, 124572 CrossRef.
  418. J. Zhang, Z. Zhu, J. Jiang and H. Li, Catalysts, 2020, 10, 373 CrossRef CAS.
  419. Z. Zhu, H. Xia, H. Li and S. Han, Inorganics, 2022, 10, 131 CrossRef CAS.
  420. C. Li, T. Sun, G. Yi, D. Zhang, Y. Zhang, X. Lin, J. Liu, Z. Shi and Q. Lin, Colloids Surf., A, 2023, 662, 131001 CrossRef CAS.
  421. C. Guo, S. Gao, J. Lv, S. Hou, Y. Zhang and J. Xu, Appl. Catal., B, 2017, 205, 68–77 CrossRef CAS.
  422. A. Behera, D. Kandi, S. Sahoo and K. Parida, J. Phys. Chem. C, 2019, 123, 17112–17126 CrossRef CAS.
  423. P. C. Sharma, A. Jain, S. Jain, R. Pahwa and M. S. Yar, J. Enzyme Inhib. Med. Chem., 2010, 25, 577–589 CrossRef CAS PubMed.
  424. Y. Zeng, D. Chen, T. Chen, M. Cai, Q. Zhang, Z. Xie, R. Li, Z. Xiao, G. Liu and W. Lv, Chemosphere, 2019, 227, 198–206 CrossRef CAS PubMed.
  425. W. Li, C. Guo, B. Su and J. Xu, J. Chem. Technol. Biotechnol., 2012, 87, 643–650 CrossRef CAS.
  426. R. Shetty, G. Kothari, A. Tambe, B. D. Kulkarni and S. P. Kamble, Indian J. Chem., Sect. A: Inorg., Bio-inorg., Phys., Theor. Anal. Chem., 2016, 55, 16–22 Search PubMed.
  427. T. A. Gad-Allah, M. E. M. Ali and M. I. Badawy, J. Hazard. Mater., 2011, 186, 751–755 CrossRef CAS PubMed.
  428. S. Li and J. Hu, Water Res., 2018, 132, 320–330 CrossRef CAS PubMed.
  429. A. Salma, S. Thoröe-Boveleth, T. C. Schmidt and J. Tuerk, J. Hazard. Mater., 2016, 313, 49–59 CrossRef CAS PubMed.
  430. A. R. Silva, P. M. Martins, S. Teixeira, S. A. C. Carabineiro, K. Kuehn, G. Cuniberti, M. M. Alves, S. Lanceros-Mendezagh and L. Pereira, RSC Adv., 2016, 6, 95494–95503 RSC.
  431. Y. Gan, Y. Wei, J. Xiong and G. Cheng, Chem. Eng. J., 2018, 349, 1–16 CrossRef CAS.
  432. Y. Li, Y. Fu and M. Zhu, Appl. Catal., B, 2020, 260, 118149 CrossRef CAS.
  433. M. R. Usman, A. Prasasti, S. Fajriyah, A. W. Marita, S. Islamiah, A. N. Firdaus, A. R. Noviyanti and D. R. Eddy, Bull. Chem. React. Eng. Catal., 2021, 16, 752–762 CrossRef CAS.
  434. M. El-Kemary, H. El-Shamy and I. El-Mehasseb, J. Lumin., 2010, 130, 2327–2331 CrossRef CAS.
  435. M. Eskandari, N. Goudarzi and S. G. Moussavi, Water Environ. J., 2018, 32, 58–66 CrossRef CAS.
  436. A. Ulyankina, T. Molodtsova, M. Gorshenkov, I. Leontyev, D. Zhigunov, E. Konstantinova, T. Lastovina, J. Tolasz, J. Henych, N. Licciardello, G. Cuniberti and N. Smirnova, J. Water Process. Eng., 2021, 40, 101809 CrossRef.
  437. S. Aghdasi and M. Shokri, Iran. J. Catal., 2016, 6, 481–487 CAS.
  438. F. Nekouei and S. Nekouei, Sci. Total Environ., 2017, 601–602, 508–517 CrossRef CAS PubMed.
  439. C. Bojer, J. Schöbel, T. Martin, M. Ertl, H. Schmalz and J. Breu, Appl. Catal., B, 2017, 204, 561–565 CrossRef CAS.
  440. V. Tiron, M. A. Ciolan, G. Bulai, G. Mihalache, F. D. Lipsa and R. Jijie, Nanomaterials, 2022, 12, 2193 CrossRef CAS PubMed.
  441. N. Finčur, P. Sfîrloagă, P. Putnik, V. Despotović, M. Lazarević, M. Uzelac, B. Abramović, P. Vlazan, C. Ianăs, T. Alapi, M. Náfrádi, I. Maksimović, M. Putnik-Delić and D. Š. Merkulov, Nanomaterials, 2021, 11, 215 CrossRef PubMed.
  442. M. Mahjoore, A. Aryafar and M. Honarmand, Journal of Mining and Engineering, 2022, 13, 155–164 Search PubMed.
  443. T. S. Roy, S. Akter, M. R. Fahim, M. A. Gafur and T. Ferdous, Heliyon, 2023, 9, e13130 CrossRef CAS PubMed.
  444. T. Suwannaruang, J. P. Hildebrand, D. H. Taffa, M. Wark, K. Kamonsuangkasem, P. Chirawatkul and K. Wantala, J. Photochem. Photobiol., A, 2020, 391, 112371 CrossRef CAS.
  445. C. A. Huerta-Aguilar, Y. S. García Gutiérrez and P. Thangarasu, Chem. Eng. J., 2020, 394, 124286 CrossRef CAS.
  446. Y. Gan, M. Zhang, J. Xiong, J. Zhu, W. Li, C. Zhang and G. Cheng, J. Taiwan Inst. Chem. Eng., 2019, 96, 229–242 CrossRef CAS.
  447. J. C. Durán-Álvarez, E. Avella, R. M. Ramírez-Zamora and R. Zanella, Catal. Today, 2016, 266, 175–187 CrossRef.
  448. T. Suwannaruang, P. Kidkhunthod, N. Chanlek, S. Soontaranon and K. Wantala, Appl. Surf. Sci., 2019, 478, 1–14 CrossRef CAS.
  449. X. Xing, Z. Du, J. Zhuang and D. Wang, J. Photochem. Photobiol., A, 2018, 359, 23–32 CrossRef CAS.
  450. X. Feng, P. Wang, J. Hou, J. Qian, Y. Ao and C. Wang, J. Hazard. Mater., 2018, 351, 196–205 CrossRef CAS PubMed.
  451. S. Vijayalakshmi, E. Kumar, P. S. Venkatesh and A. Raja, Ionics, 2020, 26, 1507–1513 CrossRef CAS.
  452. D. V. Thuan, T. B. H. Nguyen, T. H. Pham, J. Kim, T. T. Hien Chu, M. V. Nguyen, K. D. Nguyen, W. A. Al-onazi and M. S. Elshikh, Chemosphere, 2022, 308, 136408 CrossRef PubMed.
  453. Z.-H. Diao, X.-R. Xu, D. Jiang, J.-J. Liu, L.-J. Kong, G. Li, L.-Z. Zuo and Q.-H. Wu, Chem. Eng. J., 2017, 315, 167–176 CrossRef CAS.
  454. S. Das, S. Ghosh, A. J. Misra, A. J. Tamhankar, A. Mishra, C. S. Lundborg and S. K. Tripathy, Int. J. Environ. Res. Public Health, 2018, 15, 2440 CrossRef CAS PubMed.
  455. X. Yu, J. Zhang, J. Zhang, J. Niu, J. Zhao, Y. Wei and B. Yao, Chem. Eng. J., 2019, 374, 316–327 CrossRef CAS.
  456. A. Tofanello, E. Belleti, A. M. M. Brito and I. L. Nantes-Cardoso, Mater. Res., 2021, 24 DOI:10.1590/1980-5373-mr-2021-0198.
  457. L. T. Nguyen, H. T. Nguyen, T. D. Pham, T. D. Tran, H. T. Chu, H. T. Dang and B. V. Bruggen, Top. Catal., 2020, 63, 985–995 CrossRef CAS.
  458. X. Huang, W. Yang, G. Zhang, L. Yan, Y. Zhang, A. Jiang, H. Xu, M. Zhou, Z. Liu, H. Tang and D. D. Dionysiou, Catal. Today, 2021, 361, 11–16 CrossRef CAS.
  459. M. Mirzai, F. Akhlaghian and F. Rahmani, Water Environ. J., 2020, 34, 420–431 CrossRef CAS.
  460. X. Zheng, S. Xu, Y. Wang, X. Sun, Y. Gao and B. Gao, J. Colloid Interface Sci., 2018, 527, 202–213 CrossRef CAS PubMed.
  461. A. Alam, U. Rahman, Z. U. Rahman, S. A. Khan, Z. Shah, K. Shaheen, H. Suo, M. N. Qureshi, S. B. Khan, E. M. Bakhsh and K. Akhtar, J. Mater. Sci.: Mater. Electron., 2022, 33, 4255–4267 CrossRef CAS.
  462. Z. Liu, X. Liu, Q. Lu, Q. Wang and Z. Ma, J. Taiwan Inst. Chem. Eng., 2019, 96, 214–222 CrossRef CAS.
  463. A. Hassani, A. Khataee and S. Karacaa, J. Mol. Catal. A: Chem., 2015, 409, 149–161 CrossRef CAS.
  464. N. Li, J. Zhang, Y. Tian, J. Zhao, J. Zhang and W. Zuo, Chem. Eng. J., 2017, 308, 377–385 CrossRef CAS.
  465. A. Raja, P. Rajasekaran, K. Selvakumar, M. Arunpandian, K. Kaviyarasu, S. Asath Bahadur and M. Swaminathan, Sep. Purif. Technol., 2020, 233, 115996 CrossRef CAS.
