Next Article in Journal
The Role of HDACs in the Response of Cancer Cells to Cellular Stress and the Potential for Therapeutic Intervention
Next Article in Special Issue
Solid Fe Resources Separated from Rolling Oil Sludge for CO Oxidation
Previous Article in Journal
Interplay of Environmental, Individual and Genetic Factors in Rheumatoid Arthritis Provocation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Photocatalytic Degradation of Some Typical Antibiotics: Recent Advances and Future Outlooks

1
Division of Pharmacy and Optometry, Faculty of Biology, Medicine and Health, School of Health Science, University of Manchester, Oxford Road, Manchester M13 9PT, UK
2
Faculty of Biology, Medicine and Health, School of Health Science, University of Manchester, Oxford Road, Manchester M13 9PT, UK
3
State Key Laboratory of Biochemical Engineering, Institute of Process Engineering, Chinese Academy of Sciences, Beijing 100190, China
4
Helmholtz-Zentrum Berlin für Materialien und Energie GmbH, Albert-Einstein-Straße 15, 12489 Berlin, Germany
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(15), 8130; https://doi.org/10.3390/ijms23158130
Submission received: 30 June 2022 / Revised: 19 July 2022 / Accepted: 22 July 2022 / Published: 24 July 2022

Abstract

:
The existence of antibiotics in the environment can trigger a number of issues by fostering the widespread development of antimicrobial resistance. Currently, the most popular techniques for removing antibiotic pollutants from water include physical adsorption, flocculation, and chemical oxidation, however, these processes usually leave a significant quantity of chemical reagents and polymer electrolytes in the water, which can lead to difficulty post-treating unmanageable deposits. Furthermore, though cost-effectiveness, efficiency, reaction conditions, and nontoxicity during the degradation of antibiotics are hurdles to overcome, a variety of photocatalysts can be used to degrade pollutant residuals, allowing for a number of potential solutions to these issues. Thus, the urgent need for effective and rapid processes for photocatalytic degradation leads to an increased interest in finding more sustainable catalysts for antibiotic degradation. In this review, we provide an overview of the removal of pharmaceutical antibiotics through photocatalysis, and detail recent progress using different nanostructure-based photocatalysts. We also review the possible sources of antibiotic pollutants released through the ecological chain and the consequences and damages caused by antibiotics in wastewater on the environment and human health. The fundamental dynamic processes of nanomaterials and the degradation mechanisms of antibiotics are then discussed, and recent studies regarding different photocatalytic materials for the degradation of some typical and commonly used antibiotics are comprehensively summarized. Finally, major challenges and future opportunities for the photocatalytic degradation of commonly used antibiotics are highlighted.

Graphical Abstract

1. Introduction

Antibiotics are chemotherapeutic agents that cure bacterial infections [1]. Currently, antibiotics in the environment are attracting increased attention, prompting a widespread search for possible methods of containment [1,2]. This issue results in the generation of antibiotic-resistant genes and antibiotic-resistant bacteria, which expedite the spread of antibiotic resistance, creating a threat to human health and ecological systems [2,3,4]. Thus, in the 21st century, the threat to the integrity of our water resources from antibiotic pollutants is deemed to be one of the most serious environmental problems worldwide, not only because of environmental damage, but due to the potential harm to human health [5].
During the past decade, many strategies have been adopted to address the problem of wastewater antibiotics [6]. Wastewater treatment is usually considered the main method for managing these antibiotics, since wastewater collects discharge from hospitals, industry, and agriculture [7]. However, many more studies have confirmed that conventional treatments are not highly capable of removing these pollutant compounds, which are predominantly water-soluble, and are neither volatile nor biodegradable [8]. Biotic elimination and non-biotic processes, including sorption, hydrolysis, biodegradation by bacteria, and oxidation, as well as reduction, have attracted a great deal of attention [9]. Yang et al. studied the adsorption, desorption, and biodegradation performance of sulfonamide antibiotics in the existence of activated sludge with and without NaN3 biocide [10]. The experimental results showed that the antibiotics were eliminated by sorption and biodegradation via the activated sludge. Liu et al. investigated four antibiotics including norfloxacin, ofloxacin, roxithromycin, and azithromycin as target antibiotics, and adopted UV254 photolysis, ozonation, and UV/O3 approaches to conduct disposal treatments of nanofiltration, realizing the highest efficiency (>87%) in eliminating antibiotics [11]. Nevertheless, the application of these methods was highly restricted due to the high cost, low stability, and poor recycling ability. Therefore, scientists have been seeking novel methods for degrading antibiotics in wastewater, making the exploration of high-efficiency degradation techniques a popular pursuit for environmental and chemistry researchers [4].
As one of the most promising strategies for degrading antibiotic pollutants, photocatalysis has received much attention due to its low cost, efficiency, and environmental friendliness while degrading antibiotics under sunlight and ambient conditions [12,13]. Most antibiotics are resistant to decomposition owing to their robust molecular structures, thus, the development, design, and fabrication of appropriate photocatalysts with high photocatalytic activities are urgently needed [14]. Though a few catalytic processes have been discussed in the literature so far, there have not been enough examples focusing on the use of appropriate photocatalysts that possess longer wavelength absorption for the photocatalytic degradation of antibiotics, which would help to inform readers about this research field.
The photodegradation of antibiotic pollutants has been reviewed recently [5,15]. However, knowledge of the critical degradation mechanisms and underlying reaction pathways of some typical photodegradation reaction catalysts for antibiotics requires deeper discussion. Furthermore, a comprehensive overview on the possible sources and dangers of the antibiotic pollutants released through the ecological chain, particularly regarding the consequences and damages caused by antibiotic residuals on the environment and human health, is still missing. Additionally, the overall introduction of some commonly-used photocatalytic nanomaterials and their application in the degradation of some typical antibiotics is essential to confirm their practical superiority and effectiveness as photodegradation catalysts.
This review firstly summarizes the effects of antibiotics on living organisms and the environment as well as the basic mechanism of the photocatalytic degradation of antibiotics. Then, commonly used photocatalytic materials for antibiotic degradation are reviewed. Finally, the recent advances in the use of various photocatalytic materials for the degradation of antibiotics are discussed.

2. Consequences of Antibiotics in Wastewater on the Environment and Human Health

Pharmaceuticals can largely improve humans’ health and quality of life when used to treat contagious diseases, however, the misuse of drugs, especially antibiotics, has severe damage on the environment and human health [6,16]. Some results reported remarkable changes in sex ratio and fecundity of daphnia manga when exposed to antibiotics such as sulfamethoxazole and trimethoprim [17]. Meanwhile, a decrease in desire and sexual motivation was observed in experiments on male rats given cimetidine [18].
In some countries, antibiotics are not only employed for animal treatment but also to accelerate animal growth and increase production. Thus, antibiotics might be released from animal waste due to incomplete digestion, and that waste may then be used as fertilizer in agriculture or dumped into wastewater, generating a possible pathway to human harm from food or drink exposure, as shown in Figure 1 [19]. A recent study reported that chlortetracycline antibiotic was, to some extent, uptaken by onions, cabbage, and corn [20]. However, those vegetables did not uptake tyrosine antibiotic, probably due to its large molecular size. Thus, continuous release of these antibiotic pollutants into water environments and organisms has a severely negative impact on the environment by causing genetic exchange and activating drug-resistant bacteria. In particular, most antibiotic pollutants, even under low concentrations, may result in a severe risk to the ecosystem and human health [21,22].
On one hand, in terms of micro-organisms, the release of antibiotics into the environment could lead to chromosomal mutations of native bacteria, triggering the development of antibiotic-resistant bacterial strains, which may cause environmental threats such as toxicological effects on non-intended pathogens, alteration of structures, and dissemination of algal communities [23,24,25]. On the other hand, consumption of water or agricultural and sideline products containing antibiotic pollutants may induce symptoms in humans including, but not limited to, vomiting, tremors, nausea, headache, diarrhea, and nervousness [26]. Furthermore, problems such as restraining spinach growth, physiological teratogenesis, and human gene toxicity have also been reported due to the presence of antibiotic pollutants in water or food [3,27].

3. Principle and Fundamental Mechanism of Photocatalytic Degradation of Antibiotics

The steps involved in the photocatalytic degradation of antibiotics are demonstrated in Figure 2. The predominant mechanisms for antibiotic photocatalytic degradation can be summarized as three main steps: photon absorption, excitation, and reaction [5,7]. In detail, once a photocatalyst absorbs photons with an energy higher than its band gap, the electrons in the valance band (VB) can be excited and jump up into the conduction band (CB), where a hole (h+VB) is produced (Equation (1)) [6,26,28,29]. Subsequently, the photogenerated electrons and holes are efficiently separated and migrate to the surface of the photocatalyst, triggering secondary reactions with the adsorbed materials. Typically, photogenerated holes can also attack those antibiotics directly (Equation (2)), theoretically leading to significant degradation of those toxic antibiotics. In addition, two types of systematic theories about degradation pathways were proposed and recognized by researchers in this field [26]. One is a reductive pathway that happens if the CB potential of the semiconductor is negative compared to that of the O2/•O2 redox potential (−0.13 eV vs. reference hydrogen electrode (RHE)), wherein the photoexcited electrons can react with electron acceptors such as O2 deposits on the catalyst surface or dissolved in water, thereby reducing it to form superoxide radical anion •O2 (Equation (3)) [6,26]. In contrast, another pathway referring to the oxidative pathway was initiated when the holes migrated to the photocatalyst surface, accompanied by hydroxyl radical (•OH) generation upon the oxidation of H2O/OH depending on the alkalinity or acidity of the media (Equation (4)) [6,26]. After being excited, hydrogen ions could recombine with the electrons and generate heat energy (Equation (5)), which would decrease the efficiency of the photodegradation. It is noted that the standard redox potential of photocatalysts should be higher than that of •OH/OH (+1.99 eV vs. RHE) in this case [6,26]. Then, both of these reactive radicals (•OH and •O2) are highly active oxidizing agents in the photocatalytic process [30]. They can effectively mineralize any antibiotics and their intermediates to form water and carbon dioxide under prolonged exposure to high-energy UV irradiation, and eventually decompose into CO2 and H2O (Equation (6)) [28,29,31,32]. Many studies demonstrated that both pathways (reductive and oxidative) should synergistically occur to largely prevent the accumulation of electrons in the CB and significantly decrease the possibility of the recombination of electrons and positive holes compared to the pathway of direct interaction between photogenerated holes and antibiotics [33].
Photocatalyst + hυ → photocatalyst + h+ + e
h+ + antibiotics → H2O + CO2 + degradation products
O2 + e → •O2
H2O/OH + h+ → •OH + H+
H+ + e → energy
Antibiotics + •OH or •O2 → CO2 + H2O + degradation products
Considering the prediction of application and efficiency of a type of photocatalytic material, optical bandgap (Eg) is a very important factor for evaluating photoabsorption ability and photocatalytic efficiency. Mehrorang et al. put forward a method and criterion for bandgap measurement, and divided the concept of the bandgap into the two categories of photonic and electrochemical bandgap facing polyfluorene co-polymers as photocatalysts [34,35]. In addition, They concluded that, the prevention of charge recombination would accordingly lead to a higher lifetime of the active holes, thereby upgrading their antibiotic degradation activity. This proved to be a great strategy for enhancing the activity of photocatalysts under visible light, relating to the interfacial charge transfer from a separate energy surface to a molecular continuous surface from solids [34,35].
All in all, the mechanism of photocatalysis for the degradation of antibiotics can be divided into five main steps: (1) transfer of antibiotics in the fluid phase to the surface; (2) adsorption of the antibiotics; (3) reaction in the adsorbed phase; (4) desorption of the products; and (5) removal of products from the interface region [36,37]. However, photocatalytic degradation suffers the problem of electron-hole recombination in the photocatalyst when the electrons that had been excited to CB rapidly recombine with the separated holes in the VB before creating free radicals [37]. Although this depends on many flexible options such as tuned experimental conditions, the adoption of specific photocatalysts with a low CB–VB bandgap energy and photocatalyst modifications are considered as solutions for these challenges [38,39].

4. Common Photocatalytic Materials for Antibiotic Degradation

4.1. Semiconducting Metal Oxides-Based Photocatalysts

Metal oxide semiconductors have been utilized as pristine photocatalysts or as hybrids, or have been coupled/doped with other materials to facilitate the degradation of organic pollutants such as pesticides, dyes, and polycyclic aromatic hydrocarbons [40]. More importantly, the application of metal oxide-based photocatalysts for antibiotic degradation has recently drawn more interest and attention from researchers due to their good light absorption under UV, visible light, or both, combined with their biocompatibility, safety, and stability when exposed to different conditions [3,41,42]. Generally, metal oxides encounter some challenges regarding ineffectiveness or non-absorbance of photocatalytic activity because of their wide band gap (Figure 3) and faster electron–hole pair recombination [43]. For example, TiO2 is the most popular metal oxide for photocatalysis because of its good optical and electronic properties, chemical stability and reusability, non-toxicity, and low cost [44]. Additionally, ZnO is another semiconducting material that has a better quantum efficiency and higher photocatalytic efficiency compared to TiO2, particularly if used for photocatalytic antibiotic degradation at a neutral pH, however, the high recombination rate of the photogenerated electron–hole pairs limits the utilization of ZnO without any functionalization [45]. Several studies demonstrated that doping with metals like Ag and Fe, or non-metals like N and C, into ZnO enhanced the activity of photocatalytic antibiotic degradation [46,47]. WO3 is another promising metal oxide that has received remarkable attention due to its abundance, cost-effectiveness, and non-toxicity [48,49]. Furthermore, W18O49 was also considered a superior photocatalyst with a higher photocatalytic degradation efficiency compared to WO3 [3,50]. Nevertheless, it is prone to oxidization to WO3 in spite of its superior photocatalytic performance. Thus, the construction of a hybrid of W18O49 and other metal oxides can overcome this oxidization barrier [3,51]. There are many other metal oxides that play important roles in photocatalytic materials for antibiotic degradation, and we may introduce more in the next section.

4.2. Bismuth-Based Photocatalysts

Bismuth, possessing an atomic electron configuration of 6s26p3, is a metallic element from the fifth group of the sixth period in the periodic table, and is usually present in the form of Bi3+ [26,52]. A class of novel processes along with bismuth-based catalysts have been developed for antibiotic degradation, as shown in Figure 4 [26]. Bi oxides display a narrower bandgap due to the overlap of O 2p and Bi 6s orbitals in the valence band and lone-pair distortion of the Bi 6s orbital, resulting in the mobility of photoexcited charges, enhancing the visible light response performance [53]. Interestingly, the Bi5+ valence state from the oxidation of Bi3+ has good absorption of visible light once the 6s orbital is empty [54]. Basically, Bi-based photocatalysts mostly have a bandgap of less than 3 eV. There are some typical Bi-based photocatalysts attracting more attention recently, such as Bi2O3 and BiVO4 [55]. Bi2O3 is one of the most common photocatalysts, showing excellent photocatalytic performance on water-splitting and water treatment from organic wastes [56]. Bi2O3 has a bandgap ranging from 2.1 to 2.8 eV, making its utilization for visible light absorption more efficient. Bi2O3 has five different configurations: α, β, γ, δ, and ω-Bi2O3 [26]. Additionally, BiVO4, with superior physicochemical properties like ferro-elasticity and ionic conductivity, has a theoretical bandgap of 2.047 eV, which maximizes its visible light utilization [57]. BiVO4 was widely used in photocatalytic reactions for organic waste treatment and water splitting in past years [58]. Although it has been confirmed that bismuth-based photocatalysts have good photocatalytic performance and use visible light efficiently, it should be noted that some parameters, such as stability and solubility, need to be emphasized [59].