  466. S. Teixeira, H. Mora, L.-M. Blasse, P. M. Martins, S. A. C. Carabineiro, S. Lanceros-Méndez, K. Kühn and G. Cuniberti, J. Photochem. Photobiol., A, 2017, 345, 27–35 CrossRef CAS.
  467. I. Gabelica, L. Ćurković, V. Mandić, I. Panžić, D. Ljubas and K. Zadro, Catalysts, 2021, 11, 1136 CrossRef CAS.
  468. Z. Liu and Z. Ma, Mater. Res. Bull., 2019, 118, 110492 CrossRef CAS.
  469. A. J. Labrag, C. E. Bekkali, I. Es-Saidi, H. Bouyarmane, A. Laghzizil, M. Khamar and D. Robert, E3S Web Conf., 2020, 150, 02006 CrossRef.
  470. F. Tamaddon, A. Nasiri and G. Yazdanpanah, MethodsX, 2020, 7, 100764 CrossRef CAS PubMed.
  471. Z. Liu, J. Tian, D. Zeng, C. Yu, W. Huang, K. Yang, X. Liu and H. Liu, Mater. Res. Bull., 2019, 112, 336–345 CrossRef CAS.
  472. H. Wang, J. Li, P. Huo, Y. Yan and Q. Guan, Appl. Surf. Sci., 2016, 366, 1–8 CrossRef CAS.
  473. S. P. Pattnaik, A. Behera, S. Martha, R. Acharya and K. Parida, J. Mater. Sci., 2019, 54, 5726–5742 CrossRef CAS.
  474. C. Li, Z. Sun, W. Zhang, C. Yu and S. Zheng, Appl. Catal., B, 2018, 220, 272–282 CrossRef CAS.
  475. C. Chuaicham, T. Inoue, V. Balakumar, Q. Tian, B. Ohtani and K. Sasaki, J. Environ. Chem. Eng., 2022, 10, 10697 Search PubMed.
  476. Y. Yu, K. Liu, Y. Zhang, X. Xing and H. Li, Int. J. Environ. Res. Public Health, 2022, 19, 4793 CrossRef CAS PubMed.
  477. Y. Zhang, H. Pan and F. Zhang, Mater. Lett., 2019, 251, 114–117 CrossRef CAS.
  478. D. T. Sponza and P. Koyuncuoglu, Clin. Microbiol. Infect., 2019, 4, 1–10 Search PubMed.
  479. Y. He, S. Han, G. Zhao, J. Luo, C. Jia, Y. Chen, Q. Liu, J. Gao, C. Wang and J. Wang, J. Environ. Chem. Eng., 2022, 10, 107829 CrossRef CAS.
  480. M. Shi, W. Li, Q. Wang, H. Xu, Y. Zhao, G. He, Q. Meng and H. Chen, Opt. Mater., 2021, 122, 111726 CrossRef CAS.
  481. S. Kumar, R. D. Kaushik and P. Purohit, J. Hazard. Mater., 2022, 424, 127332 CrossRef CAS PubMed.
  482. J. Ni, J. Xue, J. Shen, G. He and H. Chen, Appl. Surf. Sci., 2018, 441, 599–606 CrossRef CAS.
  483. X.-F. Zeng, J.-S. Wang, Y.-N. Zhao, W.-L. Zhang and M.-H. Wang, Int. J. Miner., Metall. Mater., 2021, 28, 503–510 CrossRef CAS.
  484. S. Urus, M. Caylar, H. Eskalen and S. Ozganm, J. Mater. Sci.: Mater. Electron., 2022, 33, 4314–4329 CrossRef CAS.
  485. X. Hou, J. Liu, W. Guo, S. Li, Y. Guo, Y. Shi and C. Zhang, Catal. Commun., 2019, 121, 27–31 CrossRef CAS.
  486. A. Kumar, G. Sharma, M. Naushad, T. Ahamad, R. C. Veses and F. J. Stadler, Chem. Eng. J., 2019, 370, 148–165 CrossRef CAS.
  487. P. Huo, M. Zhou, Y. Tang, X. Liu, C. Ma, L. Yu and Y. Yan, J. Alloys Compd., 2016, 670, 198–209 CrossRef CAS.
  488. J. Du, S. Ma, Y. Yan, K. Li, F. Zhao and J. Zhou, Colloids Surf., A, 2019, 572, 237–249 CrossRef CAS.
  489. Y. Deng, L. Tang, C. Feng, G. Zeng, J. Wang, Y. Zhou, Y. Liu, B. Peng and H. Feng, J. Hazard. Mater., 2018, 344, 758–769 CrossRef CAS PubMed.
  490. X. Rong, F. Qiu, Z. Jiang, J. Rong, J. Pan, T. Zhang and D. Yang, Chem. Eng. Res. Des., 2016, 111, 253–261 CrossRef CAS.
  491. X. Wang, A. Wang and J. Ma, J. Hazard. Mater., 2017, 336, 81–92 CrossRef CAS PubMed.
  492. Z. Wu, Y. Liang, X. Yuan, D. Zou, J. Fang, L. Jiang, J. Zhang, H. Yang and Z. Xiao, Chem. Eng. J., 2020, 394, 124921 CrossRef CAS.
  493. N. Guo, Y. Zeng, H. Li, X. Xu, H. Yu and X. Han, J. Hazard. Mater., 2018, 353, 80–88 CrossRef CAS PubMed.
  494. A. Kaur, W. A. Anderson, S. Tanvir and S. K. Kansal, J. Colloid Interface Sci., 2019, 557, 236–253 CrossRef CAS PubMed.
  495. J. Zhao, Z. Zhao, N. Li, J. Nan, R. Yu and J. Du, Chem. Eng. J., 2018, 353, 805–813 CrossRef CAS.
  496. S. Shanavas, S. M. Roopan, A. Priyadharsan, D. Devipriya, S. Jayapandi, R. Acevedo and P. M. Anbarasan, Appl. Catal., B, 2019, 255, 117758 CrossRef CAS.
  497. M. Chen, J. Yao, Y. Huang, H. Gong and W. Chu, Chem. Eng. J., 2018, 334, 453–461 CrossRef CAS.
  498. T. Senasu, S. Nijpanich, S. Juabrum, N. Chanlek and S. Nanan, Appl. Surf. Sci., 2021, 567, 150850 CrossRef CAS.
  499. F. Beshkar, M. Salavati-Niasari and O. Amiri, Ind. Eng. Chem. Res., 2021, 60, 9578–9591 CrossRef CAS.
  500. M. Yu, H. Liang, R. Zhan, L. Xu and J. Niu, Chin. Chem. Lett., 2021, 32, 2155–2158 CrossRef CAS.
  501. A. S. Prasad, S. N. Kumar, M. A. Maheswari and D. Prabhakaran, Bull. Mater. Sci., 2021, 44, 99 CrossRef CAS.
  502. J. Yi, K. Wu, H. Wu, J. Guo, L. Zhang, H. Li and J. Lim, Nano, 2021, 16, 2150063 CrossRef CAS.
  503. K. K. Das, S. Patnaik, S. Mansingh, A. Behera, A. Mohanty, C. Acharya and K. M. Parida, J. Colloid Interface Sci., 2020, 561, 551–567 CrossRef CAS PubMed.
  504. L. N. Costa, F. X. Nobre, A. O. Lobo and J. M. E. de Matos, Environ. Nanotechnol., Monit. Manage., 2021, 16, 100466 CAS.
  505. M. Li, J. Zhang, L. Wang, X. Cheng, X. Gao, Y. Wang, G. Zhang, Y. Qi, H. Zhai, R. Guan and Z. Zhao, Appl. Surf. Sci., 2022, 583, 152516 CrossRef CAS.
  506. L. Wolski, K. Grzelak, M. Muńko, M. Frankowski, T. Grzyb and G. Nowaczyk, Appl. Surf. Sci., 2021, 563, 150338 CrossRef CAS.
  507. X. Hu, H. Zhao, Y. Liang, F. Chen, J. Li and R. Chen, Chemosphere, 2021, 264, 128434 CrossRef CAS PubMed.
  508. V. D. Dang, J. A. Jr, T. Annadurai, T. A. N. Bui, H. L. Tran, L.-Y. Lin and R.-A. Doong, Chem. Eng. J., 2021, 422, 130103 CrossRef CAS.
  509. K. Hu, R. Li, C. Ye, A. Wang, W. Wei, D. Hu, R. Qiu and K. Yan, J. Cleaner Prod., 2020, 253, 120055 CrossRef CAS.
  510. M. Chen, Y. Dai, J. Guo, H. Yang, D. Liu and Y. Zhai, Appl. Surf. Sci., 2019, 493, 1361–1367 CrossRef CAS.
  511. X.-J. Wen, C.-G. Niu, L. Zhang, C. Liang, H. Guo and G.-M. Zeng, J. Catal., 2018, 358, 141–154 CrossRef CAS.
  512. J. C. Durán-Álvarez, M. Méndez-Galván, L. Lartundo-Rojas, M. Rodríguez-Varela, D. Ramírez-Ortega, D. Guerrero-Araque and R. Zanella, Top. Catal., 2019, 62, 1011–1025 CrossRef.
  513. J. Mao, B. Hong, J. Wei, J. Xi, Y. Han, H. Jin, D. Jin, X. Peng, J. Li, Y. Yang, J. Gong, H. Ge and X. Wang, ChemistrySelect, 2019, 4, 13716–13723 CrossRef CAS.