4.3. Silver-Based Photocatalysts

The application of photocatalytic degradation of silver-based photocatalysts such as AgX (X = Cl, Br, I), Ag2O, Ag3PO4, and Ag2CO3 have been reported by various researchers [60,61,62]. In the case of pristine Ag2CO3, the challenge is that pristine Ag2CO3 is unstable and photocorrosive due to its possible transformation from Ag+ to metallic Ag on account of accepted photoelectrons during the photocatalytic processes [63]. Moreover, pristine Ag2O also exhibits poor stability and rapid electron-hole recombination [64]. Their superior photocatalytic performance on antibiotic degradation depends not only on the reduced electron-hole recombination but may also be from broad and strong absorption ranges in the visible region due to the localized surface plasmon resonance effects induced by Ag nanoparticles [3,7,65].

4.4. Metal-Organic Frameworks (MOFs)-Based Photocatalysts

Metal-organic frameworks (MOFs) are a new class of coordination polymers with periodic network structures formed by the self-assembly between metal ions/metal clusters and organic ligands [66]. By modifying linkers employing functional groups, highly porous structures with remarkable surface areas could be obtained with tuned surface structures [67]. MOFs were first discovered in the mid-1990′s by Omar Yaghi, and the invention of novel MOFs promised long-lived influence in the areas of chemistry, physics, biology, and the material sciences, particularly used extensively in photocatalysis due to their high surface area, adjustable porosity and pore volume [67,68,69]. Thus, MOFs promise to be highly-effective materials for the photocatalytic degradation of antibiotics in a solution [66,70]. Although various MOF-based materials have been utilized to remove antibiotics, the development of more efficient degradation agents remains a key problem facing more active MOF-based photocatalytic degradation materials with more active sites and large surface areas with group functionalization [66,67]. According to the previous report, the organic linker serves as the VB, while the metallic cluster acts as CB. Under exposure to light, MOFs behave like semiconductors, and can thus be deemed as a potential photocatalyst for highly effective degradation of antibiotics due to their superior high thermal and mechanical stability and their excellent structural characteristics [66].

4.5. Graphitic Carbon Nitrides-Based Photocatalysts

Graphitic carbon nitride (g-C3N4), a new class of polymeric semiconducting material, is another kind of promising material for photo-driven catalytic applications [26,71]. On one hand, the past report implied that the g-C3N4 has a bandgap of around 2.7 eV, and the CB–VB can also meet the requirement of overall water splitting, as demonstrated in Figure 5, which made it more popular for photocatalytic water splitting in the past decades [72].
On the other hand, g-C3N4 can also be used for the photocatalytic degradation of antibiotic pollutants under visible light. However, pure g-C3N4 exhibited a low degradation rate due to negative position, resulting in a weak oxidation ability [26,73,74]. Therefore, surface modification is necessary to overcome these limitations [26]. Researchers found that doping with noble metal ions optimized the photocatalytic performance according to several reports in the literature as a result of the higher separation of the photoproduced electrons and holes due to the excellent capacity of electron capture by the noble metallic ions [26,75]. A great deal of research is ongoing regarding g-C3N4 modification for fabricating and designing nanomaterials with different properties in order to obtain the best possible photocatalytic performance for the removal of antibiotics [76].

5. Recent Advances in Photocatalytic Degradation of Antibiotics

5.1. Photocatalytic Degradation of Ciprofloxacin

Ciprofloxacin is a second-generation fluoroquinolone antibiotic used to kill bacteria to prevent severe infection [77]. The chemical structure of ciprofloxacin is exhibited in Figure 6a [1]. Notably, ciprofloxacin dominates 73% of the total consumption, with a daily dose between 0.39 and 1.8 per 1000 patients, and has a broad antimicrobial spectrum that has impact on the DNA gyrase and topoisomerase IV of various Gram-positive and Gram-negative bacteria, thus preventing cell replication [78,79]. It has been considered a good therapy for treating digestive infections, complicated urinary tract infections, sexually transmitted diseases, pulmonary diseases, and skin infections. However, the presence of ciprofloxacin limits photosynthetic pathways and even leads to morphological deformities in higher plants. It also leads to severe damage to human health [78]. In the past, various studies reported the use of modified photocatalysts to meet the demand for higher photocatalytic efficiency in the degradation of ciprofloxacin [80].
Yu et al. [80] prepared Zn-doped Cu2O particles by a solvothermal method to achieve photocatalytic degradation of ciprofloxacin. The photocatalytic results demonstrated that Zn-doped Cu2O has better photocatalytic performance and reusability compared to the undoped Cu2O. 94.6% of ciprofloxacin was degraded in presence of Zn-doped Cu2O, even after 5 cycles, the degradation percentage still remains 91% due to the significantly enhanced absorption intensity in the visible light range, and the increased band gap than that of the undoped Cu2O (Figure 7a,b). In addition, a novel Z-scheme CeO2–Ag/AgBr photocatalyst was fabricated by Malakootian et al. [77] using in situ interspersals of AgBr on CeO2 for subsequent photoreduction process. The results also exhibited largely enhanced photocatalytic activity for the photodegradation of ciprofloxacin under visible light irradiation due to the faster interfacial charge transfer process and the largely enhanced separation of the photogenerated electron-hole pairs. Furthermore, Pattnaik et al. [81] adopted exfoliated graphitic carbon nitride into photocatalytic degradation of ciprofloxacin under solar irradiation and catalytic data have shown that photocatalytic activities of g-C3N4 have enhanced after its exfoliation because of its efficient charge separation, low recombination of photogenerated charge carriers and high surface area. They found that 1 g/L exfoliated nano g-C3N4 can degrade up to 78% of a 20 ppm solution exposed to solar light for lasted 1 h (Figure 7c,d).

5.2. Photocatalytic Degradation of Tetracycline

Tetracyclines are a series of broad-spectrum antibiotics that were first adopted in 1940, and their structures are shown in Figure 6b [1,82]. All tetracyclines have anti-inflammatory and immunosuppressive effects, and were previously used to treat rheumatism [83]. Due to the additional effect of tetracyclines against lipases and collagenases, these antibiotics were also initially used for the intrapleural treatment of malignant effusions [84,85]. Although tetracycline plays a significant role in medicine, the existence of tetracycline in aquatic media is of great concern because of its ecological impact, including carcinogenicity and toxicity to the environment [84,85]. A number of studies related to the removal or degradation of tetracycline through the use of different photocatalytic materials have been reported in past years [86]. For example, a novel TiO2/g-C3N4 core-shell quantum heterojunction prepared by a feasible strategy of polymerizing the quantum trick graphitic carbon nitride (g-C3N4) onto the surface of anatase titanium dioxide nanosheets was put forward by Wang et al. to be employed as a tetracycline degradation photocatalyst, and this catalyst exhibited the highest tetracycline degradation rate: 2.2 mg/min, which is 36% higher than that of the TiO2/g-C3N4 mixture, 2 times higher than that of TiO2, and 2.3 times higher than that of bulk g-C3N4 (Figure 8a,b) [87]. Moreover, Wang et al. synthesized a novel C–N–S tri-doped TiO2 using a facile and cost-effective sol–gel method with titanium butoxide as titanium precursor and thiourea as the dopant source, which can be used for photocatalytic degradation of tetracycline under visible light [86]. The catalytic results exhibited the highest photocatalytic degradation efficiency of tetracycline under visible light irradiation which is associated with the synergistic effects of tetracycline adsorption due to its high surface area, narrow band gap causing C–N–S tri-doping, presence of carbonaceous species functioning as a photosensitizer, and well-organized anatase phase. Additionally, Chen et al. synthesized a novel heterostructured photocatalyst AgI/BiVO4 by an in situ precipitation procedure, and the results exhibited excellent photoactivity for tetracycline decomposition under visible light irradiation, the tetracycline molecules were apparently eliminated (94.91%) within 60 min, and degradation efficiency was remarkably superior to those of bare BiVO4 (62.68%) and AgI (75.43%) under same experimental conditions (Figure 8c,d) [88].