  514. B. Zhu, D. Song, T. Jia, W. Sun, D. Wang, L. Wang, J. Guo, L. Jin, L. Zhang and H. Tao, ACS Omega, 2021, 6, 1647–1656 CrossRef CAS PubMed.
  515. Y. Zhou, M. Yu, H. Liang, J. Chen, L. Xu and J. Niu, Appl. Catal., B, 2021, 291, 120105 CrossRef CAS.
  516. N. S. Alhokbany, R. Mousa, M. Naushad, S. M. Alshehri and T. Ahamad, Int. J. Biol. Macromol., 2020, 164, 3864–3872 CrossRef CAS PubMed.
  517. M. F. R. Samsudin, C. Frebillot, Y. Kaddoury, S. Sufian and W.-J. Ong, J. Environ. Manage., 2020, 270, 110803 CrossRef CAS PubMed.
  518. C.-H. Shen, X.-J. Wen, Z.-H. Fei, Z.-T. Liu and Q.-M. Mu, Chem. Eng. J., 2020, 391, 123612 CrossRef CAS.
  519. K. Wang, Y. Li, G. Zhang, J. Li and X. Wu, Appl. Catal., B, 2019, 240, 39–49 CrossRef CAS.
  520. S. Li, J. Chen, S. Hu, H. Wang, W. Jiang and X. Chen, Chem. Eng. J., 2020, 402, 126165 CrossRef CAS.
  521. A. A. Tessema, C.-M. Wu and K. G. Motora, ACS Omega, 2022, 7, 38475–38486 CrossRef CAS PubMed.
  522. A. B. Makama, A. Salmiaton, E. B. Saion, T. S. Y. Choong and N. Abdullah, J. Nanomater., 2015, 108297,  DOI:10.1155/2015/108297.
  523. N. Lu, P. Wang, Y. Su, H. Yu, N. Liu and X. Quan, Chemosphere, 2019, 215, 444–453 CrossRef CAS PubMed.
  524. C. Lai, M. Zhang, B. Li, D. Huang, G. Zeng, L. Qin, X. Liu, H. Yi, M. Cheng, L. Li, Z. Chen and L. Chen, Chem. Eng. J., 2019, 358, 891–902 CrossRef CAS.
  525. T. H. A. Nguyen, V. T. Le, V.-D. Doan, A. V. Tran, V. C. Nguyen, A.-T. Nguyen and Y. Vasseghian, Mater. Lett., 2022, 308, 131129 CrossRef CAS.
  526. K. Wang, L. Liang, Y. Zheng, H. Li, X. Niu, D. Zhang and H. Fan, New J. Chem., 2021, 45, 16168–16178 RSC.
  527. R. A. Palominos, M. A. Mondaca, A. Giraldo, G. Penuela, M. Perez-Moya and H. D. Mansilla, Catal. Today, 2009, 144, 100–105 CrossRef CAS.
  528. X.-D. Zhu, Y.-J. Wang, R.-J. Sun and D.-M. Zhou, Chemosphere, 2013, 92, 925–932 CrossRef CAS PubMed.
  529. G. H. Safari, M. Hoseini, M. Seyedsalehi, H. Kamani, J. Jaafari and A. H. Mahvi, Int. J. Environ. Sci. Technol., 2015, 12, 603–616 CrossRef CAS.
  530. S. Wu, H. Hu, Y. Lin, J. Zhang and Y. H. Hu, Chem. Eng. J., 2020, 382, 122842 CrossRef CAS.
  531. T. Ikhlef-Taguelmimt, A. Hamiche, I. Yahiaoui, T. Bendellali, H. Lebik-Elhadi, H. A. Amar and F. Aissani-Benissad, Water Sci. Technol., 2020, 82, 1570–1578 CrossRef CAS PubMed.
  532. L. Rimoldi, D. Meroni, G. Cappelletti and S. Ardizzone, Catal. Today, 2017, 281, 38–44 CrossRef CAS.
  533. M. Malakootian, S. N. Asadzadeh, M. Mehdipootarr and D. Kalantar-Neyestanaki, Desalin. Water Treat., 2021, 222, 302–312 CrossRef CAS.
  534. S. J. Olusegun, G. Larrea, M. Osial, K. Jackowska and P. Krysinski, Catalysts, 2021, 11, 1243 CrossRef CAS.
  535. M. Saghi and K. Mahanpoor, Int. J. Ind. Chem., 2017, 8, 297–313 CrossRef CAS.
  536. H. Shi, Q. Wu, L. Jiang, L. Wang, M. Huang, B. Han, Z. Yu and Y. Zuo, Int. J. Electrochem. Sci., 2020, 15, 1539–1547 CrossRef CAS.
  537. S. K. Fanourakis, S. Q. Barroga, R. A. Mathew, J. Pena-Bahamonde, S. M. Louie, J. V. D. Perez and D. F. Rodrigues, J. Environ. Chem. Eng., 2022, 10, 107635 CrossRef CAS.
  538. Y. Jiang, C. Xing, Y. Chen, J. Shi and S. Wang, Environ. Sci. Pollut. Res., 2022, 29, 57656–57668 CrossRef CAS PubMed.
  539. M. Yan, Y. Wu and X. Liu, J. Alloys Compd., 2021, 855, 157548 CrossRef CAS.
  540. O. Linnik, E. Manuilov, S. Snegir, N. Smirnova and A. Eremenko, J. Adv. Oxid. Technol., 2009, 12, 265–270 CAS.
  541. S. Zhao, Y. Yang, R. Lu, Y. Wang, Y. Lu, R. D. Rodriguez, E. Sheremet and J. Chen, J. Environ. Chem. Eng., 2022, 10, 107107 CrossRef CAS.
  542. M. Liu, J. Jiang, H. Tan, B. Chen, J. Ou, H. Wang, J. Sun, L. Liu, F. Wang, J. Gao, C. Liu, F. Peng, Y. Liu and Y. Tu, Nanoscale, 2022, 14, 12804–12813 RSC.
  543. W. Shi, H. Ren, M. Li, K. Xu, Y. Shu, C. Yan and Y. Tang, Chem. Eng. J., 2020, 382, 122876 CrossRef CAS.
  544. H. Xu, Y. Dingi, S. H. Yang, H. Zhang, X. Wang, J. Yuan, M. Long, Q. Zhao and Y. Liu, Bull. Mater. Sci., 2021, 44, 226 CrossRef CAS.
  545. K. Pandi, M. Preeyanghaa, V. Vinesh, J. Madhavan and B. Neppolian, Environ. Res., 2022, 207, 112188 CrossRef CAS PubMed.
  546. T. Zhang, Y. Liu, Y. Rao, X. Li, D. Yuan, S. Tang and Q. Zhao, Chem. Eng. J., 2020, 384, 123350 CrossRef CAS.
  547. Y. Chen and K. Liu, Chem. Eng. J., 2016, 302, 682–696 CrossRef CAS.
  548. K. Zheng, J. Chen, X. Gao, X. Cao, S. Wu and J. Su, Water Sci. Technol., 2021, 84, 1919–1929 CrossRef CAS PubMed.
  549. N. Farhadian, R. Akbarzadeh, M. Pirsaheb, T. C. Jen, Y. Fakhri and A. Asadi, Int. J. Biol. Macromol., 2019, 132, 360–373 CrossRef CAS PubMed.
  550. P. Wang, P.-S. Yap and T.-T. Lim, Appl. Catal., A, 2011, 399, 252–261 CrossRef CAS.
  551. J. Zhang, X. Li, M. Peng, Y. Tang, A. Ke, W. Gan, X. Fu and H. Hao, Mater. Res. Express, 2018, 5, 065008 CrossRef.
  552. H. Wang, D. Wu, X. Li and P. Huo, J. Mater. Sci.: Mater. Electron., 2019, 30, 19126–19136 CrossRef CAS.
  553. A. Moradi, F. Rahmani and M. Khamforoush, Iran. J. Polym. Sci. Technol., 2021, 33, 465–478 Search PubMed.
  554. A. Bembibre, M. Benamara, M. Hjiri, E. Gómez, H. R. Alamri, R. Dhahri and A. Serrà, Chem. Eng. J., 2022, 427, 132006 CrossRef CAS.
  555. D. S. Pattanayak, D. Pal, J. Mishra, C. Thakur and K. L. Wasewar, Environ. Sci. Pollut. Res., 2023, 30, 24919–24926 CrossRef CAS PubMed.
  556. L. Jiang, X. Yuan, G. Zeng, J. Liang, Z. Wu, H. Yu, D. Mo, H. Wang, Z. Xiao and C. Zhou, J. Colloid Interface Sci., 2019, 536, 17–29 CrossRef CAS PubMed.
  557. S. Chen, F. Liu, R. Cui, B. Zhu and X. You, Water Cycle, 2022, 3, 8–17 CrossRef.
  558. T. S. Bui, P. Bansal, B.-K. Lee, T. Mahvelati-Shamsabadi and T. Soltani, Appl. Surf. Sci., 2019, 506, 144184 CrossRef.
  559. G. Li, B. Wang, J. Zhang, R. Wang and H. Liu, Chem. Eng. J., 2020, 391, 123500 CrossRef CAS.
  560. Y. Zhao, H. Qin, Z. Wang, H. Wang, Y. He, Q. Tian, Q. Luo and P. Xu, Environ. Sci. Pollut. Res., 2022, 29, 74062–74080 CrossRef CAS PubMed.
  561. W. Wang, Z. Zeng, G. Zheng, C. Zhang, R. Xiao, C. Zhou, W. Xiong, Y. Yang, L. Lei, Y. Liu, D. Huang, M. C. Y. Cheng, Y. Fu, H. Luo and Y. Zhou, Chem. Eng. J., 2019, 378, 122132 CrossRef CAS.
  562. J. Xu, Y. Chen, M. Chen, J. Wang and L. Wang, Chem. Eng. J., 2022, 442, 136208 CrossRef CAS.
  563. L. Chen, Y. Wang, S. Cheng, X. Zhao, J. Zhang, Z. Ao, C. Zhao, B. Li, S. Wang, S. Wang and H. Sun, Appl. Catal., B, 2022, 303, 120932 CrossRef CAS.
  564. N. L. M. Tri, J. Kim, B. L. Giang, T. M. A. Tahtamouni, P. T. Huong, C. Lee, N. M. Viet and D. Q. Trung, J. Ind. Eng. Chem., 2019, 80, 597–605 CrossRef.
  565. D. Jia, Y. Zhang, X. Zhang, P. Feng, L. Yang, R. Ning, H. Pan and Y. Miao, Environ. Sci.: Nano, 2021, 8, 415–431 RSC.
  566. S. J. Alyani, A. E. Pirbazari, F. E. Khalilsaraei, N. A. Kolur and N. Gilani, J. Alloys Compd., 2019, 799, 169–182 CrossRef.
  567. J. Ye, J. Liu, Z. Huang, S. Wu, X. Dai, L. Zhang and L. Cui, Chemosphere, 2019, 227, 505–513 CrossRef CAS PubMed.
  568. M. Cao, P. Wang, Y. Ao, C. Wang, J. Hou and J. Qian, J. Colloid Interface Sci., 2016, 467, 129–139 CrossRef CAS PubMed.
  569. J. Liu, H. Chen, C. Zhu, S. Han, J. Li, S. She and X. Wu, J. Colloid Interface Sci., 2022, 622, 526–538 CrossRef CAS PubMed.
  570. N. Belhoucheta, B. Hamdia, H. Chenchounid and Y. Bessekhoua, J. Photochem. Photobiol., A, 2019, 372, 196–205 CrossRef.
  571. Y. M. Hunge, A. A. Yadav, S.-W. Kang and H. Kim, J. Colloid Interface Sci., 2022, 606, 454–463 CrossRef CAS PubMed.
  572. P. Semeraro, S. Bettini, S. Sawalha, S. Pal, A. Licciulli, F. Marzo, N. Lovergine, L. Valli and G. Giancane, Nanomaterials, 2020, 10, 1458 CrossRef CAS PubMed.
  573. B. Kakavandi, N. Bahari, R. R. Kalantary and E. D. Fard, Ultrason. Sonochem., 2019, 55, 75–85 CrossRef CAS PubMed.
  574. V. H. T. Thi and B.-K. Lee, J. Hazard. Mater., 2017, 324, 329–339 CrossRef PubMed.
  575. H. Shi, Q. Wu, X. Yang, Y. Zuo, H. Yang, R. Zhang, Y. Zhang, Y. Fan, X. Du and L. Jiang, Int. J. Electrochem. Sci., 2021, 16, 210314,  DOI:10.20964/2021.03.70.
  576. S. Zhang, J. Yi, J. Chen, Z. Yin, T. Tang, W. Wei, S. Cao and H. Xu, Chem. Eng. J., 2020, 380, 122583 CrossRef CAS.
  577. E. Weidner, K. S. Ciesielczyk, D. Moszynsky, T. Jesionowski and F. Ciesielczyk, Environ. Technol. Innovation, 2021, 24, 102016 CrossRef CAS.
  578. S. Leong, D. Li, K. Hapgood, X. Zhang and H. Wang, Appl. Catal., B, 2016, 198, 224–233 CrossRef CAS.
  579. C. Lv, X. Lan, L. Wang, X. Dai, M. Zhang, J. Cui, S. Yuan, S. Wang and J. Shi, Environ. Technol., 2021, 42, 377–387 CrossRef CAS PubMed.
  580. W. Chen, L. Chang, S.-B. Ren, Z.-C. He, G.-B. Huang and X.-H. Liu, J. Hazard. Mater., 2020, 384, 121308 CrossRef CAS PubMed.
  581. L. Wang, C. Zhang, R. Cheng, J. Ali, Z. Wang, G. Mailhot and G. Pan, Catalysts, 2018, 8, 628 CrossRef.
  582. J. Xue, S. Ma, Y. Zhou, Z. Zhang and P. Jiang, RSC Adv., 2015, 5, 18832–18840 RSC.
  583. J. Wei, P. Zhu and P. Chen, Materials, 2022, 15, 1963 CrossRef CAS PubMed.
  584. H. A. Patehkhor, M. Fattahi and M. R. Khosravi-Nikou, Sci. Rep., 2021, 11, 24177 CrossRef PubMed.
  585. Y. Shi, Z. Yang, B. Wang, H. An, Z. Chen and H. Cui, Appl. Clay Sci., 2016, 119, 311–320 CrossRef CAS.
  586. S. Kaushal, A. Kumar, H. Bains and P. P. Singh, Environ. Sci. Pollut. Res., 2023, 30, 37092–37104 CrossRef CAS PubMed.
  587. X. Zhang, H. Wang, M. Gao, P. Zhao, W. Xia, R. Yang, Y. Huang, L. Wang, M. Liu, T. Wei, L. Wang, R. Yao, X. Li and Z. Fan, Chemosphere, 2022, 294, 133782 CrossRef CAS PubMed.
  588. Y. Wang, S. Zhang, Y. Ge, C. Wang, J. Hu and H. Liu, Acta Phys.-Chim. Sin., 2020, 36, 1905083 Search PubMed.
  589. C. Hu, Z.-T. Liu, P.-C. Yang, Y.-X. Ding, K.-Y. A. Lin and B.-S. Nguven, J. Taiwan Inst. Chem. Eng., 2021, 123, 219–227 CrossRef CAS.
  590. H. Shi, Y. He, Y. Li, T. He and P. Luo, Sep. Purif. Technol., 2022, 280, 119864 CrossRef CAS.
  591. D. Wang, S. Li and Q. Feng, J. Mater. Sci.: Mater. Electron., 2018, 29, 9380–9386 CrossRef CAS.
  592. J. Zhang and Z. Ma, Chin. J. Chem. Eng., 2018, 26, 753–760 CrossRef CAS.
  593. J. Liu, H. Xu, Y. Xu, Y. Song, J. Lian, Y. Zhao, L. Wang, L. Huang, H. Ji and H. Li, Appl. Catal., B, 2017, 207, 429–437 CrossRef CAS.
  594. V. T. Quyen, H. J. Kim, J. T. Kim, L. T. T. Ha, P. T. Huong, D. M. Thanh, N. M. Viet and P. Q. Thang, Sol. Energy, 2021, 214, 288–293 CrossRef.
  595. L. Jiang, X. Yuan, G. Zeng, Z. Wu, J. Liang, X. Chen, L. Leng, H. Wang and H. Wang, Appl. Catal., B, 2018, 221, 715–725 CrossRef CAS.
  596. X. Chen, L. Liu, Y. Zhao, J. Zhang, D. Li, B. Hu and X. Hai, ChemistrySelect, 2017, 2, 9256–9260 CrossRef CAS.
  597. J. Liu, Y. Song, H. Xu, X. Zhu, J. Lian, Y. Xu, Y. Zhao, L. Huang, H. Ji and H. Li, J. Colloid Interface Sci., 2017, 494, 38–46 CrossRef CAS PubMed.
  598. L. Jiang, X.-Z. Yuan, G. Zeng, X. Chen, Z. Wu, J. Liang, J. Zhang, H. Wang and H. Wang, ACS Sustainable Chem. Eng., 2017, 5, 5831–5841 CrossRef CAS.
  599. R. Chen, X. Wang and L. Zhu, Research Square, 2022, preprint,  DOI:10.21203/rs.3.rs-1172368/v1.
  600. E. Vijayakumar, M. G. Raj, M. G. Narendran, R. Preetha, R. Mohankumar, B. Neppolian and A. J. Bosco, ACS Omega, 2022, 7, 5079–5095 CrossRef CAS PubMed.
  601. X. Tang, L. Ni, J. Han and Y. Wang, Chin. J. Catal., 2017, 38, 447–457 CrossRef CAS.
  602. W. Qian, W. Hu, Z. Jiang, Y. Wu, Z. Li, Z. Diao and M. Li, Catalysts, 2022, 12, 774 CrossRef CAS.
  603. F. Wang, Z. Zhu and J. Guo, RSC Adv., 2021, 11, 35663–35672 RSC.
  604. H. Shi, T. Zhao, J. Wang, Y. Wang, Z. Chen, B. Liu, H. Ji, W. Wang, G. Zhang and Y. Li, J. Alloys Compd., 2021, 860, 157924 CrossRef CAS.
  605. H.-J. Ren, Y.-B. Tang, W.-l. Shi, F.-Y. Chen and Y.-S. Xu, New J. Chem., 2019, 43, 19172–19179 RSC.
  606. R. Zou, T. Xu, X. Lei, Q. Wu and S. Xue, Solid State Sci., 2020, 99, 106067 CrossRef CAS.
  607. A. Maridiroosi, A. R. Mahjoub and H. Fakhr, Int. J. Chem. Biomol. Sci., 2020, 14, 31–36,  DOI:10.5281/zenodo.3669194.