5.3. Photocatalytic Degradation of Norfloxacin

Norfloxacin is another antibiotic within the fluro-quinolones group, and is widely used for curing urinary tract infections [89]. Figure 6c indicates the chemical structure of norfloxacin [1]. Currently, the presence of norfloxacin in wastewater (especially in hospitals) contains high concentrations and is deemed to be one of the potential pollutants in the aquatic environment. In the past few years, fluoroquinolone antibiotics have triggered tremendous concern due to their widespread use and environmental toxicity [89,90]. There are several reports on the degradation of norfloxacin by photocatalysis using different materials [91]. Sayed et al. [92] prepared a novel immobilized TiO2/Ti film with exposed {001} facets via a facile one-pot hydrothermal route to use in the degradation of norfloxacin from aqueous media, and the experimental results demonstrated excellent photocatalytic performance toward the degradation of norfloxacin in various water matrices, with the observation that •OH is mainly involved in the photocatalytic degradation of norfloxacin by {001} faceted TiO2/Ti film (Figure 9a). Additionally, Tang et al. [93] realized excellent visible-light-driven photocatalytic performance for the degradation of norfloxacin by an as-prepared novel Z-scheme Ag/FeTiO3/Ag/BiFeO3 using a sol–gel method followed by a photo-reduction process. The results showed the photocatalytic degradation extent reaches 96.5% within 150 min when using Ag/FeTiO3/Ag/BiFeO3 at 2.0 wt.% Ag (FeTiO3: BiFeO3 = 1.0:0.5) which can be reused with excellent photocatalytic stability (Figure 9b–d). Moreover, Lv et al. [94] synthesized copper-doped bismuth oxybromide (Cu-doped BiOBr) using a solvothermal method, and assessed their ability to degrade norfloxacin under visible light. The as-prepared Cu-doped BiOBr showed high activity with a photocatalytic degradation constant of 0.64 ×10−2 min−1 in the photocatalytic degradation of norfloxacin under visible-light irradiation due to its enhanced light-harvesting properties, enhanced charge separation, and interfacial charge transfer, as well as a retention of 95% of its initial activity, even after 5 constant catalytic cycles. Similarly, Bi2WO6, another bismuth-based catalyst, was put forward by Tang et al. [95] and applied to the photodegradation of norfloxacin in a nonionic surfactant Triton-X100 (TX100)/Bi2WO6 dispersion under visible light irradiation. The results found that the degradation of barely insoluble norfloxacin could be strongly enhanced with the addition of TX100. TX100 was adsorbed strongly on the Bi2WO6 surface and promoted norfloxacin photodegradation at the critical micelle concentration (CMC = 0.25 mM).

5.4. Photocatalytic Degradation of Amoxicillin

Amoxicillin is a penicillin-type antibiotic medicine extensively used for the treatment of various bacterial infections such as dental infections, chest infections, and other infections (ear, throat, and sinus). Figure 6d exhibits the chemical structure of amoxicillin [1]. However, amoxicillin in water or in the ecological environment is considered an emerging pollutant because it can cause several health effects to aquatic life in the presence of solved molecules in water [16]. There are some reports regarding the use of different photocatalytic materials to degrade amoxicillin [96]. For instance, the photocatalytic degradation of amoxicillin by as-prepared titanium dioxide nanoparticles loaded on graphene oxide (GO/TiO2) by the chemical hydrothermal method was evaluated under UV light by Balarak et al., and the experimental data exhibited that key indexes such as initial pH, GO/TiO2 dosage, UV intensity, and initial amoxicillin concentration all had a significant impact on amoxicillin degradation (Figure 10a) [97]. The efficiency of amoxicillin degradation collection was measured to be more than 99% at specific conditions, including a pH of 6, a GO/TiO2 dosage of 0.4 g/L, an amoxicillin concentration of 50 mg/L, and an intensity of 36 W. Additionally, Mirzaei et al. [98] synthesized a new fluorinated graphite carbon nitride photocatalyst with magnetic properties by a gentle hydrothermal method that can be used for the degradation of amoxicillin in water. Compared to the bulk g-C3N4, magnetic fluorinated Fe3O4/g-C3N4 with a high specific surface area (243 m2g−1) resulted in improved photocatalytic activity regarding amoxicillin degradation and the mineralization of the solution. Furthermore, Huang et al. [99] also prepared novel carbon-rich g-C3N4 nanosheets with large surface areas by a facile thermal polymerization method, which displayed superior photocatalytic activity for amoxicillin degradation under solar light (Figure 10b–d). Meanwhile, the catalyst showed high stability and amoxicillin degradation ability under various media conditions, indicating its high applicability for amoxicillin treatment.
Finally, we list a table to compare the different photocatalysts discussed above, and note that the information in Table 1 shows most of the catalysts can easily and effectively remove the antibiotic contaminations within 2 h with relatively high degradation efficiency, which fully proves the superiority of rapid and effective antibiotic removal by photodegradation.

6. Conclusions and Future Perspective

In this review, the photocatalytic degradation of antibiotics was summarized. Firstly, the mechanism of photocatalytic degradation of antibiotics depending on the formation of free radicals and active oxygen species, and the consequences of antibiotics in wastewater on the environment and human health were reviewed. Some widely used antibiotics were then analyzed, and a number of commonly used photocatalysts were introduced. Heteroatom doping is generally used as a strategy to enhance the photocatalytic performance of a photocatalyst, particularly metal atoms as dopants. However, it should be noted that metal dopants could serve as recombination centers at higher concentrations, which can reduce the efficiency of a photocatalyst. Consequently, future research should also focus on other options, including doping with non-metals such as nitrogen, boron, sulfur, and phosphorus. Meanwhile, the formation of a heterojunction with other semiconductors can also play a significant role in the modification of photocatalysts on the degradation of antibiotics due to other semiconductors possibly serving as photosensitizers while simultaneously inhibiting electron-hole recombination. Thus, these methods can achieve visible light-driven photocatalysts with enhanced photocatalytic activity by narrowing the band gap of the photocatalyst or by increasing the activity of charge separation.
Physicochemical properties such as morphology and surface areas are also very critical factors in the performance of catalysts during photodegradation studies. As mentioned above, further studying photocatalysts with different morphologies and surface areas can effectively enhance the performance of catalysts. Furthermore, the degradation pathway also provides a clear introduction to the fate and transformation of antibiotics during the photocatalytic degradation process. Thus, exploring the photocatalytic degradation mechanism at the atomic level is also necessary for accelerating the efficiency of antibiotic degradation.
The utilization of solar radiation and visible light sources to activate photocatalysts during the photodegradation of antibiotics such as ciprofloxacin, tetracycline, norfloxacin, and amoxicillin is still limited. Therefore, the exploration and development of photodegradation induced by UV light sources are still in urgent demand.
Firstly, in the long term, although the removal rate of antibiotics is still being optimized, the removal rate of the chemical oxygen demand is still relatively high during the degradation of antibiotics by photocatalysts, and thus it confirms that the mineralization degree of antibiotics needs to be optimal. There are many intermediates during the process of photodegradation, therefore a deep study into intermediates is also critical for improving the performance of catalysts. Secondly, most experiments involve regular and constant stirring to prevent the agglomeration of materials in media during the degradation of antibiotics by photocatalysts, which requires additional energy consumption. Thirdly, the problems of antibiotics are not only induced by water quality but also by the accumulation of antibiotics in water, which lead to the generation of microbial resistance genes. There is currently still a lack of research on photocatalysts’ limited resistance genes. Finally, The recycling ability of a photocatalyst is a significant index for evaluating its cost-effectiveness and feasibility for practical application in the degradation of antibiotics. To minimize any possible waste, the design of photocatalysts with quasi-same photoactivity during each cycle is preferred. Also, it is important to design photocatalysts that are easier to separate and recycle in order not to avoid losing any worthy materials during the photocatalytic reaction. Thus, the separation of photocatalysts from the aqueous phase is crucial from an economic standpoint. It is noted that the operating cost of a photocatalytic reaction mainly originates from its being a single-use photocatalyst, unable to be recycled. Regarding the repetitive usage of photocatalysts, deep research on how to sustainably use recyclable photocatalysts for antibiotic degradation is still urgently needed.