  608. D. Qiao, Z. Li, J. Duan and X. He, Chem. Eng. J., 2020, 400, 125952 CrossRef CAS.
  609. B. Li, X. Yu, X. Yu, R. Du, L. Liu and Y. Zhang, Appl. Surf. Sci., 2019, 478, 991–997 CrossRef CAS.
  610. K. Chakraborty, T. Pal and S. Ghosh, ACS Appl. Nano Mater., 2018, 1, 3137–3144 CrossRef CAS.
  611. X. Tang, Z. Wang and Y. Wang, Chem. Phys. Lett., 2018, 691, 408–414 CrossRef CAS.
  612. K. V. A. Kumar, B. Lakshminarayana, D. Suryakala and C. Subrahmanyam, RSC Adv., 2020, 10, 20494–20503 RSC.
  613. J. Shan, X. Wu, C. Li, J. Hu, Z. Zhang, H. Liu, P. Xia and X. Huang, Research Square, 2022, preprint,  DOI:10.21203/rs.3.rs-1721079/v1.
  614. S. M. Ghoreishian, G. S. R. Raju, E. Pavitra, C. H. Kwak, Y.-K. Han and Y. S. Huh, Appl. Surf. Sci., 2019, 489, 110–122 CrossRef CAS.
  615. S. Zhang, J. Xu, J. Hu, O. Cui and H. Liu, Langmuir, 2017, 33, 5015–5024 CrossRef CAS PubMed.
  616. R. Zou, T. Xu, X. Lei, Q. Wu and S. Xue, Solid State Sci., 2020, 99, 106067 CrossRef CAS.
  617. S. K. Ibrahim, T. Pal and S. Ghosh, AIP Conf. Proc., 2019, 2115, 030188,  DOI:10.1063/1.5113027.
  618. Y. Hou, S. Pu, Q. Shi, S. Mandal, H. Ma, S. Xue, G. Cai and Y. Bai, J. Taiwan Inst. Chem. Eng., 2019, 104, 139–150 CrossRef CAS.
  619. F.-S. Tabatabai-Yazdi, A. E. Pirbazari, F. E. Khalilsaraei, N. A. Kolur and N. Gilani, Water Environ. Res., 2020, 92, 662–676 CrossRef CAS PubMed.
  620. F.-S. Tabatabai-Yazdi, A. E. Pirbazari, F. E. K. Saraei and N. Gilani, Phys. B, 2021, 608, 412869 CrossRef CAS.
  621. L.-L. Qu, N. Wang, Y.-Y. Li, D.-D. Bao, G.-H. Yang and H.-T. Li, J. Mater. Sci., 2017, 52, 8311–8320 CrossRef CAS.
  622. J. Song, X. Wu, M. Zhang, C. Liu, J. Yu, G. Sun, Y. Si and B. Ding, Chem. Eng. J., 2020, 379, 122269 CrossRef CAS.
  623. L. Huang, D. Bao, J. Li and X. Jiang, Appl. Surf. Sci., 2021, 555, 149696 CrossRef CAS.
  624. A. Iqbal, U. Saidu, F. Adam, S. Sreekantan, N. Yahaya, M. N. Ahmad, R. J. Ramalingam and L. D. Wilson, Molecules, 2021, 26, 2509 CrossRef CAS PubMed.
  625. Y. Zhao, A. Zada, Y. Yang, J. Pan, Y. Wang, Z. Yan, Z. Xu and K. Qi, Front. Chem., 2021, 9, 797738 CrossRef CAS PubMed.
  626. H. Jingyu, Y. Ran, L. Zhaohui, S. Yuanqiang, Q. Lingbo and A. N. Kani, Solid State Sci., 2019, 92, 60–67 CrossRef.
  627. B. Guo, B. Liu, C. Wang, Y. Wang, S. Yin, M. S. Javed and W. Han, J. Environ. Chem. Eng., 2022, 10, 107118 CrossRef CAS.
  628. P. Zhu, M. Duan, R. Wang, J. Xu, P. Zou and H. Jia, Colloids Surf., A, 2020, 602, 125118 CrossRef CAS.
  629. J. Gu, H. Jia, S. Ma, Z. Ye, J. Pan, R. Dong, Y. Zong and J. Xue, ACS Omega, 2020, 5, 30980–30988 CrossRef CAS PubMed.
  630. X. Liao, T.-T. Li, H.-T. Ren, Z. Mao, X. Zhang, J.-H. Lin and C.-W. Lou, Ceram. Int., 2021, 47, 10786–10795 CrossRef CAS.
  631. S. Li, S. Hu, K. Xu, W. Jiang, Y. Liu, Z. Leng and J. Liu, J. Colloid Interface Sci., 2017, 504, 561–569 CrossRef CAS PubMed.
  632. D. Chen, S. Wu, J. Fang, S. Lu, G. Y. Zhou, W. Feng, F. Yang, Y. Chen and Z. Q. Fang, Sep. Purif. Technol., 2018, 193, 232–241 CrossRef CAS.
  633. W. Wang, Q. Han, Z. Zhu, L. Zhang, S. Zhong and B. Liu, Adv. Powder Technol., 2019, 30, 1882–1896 CrossRef CAS.
  634. Y. Zhou, J. Li, C. Liu, P. Huo and H. Wang, Appl. Surf. Sci., 2018, 458, 586–596 CrossRef CAS.
  635. Z. Liu, A. Zhang, Y. Liu, Y. Fu and Y. Du, Colloids Surf., A, 2022, 643, 128818 CrossRef CAS.
  636. H. Liu, W. Huo, T. C. Zhang, L. Ouyang and S. Yuan, Mater. Today Chem., 2022, 23, 100729 CrossRef CAS.
  637. M. Al-Haddad, A. Shawky and I. A. Mkhalid, J. Taiwan Inst. Chem. Eng., 2021, 123, 284–292 CrossRef CAS.
  638. H. M. Lwin, W. Zhan, S. Song, F. Jia and J. Zhou, Chem. Phys. Lett., 2019, 736, 136806 CrossRef CAS.
  639. X. Zhang, D. Han, M. Dai, K. Chen, Z. Han, Y. Fan, Y. He, D. Han and L. Niu, Catal. Sci. Technol., 2021, 11, 6248 RSC.
  640. C. Jin, J. Kang, Z. Li, M. Wang, Z. Wu and Y. Xie, Appl. Surf. Sci., 2020, 514, 146076 CrossRef CAS.
  641. Q. Chen, W. Yang, J. Zhu, L. Fu, D. Li and L. Zhou, J. Hazard. Mater., 2020, 384, 121275 CrossRef CAS PubMed.
  642. A. O. Oluwole and O. S. Olatunji, Environ. Sci. Eur., 2022, 34, 5 CrossRef CAS.
  643. S. Kumar, V. Sharma, K. Bhattacharyya and V. Krishnan, Mater. Chem. Front., 2017, 1, 1093–1106 RSC.
  644. T. Xiao, Z. Tang, Y. Yang, L. Tang, Y. Zhou and Z. Zou, Appl. Catal., B, 2018, 220, 417–428 CrossRef CAS.
  645. C. Li, S. Yu, H. Dong, C. Liu, H. Wu, H. Che and G. Chen, Appl. Catal., B, 2018, 238, 284–293 CrossRef CAS.
  646. X. Jiang, S. Lai, W. Xu, J. Fang, X. Chen, J. Beiyuan, X. Zhou, K. Lin, J. Liu and G. Guan, J. Alloys Compd., 2019, 809, 151804 CrossRef CAS.
  647. J. Ni, W. Wang, D. Liu, Q. Zhu, J. Jia, J. Tian, Z. Li, X. Wang and Z. Xing, J. Hazard. Mater., 2021, 408, 124432 CrossRef CAS PubMed.
  648. Y. Liu, J. Tian, L. Wei, Q. Wang, C. Wang, Z. Xing, X. Li, W. Yang and C. Yang, Sep. Purif. Technol., 2021, 257, 117976 CrossRef CAS.
  649. C. Li, S. Yu, H. Che, X. Zhang, J. Han, Y. Mao, Y. Wang, C. Liu and H. Dong, ACS Sustainable Chem. Eng., 2018, 6, 16437–16447 CrossRef CAS.
  650. Y. Hong, Y. Meng, G. Zhang, B. Yin, Y. Zhao, W. Shi and C. Li, Sep. Purif. Technol., 2016, 171, 229–237 CrossRef CAS.
  651. Z. Yang, X. Xia, L. Shao, L. Wang and Y. Liu, Chem. Eng. J., 2021, 410, 128454 CrossRef CAS.
  652. Y. Li, Z. Lai, Z. Huang, H. Wang, C. Zhao, G. Ruan and F. Du, Appl. Surf. Sci., 2021, 550, 149342 CrossRef CAS.
  653. C. Du, Z. Zhang, S. Tan, G. Yu, H. Chen, L. Zhou, L. Yu, Y. Su, Y. Zhang, F. Deng and S. Wang, Environ. Res., 2021, 200, 111427 CrossRef CAS PubMed.
  654. L. Yan, W. Li, Q. Zhao, Z. Zhu, C. Hu and B. Liu, Colloids Surf., A, 2021, 524, 126783 CrossRef.
  655. J. Chen, X. Xiao, Y. Wang, M. Lu and X. Zeng, J. Alloys Compd., 2019, 800, 88–98 CrossRef CAS.
  656. F. Chen, Q. Yang, X. Li, G. Zeng, D. Wang, C. Niu, J. Zhao, H. An, T. Xie and Y. Deng, Appl. Catal., B, 2017, 200, 330–342 CrossRef CAS.
  657. H. Jing, R. Ou, H. Yu, Y. Zhao, Y. Lu, M. Huo, H. Huo and X. Wang, Sep. Purif. Technol., 2021, 255, 117646 CrossRef CAS.