Funding

This work is supported by the National Natural Science Foundation of China (Grant No. 51772296).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No data.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Abdurahman, M.H.; Abdullah, A.Z.; Shoparwe, N.F. A Comprehensive Review on Sonocatalytic, Photocatalytic, and Sonophotocatalytic Processes for the Degradation of Antibiotics in Water: Synergistic Mechanism and Degradation Pathway. Chem. Eng. J. 2021, 413, 127412. [Google Scholar] [CrossRef]
  2. Calvete, M.J.F.; Piccirillo, G.; Vinagreiro, C.S.; Pereira, M.M. Hybrid Materials for Heterogeneous Photocatalytic Degradation of Antibiotics. Coord. Chem. Rev. 2019, 395, 63–85. [Google Scholar] [CrossRef]
  3. Shurbaji, S.; Huong, P.T.; Altahtamouni, T.M. Review on the Visible Light Photocatalysis for the Decomposition of Ciprofloxacin, Norfloxacin, Tetracyclines, and Sulfonamides Antibiotics in Wastewater. Catalysts 2021, 11, 437. [Google Scholar] [CrossRef]
  4. Li, D.; Shi, W. Recent Developments in Visible-Light Photocatalytic Degradation of Antibiotics. Chin. J. Catal. 2016, 37, 792–799. [Google Scholar] [CrossRef]
  5. Wei, Z.; Liu, J.; Shangguan, W. A Review on Photocatalysis in Antibiotic Wastewater: Pollutant Degradation and Hydrogen Production. Chin. J. Catal. 2020, 41, 1440–1450. [Google Scholar] [CrossRef]
  6. Velempini, T.; Prabakaran, E.; Pillay, K. Recent Developments in the Use of Metal Oxides for Photocatalytic Degradation of Pharmaceutical Pollutants in Water—A Review. Mater. Today Chem. 2021, 19, 100380. [Google Scholar] [CrossRef]
  7. Shehu Imam, S.; Adnan, R.; Mohd Kaus, N.H. Photocatalytic Degradation of Ciprofloxacin in Aqueous Media: A Short Review. Toxicol. Environ. Chem. 2018, 100, 518–539. [Google Scholar] [CrossRef]
  8. Cuerda-Correa, E.M.; Alexandre-Franco, M.F.; Fernández-González, C. Advanced Oxidation Processes for the Removal of Antibiotics from Water. An Overview. Water 2020, 12, 102. [Google Scholar] [CrossRef] [Green Version]
  9. Yang, Q.; Gao, Y.; Ke, J.; Show, P.L.; Ge, Y.; Liu, Y.; Guo, R.; Chen, J. Antibiotics: An Overview on the Environmental Occurrence, Toxicity, Degradation, and Removal Methods. Bioengineered 2021, 12, 7376–7416. [Google Scholar] [CrossRef]
  10. Yang, S.-F.; Lin, C.-F.; Yu-Chen Lin, A.; Andy Hong, P.-K. Sorption and Biodegradation of Sulfonamide Antibiotics by Activated Sludge: Experimental Assessment Using Batch Data Obtained Under Aerobic Conditions. Water Res. 2011, 45, 3389–3397. [Google Scholar] [CrossRef]
  11. Liu, P.; Zhang, H.; Feng, Y.; Yang, F.; Zhang, J. Removal of Trace Antibiotics from Wastewater: A Systematic Study of Nanofiltration Combined with Ozone-Based Advanced Oxidation Processes. Chem. Eng. J. 2014, 240, 211–220. [Google Scholar] [CrossRef]
  12. Boxi, S.S.; Paria, S. Visible Light Induced Enhanced Photocatalytic Degradation of Organic Pollutants in Aqueous Media Using Ag Doped Hollow TiO2 Nanospheres. RSC Adv. 2015, 5, 37657–37668. [Google Scholar] [CrossRef]
  13. Kushwaha, H.S.; Halder, A.; Jain, D.; Vaish, R. Visible Light-Induced Photocatalytic and Antibacterial Activity of Li-Doped Bi0.5Na0.45K0.5TiO3–BaTiO3 Ferroelectric Ceramics. J. Electron. Mater. 2015, 44, 4334–4342. [Google Scholar] [CrossRef]
  14. Huo, P.; Lu, Z.; Liu, X.; Wu, D.; Liu, X.; Pan, J.; Gao, X.; Guo, W.; Li, H.; Yan, Y. Preparation Photocatalyst of Selected Photodegradation Antibiotics by Molecular Imprinting Technology onto TiO2/Fly-Ash Cenospheres. Chem. Eng. J. 2012, 189–190, 75–83. [Google Scholar] [CrossRef]
  15. Yang, X.; Chen, Z.; Zhao, W.; Liu, C.; Qian, X.; Zhang, M.; Wei, G.; Khan, E.; Hau Ng, Y.; Sik Ok, Y. Recent Advances in Photodegradation of Antibiotic Residues in Water. Chem. Eng. J. 2021, 405, 126806. [Google Scholar] [CrossRef]
  16. Kraemer, S.A.; Ramachandran, A.; Perron, G.G. Antibiotic Pollution in the Environment: From Microbial Ecology to Public Policy. Microorganisms 2019, 7, 180. [Google Scholar] [CrossRef] [Green Version]
  17. De Liguoro, M.; Fioretto, B.; Poltronieri, C.; Gallina, G. The Toxicity of Sulfamethazine to Daphnia Magna and Its Additivity to Other Veterinary Sulfonamides and Trimethoprim. Chemosphere 2009, 75, 1519–1524. [Google Scholar] [CrossRef]
  18. Bialy, M.; Bogacki-Rychlik, W.; Przybylski, J.; Zera, T. The Sexual Motivation of Male Rats as a Tool in Animal Models of Human Health Disorders. Front. Behav. Neurosci. 2019, 13, 257. [Google Scholar] [CrossRef]
  19. Kar, P.; Shukla, K.; Jain, P.; Sathiyan, G.; Gupta, R.K. Semiconductor Based Photocatalysts for Detoxification of Emerging Pharmaceutical Pollutants from Aquatic Systems: A Critical Review. Nano Mater. Sci. 2021, 3, 25–46. [Google Scholar] [CrossRef]
  20. Manyi-Loh, C.; Mamphweli, S.; Meyer, E.; Okoh, A. Antibiotic Use in Agriculture and Its Consequential Resistance in Environmental Sources: Potential Public Health Implications. Molecules 2018, 23, 795. [Google Scholar] [CrossRef] [Green Version]
  21. Fletcher, S. Understanding the Contribution of Environmental Factors in the Spread of Antimicrobial Resistance. Environ. Health Prev. Med. 2015, 20, 243–252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Larsson, D.G.J.; Flach, C.-F. Antibiotic Resistance in the Environment. Nat. Rev. Microbiol. 2022, 20, 257–269. [Google Scholar] [CrossRef] [PubMed]