  658. F. Guo, W. Shi, M. Li, Y. Shi and H. Wen, Sep. Purif. Technol., 2019, 210, 608–615 CrossRef CAS.
  659. P. Zhu, M. Hu, M. Duan, L. Xie and M. Zhao, J. Alloys Compd., 2020, 840, 155714 CrossRef CAS.
  660. Z. Zhang, X. Xue and X. Chen, Dalton Trans., 2022, 51, 8015–8027 RSC.
  661. X. Zhou, Y. Chen, C. Li, L. Zhang, X. Zhang, X. Ning, L. Zhan and J. Luo, Sep. Purif. Technol., 2019, 211, 179–188 CrossRef CAS.
  662. B. Zhang, X. He, X. Ma, Q. Chen, G. Liu, Y. Zhou, D. Ma, C. Cui, J. Ma and Y. Xin, Sep. Purif. Technol., 2020, 247, 116932 CrossRef CAS.
  663. L. Lonappan, S. K. Brar, R. K. Das, M. Verma and R. Y. Surampalli, Environ. Int., 2016, 96, 127–138 CrossRef CAS PubMed.
  664. A. S. Mestre and A. P. Carvalho, Molecules, 2019, 24, 3702,  DOI:10.3390/molecules24203702.
  665. K. C. Hui, W. L. Ang, W. Z. N. Yahya and N. S. Sambudi, Chemosphere, 2022, 290, 133377 CrossRef CAS PubMed.
  666. D. Fabbri, M. J. López-Muñoz, A. Daniele, C. Medana and P. Calza, Photochem. Photobiol. Sci., 2019, 18, 845–852 CrossRef CAS PubMed.
  667. L. Rizza, S. Meric, D. Kassinos, M. Guida, F. Russo and V. Belgiorno, Water Res., 2009, 43, 979–988 CrossRef PubMed.
  668. W. Z. Khan, I. Najeeb, S. Ishtiaque and S. Jabeen, International Organization of Scientific Research (IOSR) Journal of Engineering, 2016, 6, 36–46 Search PubMed.
  669. A. C. Khraibet, N. J. Imran, H. M. Majeed, M. A. Ehmood, G. S. Hasoon, Z. F. Nathim, A. K. Alwan and A. S. A. Alsada, J. Genet. Environ. Resour. Conserv., 2022, 10, 41–48 Search PubMed.
  670. U. Schulze-Hennings, I. Brückner, W. Gebhardt, M. Groteklaes, S. P. Blöβ, M. Wett, V. Linnemann, D. Montag and J. Pinnekamp, Water Environ. J., 2017, 31, 508–514 CrossRef CAS.
  671. M. Baniamer, A. Almasi and S. Sharifnia, Iran. J. Chem. Chem. Eng., 2018, 15, 3–16 Search PubMed.
  672. A. Achilleos, E. Hapeshi, N. P. Xekoukoulotakis, D. Mantzavinos and D. Fatta-Kassinos, Chem. Eng. J., 2010, 161, 53–59 CrossRef CAS.
  673. M. V. Bagal and P. R. Gogate, Ultrason. Sonochem., 2014, 21, 1035–1043 CrossRef CAS PubMed.
  674. M. Tanveer, G. Tezcanli, M. T. Sadiq, S. M. Kazmi, N. Noshad, G. Abbas and A. Ali, Eng. Proc., 2022, 12, 76 Search PubMed.
  675. I. Mimouni, A. Bouziani, Y. Naciri, M. Boujnah, M. Alaoui, E. Belghiti and M. E. Azzouzi, Environ. Sci. Pollut. Res., 2022, 29, 7984–7996 CrossRef CAS PubMed.
  676. D. Meroni, C. L. Bianchi, D. C. Boffito, G. Cerrato, A. Bruni, M. A. Sartirana and E. Falletta, Ultrason. Sonochem., 2021, 75, 105615 CrossRef CAS PubMed.
  677. I. Berruti, N. P. F. Gonçalves, P. Calza, M. C. Paganini, I. Oller and M. I. Polo-López, Chemosphere, 2022, 303, 135017 CrossRef CAS PubMed.
  678. H. Mohan, P. M. Sathya, S. Vadivel, G. H. Ha, H. S. Oh, G. Kim, K.-K. Seralathan and T. Shin, Chemosphere, 2022, 301, 134699 CrossRef CAS PubMed.
  679. M. V. Gerbaldo, S. G. Marchetti, V. R. Elías, S. N. Mendieta and M. E. Crivello, Chem. Eng. Res. Des., 2021, 166, 237–247 CrossRef CAS.
  680. M. A. H. Karim and B. K. Aziz, React. Kinet., Mech. Catal., 2022, 135, 1113–1124 CrossRef CAS.
  681. A. Rey, E. Mena, A. M. Chavez, F. J. Beltron and F. Medina, Chem. Eng. Sci., 2015, 126, 80–90 CrossRef CAS.
  682. H. Chakhtouna, H. Benzeid, N. Zari, A. E. K. Qaiss and R. Bouhfid, Environ. Sci. Pollut. Res., 2021, 28, 44638–44666 CrossRef CAS PubMed.
  683. M. R. Espino-Estevez, C. Fernandez-Rodríguez, O. M. Gonzalez-Díaz, J. Arana, J. P. Espinos, J. A. Ortega-Mendez and J. M. Dona-Rodríguez, Chem. Eng. J., 2016, 298, 82–95 CrossRef CAS.
  684. E. Torad, E. H. Ismail, M. M. Mohamed and M. M. H. Khalil, Mater. Res. Bull., 2021, 137, 111193 CrossRef CAS.
  685. V.-H. Nguyena, Q. B. Tran, X. C. Nguyen, L. T. Hai, T. T. T. Ho, M. Shokouhimehr, D.-V. N. Vo, S. S. Lami, V.-H. Nguyena, Q. B. Tranc, X. C. Nguyenc, L. T. Hai, T. T. T. Ho, M. Shokouhimehr, D.-V. N. Vo, S. S. Lam, H. P. Nguyenj, C. T. Hoang, Q. V. Ly, W. Peng, S. Y. Kim, T. V. Tunge and Q. V. Lec, Process Saf. Environ. Prot., 2020, 142, 229–237 CrossRef.
  686. A. Surenjan, B. Sambandam, T. Pradeep and L. Philip, J. Environ. Chem. Eng., 2017, 5, 757–767 CrossRef CAS.
  687. I. Tbessi, M. Benito, J. Llorca, E. Molins, S. Sayadi and W. Najjar, J. Alloys Compd., 2019, 779, 314–325 CrossRef CAS.
  688. L. Rueda-Salaya, A. Hernández-Ramírez, L. Hinojosa-Reyes, J. L. Guzmán-Mar, M. Villanueva-Rodríguez and E. Sánchez-Cervantes, J. Photochem. Photobiol., A, 2020, 391, 112364 CrossRef CAS.
  689. G. Vitiello, G. Iervolino, C. Imparato, I. Rea, F. Borbone, L. D. Stefano, A. Aronne and V. Vaiano, Sci. Total Environ., 2021, 762, 143066 CrossRef CAS PubMed.
  690. S. M. Chaudhari, O. S. Gonsalves and R. Nemade, Mater. Res. Bull., 2021, 143, 111463 CrossRef CAS.
  691. M. Thiruppathi, P. S. Kumar, P. Devendran, C. Ramalingan, M. Swaminathan and E. R. Nagarajan, J. Alloys Compd., 2018, 735, 728–734 CrossRef CAS.
  692. P. Apopei, C. Orha, M. I. Popescu, C. Lazau, F. Manea, C. Catrinescu and C. Teodosiu, Process Saf. Environ. Prot., 2020, 138, 324–336 CrossRef CAS.
  693. W. Buda and B. Czech, Water Sci. Technol., 2013, 68, 1322 CrossRef CAS PubMed.
  694. I. Tbessi, M. Benito, E. Molins, J. Liorca, A. Touati, S. Sayadi and W. Najjar, Solid State Sci., 2019, 88, 20–28 CrossRef CAS.
  695. W. S. Koe, W. C. Chong, Y. L. Pang, C. H. Koo, M. Ebrahim and A. W. Mohammad, J. Water Process. Eng., 2020, 33, 101068 CrossRef.
  696. A. Gil, A. M. García, M. Fernández, M. A. Vicente, B. González-Rodríguez, V. Rives and S. A. Korili, J. Ind. Eng. Chem., 2017, 53, 183–191 CrossRef CAS.
  697. S. Ramandi, M. H. Entezari and N. Ghows, Ultrason. Sonochem., 2017, 38, 234–245 CrossRef CAS PubMed.
  698. E. Mugunthan, M. B. Saidutta and P. E. Jagadeeshbabu, Environ. Nanotechnol., Monit. Manage., 2018, 10, 322–330 Search PubMed.
  699. A. Cordero-García, G. T. Palomino, L. Hinojosa-Reyes, J. L. Guzmán-Mar, L. MayaTeviño and A. Hernández-Ramírez, Environ. Sci. Pollut. Res., 2017, 24, 4613–4624 CrossRef PubMed.
  700. E. Mugunthan, M. B. Saidutta and P. E. Jagadeeshbabu, J. Photochem. Photobiol., A, 2019, 383, 111993 CrossRef CAS.
  701. A. Cordero-García, J. L. Guzmán-Mar, L. Hinojosa-Reyes, E. Ruiz-Ruiz and A. Hernández-Ramírez, Ceram. Int., 2016, 42, 9796–9803 CrossRef.
  702. S. Murgolo, I. S. Moreira, C. Piccirillo, P. M. L. Castro, G. Ventrella, C. Cocozza and G. Mascolo, Materials, 2018, 11, 1779 CrossRef PubMed.
  703. W. Sun, H. Chu, B. Dong, D. Cao and S. Zheng, Int. J. Electrochem. Sci., 2014, 9, 4566–4573 CrossRef.
  704. C. T. Mehmood, Z. Zhong, H. Zhou, C. Zhang and Y. Xiao, RSC Adv., 2020, 10, 36349–36362 RSC.
  705. Y. Cui, Q. Ma, X. Deng, Q. Meng, X. Cheng, M. Xie, X. Li, Q. Cheng and H. Liu, Appl. Catal., B, 2017, 206, 136–145 CrossRef CAS.
  706. S. Silvestri, C. D. Ferreira, V. Oliveira, J. M. T. B. Varejão, J. A. Labrincha and D. M. Tobaldi, J. Photochem. Photobiol., A, 2019, 375, 261–269 CrossRef CAS.
  707. L. Das, S. K. Barodia, S. Sengupta and J. K. Basu, Int. J. Environ. Sci. Technol., 2015, 12, 317–326 CrossRef CAS.
  708. E. Mugunthan, M. B. Saidutta and P. E. Jagadeeshbabu, Environ. Technol., 2019, 40, 929–941 CrossRef CAS PubMed.
  709. E. I. Moreno-Valencia, S. P. Paredes-Carrera, J. C. Sánchez-Ochoa, S. O. Flores-Valle and J. R. Avendaño-Gómez, Mater. Res. Express, 2017, 4, 115026 CrossRef.
  710. S. Li, Z. Wang, X. Zhang, J. Zhao, Z. Hu, Z. Wang and X. Xie, Chem. Eng. J., 2019, 378, 122169 CrossRef CAS.
  711. N. Ahmadpour, M. H. Sayadi, S. Sobhani and M. Hajiani, J. Environ. Manage., 2020, 271, 110964 CrossRef CAS PubMed.
  712. K. Fischer, M. Kühnert, R. Gläser and A. Schulze, RSC Adv., 2015, 5, 16340–16348 RSC.
  713. J. E. Casillas, J. Campa-Molina, F. Tzompantzi, G. G. C. Arízaga, A. López-Gaona, S. Ulloa-Godínez, M. E. Cano and A. Barrera, Materials, 2020, 13, 1345 CrossRef CAS PubMed.
  714. S. Salaeh, D. J. Perisic, M. Biosic, H. Kusic, S. Babic, U. Lavrencic Stangar, D. D. Dionysiou and A. L. Bozic, Chem. Eng. J., 2016, 304, 289–302 CrossRef CAS.
  715. W. Liu, Y. Li, F. Liu, W. Jiang, D. Zhang and J. Liang, Water Res., 2019, 151, 8–19 CrossRef CAS PubMed.
  716. X. Liu, F. Li, Y. Liu, P. Li, L. Chen, B. Li, T. Qian and W. Liu, J. Environ. Chem. Eng., 2022, 10, 107545 CrossRef CAS.
  717. Z. Xiu, H. Bo, Y. Wu and X. Hao, Appl. Surf. Sci., 2014, 289, 394–399 CrossRef CAS.
  718. M. Penas-Garzon, W. H. M. Abdelraheem, C. Belver, J. J. Rodriguez, J. Bedia and D. D. Dionysiou, Sep. Purif. Technol., 2021, 275, 119169 CrossRef CAS.
  719. Z. Hu, X. Cai, Z. Wang, S. Li, Z. Wang and X. Xie, J. Hazard. Mater., 2019, 380, 120812 CrossRef CAS PubMed.
  720. W. Zhang, L. Zhou, J. Shi and H. Deng, J. Colloid Interface Sci., 2017, 496, 167–176 CrossRef CAS PubMed.
  721. W. Zhang, L. Zhou and H. Deng, J. Mol. Catal. A: Chem., 2016, 423, 270–276 CrossRef CAS.
  722. E. Valadez-Renteria, R. Perez-Gonzalez, C. Gomez-Solis, L. A. Diaz-Torres, A. Encinas, J. Oliva and V. Rodriguez-Gonzalez, J. Environ. Sci., 2023, 126, 575–589 CrossRef CAS PubMed.
  723. V. Muelas-Ramos, M. J. Sampaio, C. G. Silva, J. Bedia, J. J. Rodriguez, J. L. Faria and C. Belver, J. Hazard. Mater., 2021, 416, 126199 CrossRef CAS PubMed.
  724. O. Moradi, H. Alizadeh and S. Sedaghat, Chemosphere, 2022, 299, 134435 CrossRef CAS PubMed.
  725. M. Kovacic, K. Perovic, J. Papac, A. Tomic, L. Matoh, B. [Z with combining breve]ener, T. Brodar, I. Capan, A. K. Surca, H. Kuši, U. L. Štangar and A. L. Božic, Materials, 2020, 13, 1621 CrossRef CAS PubMed.
  726. D. John, J. Jose, S. G. Bhat and V. S. Achari, Heliyon, 2021, 7, e07451 CrossRef CAS PubMed.
  727. J. Rashid, S. Oooim, R. Kumar, M. A. Barakat, B. Akram, N. Hussain, H. B. Bin and M. Xu, Sci. Rep., 2020, 10, 14191 CrossRef CAS PubMed.
  728. W. Li, R. Yu, M. Li, N. Guo, H. Yu and Y. Yu, Chemosphere, 2019, 218, 966–973 CrossRef CAS PubMed.
  729. A. Esmaeili and M. H. Entezari, RSC Adv., 2015, 5, 97027 RSC.
  730. O. C. Olatunde and D. C. Onwudiwe, Results Chem., 2022, 4, 100273 CrossRef CAS.
  731. M. E. Malefane, U. Feleni and A. T. Kuvarega, J. Environ. Chem. Eng., 2020, 8, 103560 CrossRef CAS.
  732. A. Kumar, S. K. Sharma, A. Kumar, G. Sharma, N. Almasoud, T. S. Alomar, M. Naushad, Z. A. Alothman and F. J. Stadler, J. Cleaner Prod., 2021, 315, 128137 CrossRef CAS.
  733. L. Zhang, L. Liu, X. Zhang, C. Ge and J. Liao, Adv. Energy Sustainability Res., 2022, 3, 2200138 CrossRef CAS.
  734. D. Kulhary and S. Singh, ChemistrySelect, 2022, 7, e202201964 CrossRef CAS.
  735. M. Irandost, R. Akbarzadeh, M. Pirsaheb, A. Asadi, P. Mohammadi and M. Sillanpää, J. Mol. Liq., 2019, 291, 111342 CrossRef CAS.
  736. M. D. Nguyen, T. B. Nguyen, L. H. Tran, T. G. Nguyen, I. Fatimah, E. P. Kuncoro and R.-A. Doong, Chem. Eng. J., 2023, 452, 139249 CrossRef.
  737. N. Aghababaei, M. Abdouss, H. Hosseini-Monfared and F. Ghanbari, Journal of Water Process Engineering, 2023, 53, 103702 CrossRef.
  738. O. Sacco, J. J. Murcia, A. E. Lara, M. Hernández-Laverde, H. Rojas, J. A. Navío, M. C. Hidalgo and V. Vaiano, Mater. Sci. Semicond. Process., 2020, 107, 104839 CrossRef CAS.
  739. S. Wu, X. Yu, J. Zhang, Y. Zhang, Y. Zhu and M. Zhu, Chem. Eng. J., 2021, 411, 128555 CrossRef CAS.
  740. X. Chen, C. Yu, R. Zhu, N. Li, J. Chen, S. Li, W. Xia, S. Xu, H. Wang and X. Chen, Nanomaterials, 2019, 9, 959 CrossRef CAS PubMed.
  741. X. Chen, C. Yu, R. Zhu, N. Li, J. Chen, Q. Lin, S. Xu, X. Chen and H. Wang, Sci. Total Environ., 2020, 711, 134643 CrossRef CAS PubMed.
  742. P. John, K. Johari, G. S. Nirmala, A. Appusamy and M. Thanabalan, Environ. Technol. Innovation, 2020, 22, 101412 CrossRef.
  743. M. Elangovan, S. M. Bharathaiyengar and J. P. Ettiyappan, Environ. Sci. Pollut. Res., 2021, 28, 18186–18200 CrossRef CAS PubMed.
  744. N. Li, W. Fu, G. Sun, M. Shi, M. Wu, W. Shen, Q. Li and J. Ma, Appl. Organomet. Chem., 2023, 37, e7170 CrossRef CAS.
  745. W. Zhang, L. Zhou, J. Shi and H. Deng, Catalysts, 2018, 8, 45 CrossRef.
  746. A. N. Oliveros, J. A. I. Pimentel, M. D. G. de Luna, S. Garcia-Segura, R. R. M. Abarca and R.-A. Doong, Chem. Eng. J., 2021, 403, 126213 CrossRef CAS.
  747. M. Cai, R. Li, Z. Xie, J. Huang, Y. Zeng, Q. Zhang, H. Liu, W. Lv and G. Liu, Appl. Catal., B, 2019, 259, 118033 CrossRef CAS.
  748. G. A. Ashraf, R. T. Rasool, R. U. Rasool, M. F. Saleem, J. Ali, D. Ghernaout, M. Hassan, A. M. Aljuwayid, M. A. Habila and H. Guo, J. Water Process Eng., 2023, 51, 103435 CrossRef.