  23. Uddin, T.M.; Chakraborty, A.J.; Khusro, A.; Zidan, B.R.M.; Mitra, S.; Emran, T.B.; Dhama, K.; Ripon, K.H.; Gajdács, M.; Sahibzada, M.U.K.; et al. Antibiotic Resistance in Microbes: History, Mechanisms, Therapeutic Strategies and Future Prospects. J. Infect. Public Health 2021, 14, 1750–1766. [Google Scholar] [CrossRef] [PubMed]
  24. Davies Julian; Davies Dorothy Origins and Evolution of Antibiotic Resistance. Microbiol. Mol. Biol. Rev. 2010, 74, 417–433. [CrossRef] [Green Version]
  25. Peterson, E.; Kaur, P. Antibiotic Resistance Mechanisms in Bacteria: Relationships Between Resistance Determinants of Antibiotic Producers, Environmental Bacteria, and Clinical Pathogens. Front. Microbiol. 2018, 9, 2928. [Google Scholar] [CrossRef]
  26. Qin, K.; Zhao, Q.; Yu, H.; Xia, X.; Li, J.; He, S.; Wei, L.; An, T. A Review of Bismuth-Based Photocatalysts for Antibiotic Degradation: Insight into the Photocatalytic Degradation Performance, Pathways and Relevant Mechanisms. Environ. Res. 2021, 199, 111360. [Google Scholar] [CrossRef]
  27. Founou, L.L.; Founou, R.C.; Essack, S.Y. Antibiotic Resistance in the Food Chain: A Developing Country-Perspective. Front. Microbiol. 2016, 7, 1881. [Google Scholar] [CrossRef]
  28. Gunawan, M.A.; Hierso, J.-C.; Poinsot, D.; Fokin, A.A.; Fokina, N.A.; Tkachenko, B.A.; Schreiner, P.R. Diamondoids: Functionalization and Subsequent Applications of Perfectly Defined Molecular Cage Hydrocarbons. New J. Chem. 2014, 38, 28–41. [Google Scholar] [CrossRef]
  29. Li, Y.; Fu, Y.; Zhu, M. Green Synthesis of 3d Tripyramid TiO2 Architectures with Assistance of Aloe Extracts for Highly Efficient Photocatalytic Degradation of Antibiotic Ciprofloxacin. Appl. Catal. B Environ. 2020, 260, 118149. [Google Scholar] [CrossRef]
  30. Soutsas, K.; Karayannis, V.; Poulios, I.; Riga, A.; Ntampegliotis, K.; Spiliotis, X.; Papapolymerou, G. Decolorization and Degradation of Reactive Azo Dyes via Heterogeneous Photocatalytic Processes. Desalination 2010, 250, 345–350. [Google Scholar] [CrossRef]
  31. Chen, X.; Yao, J.; Xia, B.; Gan, J.; Gao, N.; Zhang, Z. Influence of pH and DO on the Ofloxacin Degradation in Water by UVA-LED/TiO2 Nanotube Arrays Photocatalytic Fuel Cell: Mechanism, ROSs Contribution and Power Generation. J. Hazard. Mater. 2020, 383, 121220. [Google Scholar] [CrossRef] [PubMed]
  32. Pinto, M.F.; Olivares, M.; Vivancos, Á.; Guisado-Barrios, G.; Albrecht, M.; Royo, B. (Di)triazolylidene Manganese Complexes in Catalytic Oxidation of Alcohols to Ketones and Aldehydes. Catal. Sci. Technol. 2019, 9, 2421–2425. [Google Scholar] [CrossRef]
  33. Muruganandham, M.; Swaminathan, M. Solar Photocatalytic Degradation of a Reactive Azo Dye in TiO2-Suspension. Sol. Energy Mater. Sol. Cells 2004, 81, 439–457. [Google Scholar] [CrossRef]
  34. Taghizadeh, A.; Taghizadeh, M.; Sabzehmeidani, M.M.; Sadeghfar, F.; Ghaedi, M. Chapter 1—Electronic Structure: From Basic Principles to Photocatalysis. In Interface Science and Technology; Ghaedi, M., Ed.; Elsevier: Amsterdam, The Netherlands, 2021; Volume 32, pp. 1–53. ISBN 1573-4285. [Google Scholar]
  35. Sadeghfar, F.; Zalipour, Z.; Taghizadeh, M.; Taghizadeh, A.; Ghaedi, M. Chapter 2—Photodegradation Processes. In Interface Science and Technology; Ghaedi, M., Ed.; Elsevier: Amsterdam, The Netherlands, 2021; Volume 32, pp. 55–124. ISBN 1573-4285. [Google Scholar]
  36. Sacco, O.; Vaiano, V.; Han, C.; Sannino, D.; Dionysiou, D.D. Photocatalytic Removal of Atrazine Using N-Doped TiO2 Supported on Phosphors. Appl. Catal. B Environ. 2015, 164, 462–474. [Google Scholar] [CrossRef]
  37. Zhao, G.; Ding, J.; Zhou, F.; Chen, X.; Wei, L.; Gao, Q.; Wang, K.; Zhao, Q. Construction of a Visible-Light-Driven Magnetic Dual Z-Scheme BiVO4/g-C3n4/NiFe2O4 Photocatalyst for Effective Removal of Ofloxacin: Mechanisms and Degradation Pathway. Chem. Eng. J. 2021, 405, 126704. [Google Scholar] [CrossRef]
  38. Topkaya, E.; Konyar, M.; Yatmaz, H.C.; Öztürk, K. Pure ZnO and Composite ZnO/TiO2 Catalyst Plates: A Comparative Study for the Degradation of Azo Dye, Pesticide and Antibiotic in Aqueous Solutions. J. Colloid Interface Sci. 2014, 430, 6–11. [Google Scholar] [CrossRef]
  39. Ni, M.; Leung, M.K.H.; Leung, D.Y.C.; Sumathy, K. A Review and Recent Developments in Photocatalytic Water-Splitting Using TiO2 for Hydrogen Production. Renew. Sustain. Energy Rev. 2007, 11, 401–425. [Google Scholar] [CrossRef]
  40. Djurišić, A.B.; He, Y.; Ng, A.M.C. Visible-Light Photocatalysts: Prospects and Challenges. APL Mater. 2020, 8, 030903. [Google Scholar] [CrossRef] [Green Version]
  41. Soni, V.; Khosla, A.; Singh, P.; Nguyen, V.-H.; Le, Q.V.; Selvasembian, R.; Hussain, C.M.; Thakur, S.; Raizada, P. Current Perspective in Metal Oxide Based Photocatalysts for Virus Disinfection: A Review. J. Environ. Manag. 2022, 308, 114617. [Google Scholar] [CrossRef]
  42. Zeng, J.; Li, Z.; Jiang, H.; Wang, X. Progress on Photocatalytic Semiconductor Hybrids for Bacterial Inactivation. Mater. Horiz. 2021, 8, 2964–3008. [Google Scholar] [CrossRef]
  43. Haque, F.; Daeneke, T.; Kalantar-zadeh, K.; Ou, J.Z. Two-Dimensional Transition Metal Oxide and Chalcogenide-Based Photocatalysts. Nano-Micro Lett. 2017, 10, 23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. FUJISHIMA, A.; HONDA, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  45. Chen, X.; He, Y.; Zhang, Q.; Li, L.; Hu, D.; Yin, T. Fabrication of Sandwich-Structured ZnO/Reduced Graphite Oxide Composite and Its Photocatalytic Properties. J. Mater. Sci. 2010, 45, 953–960. [Google Scholar] [CrossRef]