  749. M. V. Krishna, G. Madhavi, N. F. Idris, S. A. M. Idris and L. R. K. Chowdary, Arabian J. Chem., 2019, 12, 1290–1297 CrossRef.
  750. F. Pomati, S. Castiglioni, E. Zuccato, R. Fanelli, D. Vigetti, C. Rossetti and D. Calamari, Environ. Sci. Technol., 2006, 40, 2442 CrossRef CAS PubMed.
  751. Z. Bensaadi, N. Yeddou-Mezenner, M. Trari and F. Medjene, J. Environ. Chem. Eng., 2014, 2, 1371–1377 CrossRef CAS.
  752. H. Yang, T. An, G. Y. Li, W. Song, W. J. Cooper, H. Luo and X. Guo, J. Hazard. Mater., 2010, 179, 834–839 CrossRef CAS PubMed.
  753. E. Hapeshi, A. Achilleos, M. I. Vasquez, C. Michael, N. P. Xekoukoulotakis, D. Mantzavinos and D. Kassinos, Water Res., 2010, 44, 1737–1746 CrossRef CAS PubMed.
  754. Z. Ran, L. Wang, Y. Fang, C. Ma and S. Li, Catalysts, 2019, 9, 876 CrossRef CAS.
  755. L. A. Ioannou, E. Hapeshi, M. I. Vasquez, D. Mantzavinos and D. Fatta-Kassinos, Sol. Energy, 2011, 85, 1915–1926 CrossRef CAS.
  756. D. A. Pino-Sandoval, L. Hinojosa-Reyes, J. L. Guzmán-Mar, J. C. Murillo-Sierra and A. Hernandez-Ramirez, Water, Air, Soil Pollut., 2022, 233, 17,  DOI:10.1007/s11270-021-05484-7.
  757. L. Rimoldi, D. Meroni, E. Falletta, V. Pifferi, L. Falciola, G. Cappelletti and S. Ardizzone, Photochem. Photobiol. Sci., 2017, 16, 60 CrossRef CAS PubMed.
  758. A. Ponkshe and P. Thakur, Mater. Today: Proc., 2019, 18, 1162–1175 CAS.
  759. V. Rogé, C. Guignard, G. Lamblin, F. Laporte, I. Fechete, F. Garin, A. Dinia and D. Lenoble, Catal. Today, 2018, 306, 215–222 CrossRef.
  760. Y. Ji, L. Zhou, C. Ferronato, X. Yang, A. Salvador, C. Zeng and J.-M. Chovelon, J. Photochem. Photobiol., A, 2013, 254, 35–44 CrossRef CAS.
  761. C. Medana, P. Calza, F. Carbone, E. Pelizzetti, H. Hidaka and C. Baiocchi, Rapid Commun. Mass Spectrom., 2009, 23, 206 CrossRef CAS.
  762. Z. Mehrabadi and H. Faghihian, J. Photochem. Photobiol., A, 2018, 356, 102–111 CrossRef CAS.
  763. M. Eskandari, N. Goudarzi, M. A. Chamjangali and S. G. Moussavi, Journal of Sabzevar University of Medical Science, 2020, 27, 131–141 Search PubMed.
  764. B. Ramasamy, J. Jeyanthi and P. Chinnaiyan, Environ. Nanotechnol., Monit. Manage., 2023, 19, 100779 CAS.
  765. R. Bhuvaneswari, J. Jeyanthi and S. M. Kumar, Optik, 2021, 239, 166658 CrossRef.
  766. Y. Ling, G. Liao, Y. Xie, J. Yin, J. Huang, W. Feng and L. Li, J. Photochem. Photobiol., A, 2016, 329, 280–286 CrossRef CAS.
  767. Z. A. Alothman, A. Y. Badjah, O. M. Alharbi and I. Ali, Environ. Sci. Pollut. Res., 2021, 28, 7423–7430 CrossRef CAS PubMed.
  768. B. Bina, A. Fatehizadeh, E. Taheri, M. Heydari, M. Darvishmotevalli and A. Bazmeh, Int. J. Environ. Anal. Chem., 2022, 1–12,  DOI:10.1080/03067319.2022.2085045.
  769. S. Stojanović, M. Vranješ, Z. Šaponjić, V. Rac, V. Rakić, L. Ignjatović and L. Damjanović-Vasilić, Int. J. Environ. Sci. Technol., 2023, 20, 1–16 CrossRef.
  770. R. Corchero, R. Rodil, A. Soto and E. Rodil, Nanomaterials, 2021, 11, 411 CrossRef CAS PubMed.
  771. Y. Shi, H. Chen, Y. Wu and W. Dong, Environ. Sci. Pollut. Res., 2017, 25, 693–703 CrossRef PubMed.
  772. V. Pistková, M. Tasbihi, M. Vávrová and U. L. Stangar, J. Photochem. Photobiol., A, 2015, 305, 19–28 CrossRef.
  773. V. Bhatia, G. Malekshoar, A. Dhir and A. K. Ray, J. Photochem. Photobiol., A, 2017, 332, 182–187 CrossRef CAS.
  774. V. Bhatia, K. Manoli, A. Dhir and A. K. Ray, Can. J. Chem. Eng., 2020, 98, 1767–1775 CrossRef CAS.
  775. V. Bhatia, A. Dhar and A. K. Ray, J. Photochem. Photobiol., A, 2021, 409, 113136 CrossRef CAS.
  776. N. F. F. Moreira, M. J. Sampaio, A. R. Ribeiro, C. G. Silva, J. L. Faria and A. M. T. Silva, Appl. Catal., B, 2019, 248, 184–192 CrossRef CAS.
  777. Y. Zhang, S. Yang, L. Wei, X. Zhao and X. Lu, Environ. Technol., 2024, 45, 972–987 CrossRef CAS PubMed.
  778. A. Kumar, A. Kumar, G. Sharma, M. Naushad, T. Ahamad and F. J. Stadler, Chemosphere, 2018, 209, 457–469 CrossRef CAS PubMed.
  779. Y. Wang, J. Niu, X. Gao and Y. Zhang, Appl. Surf. Sci., 2020, 533, 147458 CrossRef.
  780. M. R. Rajeshwari, A. Syed, A. H. Bahkali, A. M. Elgorban, M. K. Rahiman, R. S. Varma and S. S. Khan, J. Ind. Eng. Chem., 2022, 115, 402–415 CrossRef CAS.
  781. A. H. Shah and M. A. Rather, Adv. Nano Res., 2021, 10, 397–414 Search PubMed.
  782. S. Sun, X. Yu, Q. Yang, Z. Yang and S. Liang, Nanoscale Adv., 2019, 1, 34–63 RSC.
  783. Photocatalytic Functional Materials for Environmental Remediation, ed. A. Pandikumar and K. Jothivenkatachalam, 2019, Wiley Search PubMed.
  784. A. B. D. Nandiyanto, R. Zaen and R. Oktiani, Arabian J. Chem., 2020, 13, 1283–1296 CrossRef CAS.
  785. Z. Cui, L. Zhang, Y. Xue, Y. Feng, M. Wang, J. Chen, B. Ji, C. Wang and Y. Xue, Int. J. Miner., Metall. Mater., 2022, 29, 2221–2231 CrossRef CAS.
  786. Z.-H. Diao, J.-C. Jin, M.-Y. Zou, H. Liu, J.-Q. Qin, X.-H. Zhou, W. Qian, P.-R. Guo, L.-J. Kong and W. Chu, Sep. Purif. Technol., 2021, 278, 119620 CrossRef.
  787. L. Han, B. Li, H. Wen, Y. Guo and Z. Lin, J. Mater. Sci. Technol., 2021, 70, 176–184 CrossRef CAS.
  788. E. Omrani, A. Ahmadpour, M. Heravi and T. R. Bastami, J. Water Process Eng., 2022, 47, 102581 CrossRef.
  789. A. Majumder, A. K. Gupta and M. Sillanpää, Colloids Surf., A, 2022, 648, 129250 CrossRef CAS.
  790. S. Das, M. Sanjay, S. Kumar, S. Sarkar, C. S. Tiwary and S. Chowdhury, Chem. Eng. J., 2023, 476, 146719 CrossRef CAS.
  791. T. Zhu, J. Yang, Y. Shen, T. Liang, S. Fan, S. Zhang, Z. Yu, S. Wang and Y. Hou, Colloids Surf., A, 2024, 681, 132802 CrossRef CAS.
  792. M. Wang, B. Liang, S. Lin, Z. Zhao, H. He, X. Li and S. Liang, Appl. Surf. Sci., 2024, 657, 15979 Search PubMed.
  793. C. Yang, X. Zhang, Z. Li, Z. Wang, J. Shi, H. Xie, T. Li and Y. Zhou, Mater. Res. Bull., 2024, 174, 112709 CrossRef CAS.
  794. Y. Xia, H. Liu, F. Sun, B. Yue, X. Wang, F. Guo, Y. Zhu, H. Yu, G. Liu, W. Yu and X. Dong, J. Cleaner Prod., 2024, 434, 140445 CrossRef CAS.
  795. J. J. Rueda-Marquez, I. Levchuk, P. F. Ibañez and M. Sillanpää, J. Cleaner Prod., 2020, 258, 120694 CrossRef CAS.
  796. C. Blaise, F. Gagné, J. F. Férard and P. Eullaffroy, Environ. Toxicol., 2008, 23, 591–598 CrossRef CAS PubMed.
  797. N. B. Turan, H. S. Erkan, G. O. Engin and M. S. Bilgili, Process Saf. Environ. Prot., 2019, 130, 238–249 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3lf00142c

This journal is © The Royal Society of Chemistry 2024