  46. Semeraro, P.; Bettini, S.; Sawalha, S.; Pal, S.; Licciulli, A.; Marzo, F.; Lovergine, N.; Valli, L.; Giancane, G. Photocatalytic Degradation of Tetracycline by ZnO/γ-Fe2O3 Paramagnetic Nanocomposite Material. Nanomaterials 2020, 10, 1458. [Google Scholar] [CrossRef] [PubMed]
  47. Mirzaeifard, Z.; Shariatinia, Z.; Jourshabani, M.; Rezaei Darvishi, S.M. ZnO Photocatalyst Revisited: Effective Photocatalytic Degradation of Emerging Contaminants Using S-Doped ZnO Nanoparticles Under Visible Light Radiation. Ind. Eng. Chem. Res. 2020, 59, 15894–15911. [Google Scholar] [CrossRef]
  48. Gholami, P.; Khataee, A.; Bhatnagar, A. Photocatalytic Degradation of Antibiotic and Hydrogen Production Using Diatom-Templated 3d Wo3-x@mesoporous Carbon Nanohybrid Under Visible Light Irradiation. J. Clean. Prod. 2020, 275, 124157. [Google Scholar] [CrossRef]
  49. Nguyen, T.T.; Nam, S.-N.; Son, J.; Oh, J. Tungsten Trioxide (Wo3)-Assisted Photocatalytic Degradation of Amoxicillin by Simulated Solar Irradiation. Sci. Rep. 2019, 9, 9349. [Google Scholar] [CrossRef]
  50. Wang, J.; Fang, X.-C.; Liu, Y.; Fu, S.-L.; Zhuo, M.-P.; Yao, M.-D.; Wang, Z.-S.; Chen, W.-F.; Liao, L.-S. W18o49/N-Doped Reduced Graphene Oxide Hybrid Architectures for Full-Spectrum Photocatalytic Degradation of Organic Contaminants in Water. J. Mater. Chem. C 2021, 9, 829–835. [Google Scholar] [CrossRef]
  51. Yang, Y.; Qiu, M.; Chen, F.; Qi, Q.; Yan, G.; Liu, L.; Liu, Y. Charge-Transfer-Mediated Photocatalysis of W18o49@CdS Nanotubes to Boost Photocatalytic Hydrogen Production. Appl. Surf. Sci. 2021, 541, 148415. [Google Scholar] [CrossRef]
  52. Gurunathan, K. Photocatalytic Hydrogen Production Using Transition Metal Ions-Doped γ-Bi2O3 Semiconductor Particles. Int. J. Hydrogen Energy 2004, 29, 933–940. [Google Scholar] [CrossRef]
  53. Chen, P.; Zhang, Q.; Zheng, X.; Tan, C.; Zhuo, M.; Chen, T.; Wang, F.; Liu, H.; Liu, Y.; Feng, Y.; et al. Phosphate-Modified m-Bi2O4 Enhances the Absorption and Photocatalytic Activities of Sulfonamide: Mechanism, Reactive Species, and Reactive Sites. J. Hazard. Mater. 2020, 384, 121443. [Google Scholar] [CrossRef] [PubMed]
  54. Tang, J.; Zou, Z.; Ye, J. Efficient Photocatalysis on BaBiO3 Driven by Visible Light. J. Phys. Chem. C 2007, 111, 12779–12785. [Google Scholar] [CrossRef]
  55. He, R.; Cao, S.; Zhou, P.; Yu, J. Recent Advances in Visible Light Bi-Based Photocatalysts. Chin. J. Catal. 2014, 35, 989–1007. [Google Scholar] [CrossRef]
  56. Huang, Y.; Fan, W.; Long, B.; Li, H.; Zhao, F.; Liu, Z.; Tong, Y.; Ji, H. Visible Light Bi2S3/Bi2O3/Bi2O2CO3 Photocatalyst for Effective Degradation of Organic Pollutions. Appl. Catal. B Environ. 2016, 185, 68–76. [Google Scholar] [CrossRef]
  57. Gotić, M.; Musić, S.; Ivanda, M.; Šoufek, M.; Popović, S. Synthesis and Characterisation of Bismuth(III) Vanadate. J. Mol. Struct. 2005, 744–747, 535–540. [Google Scholar] [CrossRef]
  58. Wang, Z.; Huang, X.; Wang, X. Recent Progresses in the Design of BiVO4-Based Photocatalysts for Efficient Solar Water Splitting. Catal. Today 2019, 335, 31–38. [Google Scholar] [CrossRef]
  59. Wang, G.; Sun, Q.; Liu, Y.; Huang, B.; Dai, Y.; Zhang, X.; Qin, X. A Bismuth-Based Metal–Organic Framework as an Efficient Visible-Light-Driven Photocatalyst. Chem. A Eur. J. 2015, 21, 2364–2367. [Google Scholar] [CrossRef]
  60. Liu, W.; Liu, D.; Wang, K.; Yang, X.; Hu, S.; Hu, L. Fabrication of Z-Scheme Ag3PO4/TiO2 Heterostructures for Enhancing Visible Photocatalytic Activity. Nanoscale Res. Lett. 2019, 14, 203. [Google Scholar] [CrossRef] [Green Version]
  61. Wang, H.; Li, J.; Huo, P.; Yan, Y.; Guan, Q. Preparation of Ag2O/Ag2CO3/MWNTs Composite Photocatalysts for Enhancement of Ciprofloxacin Degradation. Appl. Surf. Sci. 2016, 366, 1–8. [Google Scholar] [CrossRef]
  62. Li, W.; Chen, Q.; Lei, X.; Gong, S. Fabrication of Ag/AgBr/Ag3VO4 Composites with High Visible Light Photocatalytic Performance. RSC Adv. 2019, 9, 5100–5109. [Google Scholar] [CrossRef] [Green Version]
  63. Dai, G.; Yu, J.; Liu, G. A New Approach for Photocorrosion Inhibition of Ag2CO3 Photocatalyst with Highly Visible-Light-Responsive Reactivity. J. Phys. Chem. C 2012, 116, 15519–15524. [Google Scholar] [CrossRef]
  64. Tian, L.; Sun, K.; Rui, Y.; Cui, W.; An, W. Facile Synthesis of an Ag@AgBr Nanoparticle-Decorated K4Nb6O17 Photocatalyst with Improved Photocatalytic Properties. RSC Adv. 2018, 8, 29309–29320. [Google Scholar] [CrossRef] [Green Version]
  65. Khanam, S.; Rout, S.K. A Photocatalytic Hydrolysis and Degradation of Toxic Dyes by Using Plasmonic Metal–Semiconductor Heterostructures: A Review. Chemistry 2022, 4, 454–479. [Google Scholar] [CrossRef]
  66. Du, C.; Zhang, Z.; Yu, G.; Wu, H.; Chen, H.; Zhou, L.; Zhang, Y.; Su, Y.; Tan, S.; Yang, L.; et al. A Review of Metal Organic Framework (MOFs)-Based Materials for Antibiotics Removal via Adsorption and Photocatalysis. Chemosphere 2021, 272, 129501. [Google Scholar] [CrossRef] [PubMed]
  67. Zhang, J.; Xiang, S.; Wu, P.; Wang, D.; Lu, S.; Wang, S.; Gong, F.; Wei, X.; Ye, X.; Ding, P. Recent Advances in Performance Improvement of Metal-Organic Frameworks to Remove Antibiotics: Mechanism and Evaluation. Sci. Total Environ. 2022, 811, 152351. [Google Scholar] [CrossRef] [PubMed]
  68. Wang, F.; Xue, R.; Ma, Y.; Ge, Y.; Wang, Z.; Qiao, X.; Zhou, P. Study on the Performance of a MOF-808-Based Photocatalyst Prepared by a Microwave-Assisted Method for the Degradation of Antibiotics. RSC Adv. 2021, 11, 32955–32964. [Google Scholar] [CrossRef]
  69. Gheytanzadeh, M.; Baghban, A.; Habibzadeh, S.; Jabbour, K.; Esmaeili, A.; Mohaddespour, A.; Abida, O. An Insight into Tetracycline Photocatalytic Degradation by MOFs Using the Artificial Intelligence Technique. Sci. Rep. 2022, 12, 6615. [Google Scholar] [CrossRef]
  70. Li, X.; Zhang, D.; Bai, R.; Mo, R.; Yang, C.; Li, C.; Han, Y. Zr-MOFs Based BiOBr/UiO-66 Nanoplates with Enhanced Photocatalytic Activity for Tetracycline Degradation Under Visible Light Irradiation. AIP Adv. 2020, 10, 125228. [Google Scholar] [CrossRef]
  71. Dong, G.; Zhang, Y.; Pan, Q.; Qiu, J. A Fantastic Graphitic Carbon Nitride (g-C3n4) Material: Electronic Structure, Photocatalytic and Photoelectronic Properties. J. Photochem. Photobiol. C Photochem. Rev. 2014, 20, 33–50. [Google Scholar] [CrossRef]
  72. Wang, X.; Maeda, K.; Thomas, A.; Takanabe, K.; Xin, G.; Carlsson, J.M.; Domen, K.; Antonietti, M. A Metal-Free Polymeric Photocatalyst for Hydrogen Production from Water Under Visible Light. Nat. Mater. 2009, 8, 76–80. [Google Scholar] [CrossRef]
  73. Alaghmandfard, A.; Ghandi, K. A Comprehensive Review of Graphitic Carbon Nitride (g-C3n4)–Metal Oxide-Based Nanocomposites: Potential for Photocatalysis and Sensing. Nanomaterials 2022, 12, 294. [Google Scholar] [CrossRef] [PubMed]
  74. Li, G.; Wang, B.; Zhang, J.; Wang, R.; Liu, H. Er-Doped g-C3n4 for Photodegradation of Tetracycline and Tylosin: High Photocatalytic Activity and Low Leaching Toxicity. Chem. Eng. J. 2020, 391, 123500. [Google Scholar] [CrossRef]
  75. Pattanayak, D.S.; Pal, D.; Mishra, J.; Thakur, C. Noble Metal–Free Doped Graphitic Carbon Nitride (g-C3n4) for Efficient Photodegradation of Antibiotics: Progress, Limitations, and Future Directions. Environ. Sci. Pollut. Res. 2022, 1–13. [Google Scholar] [CrossRef] [PubMed]
  76. Dong, J.; Zhang, Y.; Hussain, M.I.; Zhou, W.; Chen, Y.; Wang, L.-N. G-C3n4: Properties, Pore Modifications, and Photocatalytic Applications. Nanomaterials 2022, 12, 121. [Google Scholar] [CrossRef] [PubMed]
  77. Malakootian, M.; Nasiri, A.; Amiri Gharaghani, M. Photocatalytic Degradation of Ciprofloxacin Antibiotic by TiO2 Nanoparticles Immobilized on a Glass Plate. Chem. Eng. Commun. 2020, 207, 56–72. [Google Scholar] [CrossRef]
  78. Campoli-Richards, D.M.; Monk, J.P.; Price, A.; Benfield, P.; Todd, P.A.; Ward, A. Ciprofloxacin. Drugs 1988, 35, 373–447. [Google Scholar] [CrossRef]
  79. Fief, C.A.; Hoang, K.G.; Phipps, S.D.; Wallace, J.L.; Deweese, J.E. Examining the Impact of Antimicrobial Fluoroquinolones on Human DNA Topoisomerase IIα and IIβ. ACS Omega 2019, 4, 4049–4055. [Google Scholar] [CrossRef]
  80. Yu, X.; Zhang, J.; Zhang, J.; Niu, J.; Zhao, J.; Wei, Y.; Yao, B. Photocatalytic Degradation of Ciprofloxacin Using Zn-Doped Cu2O Particles: Analysis of Degradation Pathways and Intermediates. Chem. Eng. J. 2019, 374, 316–327. [Google Scholar] [CrossRef]
  81. Pattnaik, S.P.; Behera, A.; Martha, S.; Acharya, R.; Parida, K. Facile Synthesis of Exfoliated Graphitic Carbon Nitride for Photocatalytic Degradation of Ciprofloxacin Under Solar Irradiation. J. Mater. Sci. 2019, 54, 5726–5742. [Google Scholar] [CrossRef]
  82. Chopra Ian; Roberts Marilyn Tetracycline Antibiotics: Mode of Action, Applications, Molecular Biology, and Epidemiology of Bacterial Resistance. Microbiol. Mol. Biol. Rev. 2001, 65, 232–260. [CrossRef] [Green Version]
  83. Florou, D.T.; Mavropoulos, A.; Dardiotis, E.; Tsimourtou, V.; Siokas, V.; Aloizou, A.-M.; Liaskos, C.; Tsigalou, C.; Katsiari, C.; Sakkas, L.I.; et al. Tetracyclines Diminish In Vitro IFN-γ and IL-17-Producing Adaptive and Innate Immune Cells in Multiple Sclerosis. Front. Immunol. 2021, 12, 4099. [Google Scholar] [CrossRef] [PubMed]
  84. Tettey, M.; Sereboe, L.; Edwin, F.; Frimpong-Boateng, K. Tetracycline Pleurodesis for Malignant Pleural Effusion-A Review of 38 Cases. Ghana Med. J. 2005, 39, 128–131. [Google Scholar]
  85. Daghrir, R.; Drogui, P. Tetracycline Antibiotics in the Environment: A Review. Environ. Chem. Lett. 2013, 11, 209–227. [Google Scholar] [CrossRef]
  86. Wang, P.; Yap, P.-S.; Lim, T.-T. C–N–S Tridoped TiO2 for Photocatalytic Degradation of Tetracycline Under Visible-Light Irradiation. Appl. Catal. A Gen. 2011, 399, 252–261. [Google Scholar] [CrossRef]
  87. Wang, W.; Fang, J.; Shao, S.; Lai, M.; Lu, C. Compact and Uniform TiO2@g-C3n4 Core-Shell Quantum Heterojunction for Photocatalytic Degradation of Tetracycline Antibiotics. Appl. Catal. B Environ. 2017, 217, 57–64. [Google Scholar] [CrossRef]
  88. Chen, F.; Yang, Q.; Sun, J.; Yao, F.; Wang, S.; Wang, Y.; Wang, X.; Li, X.; Niu, C.; Wang, D.; et al. Enhanced Photocatalytic Degradation of Tetracycline by AgI/BiVO4 Heterojunction Under Visible-Light Irradiation: Mineralization Efficiency and Mechanism. ACS Appl. Mater. Interfaces 2016, 8, 32887–32900. [Google Scholar] [CrossRef]
  89. Ezelarab, H.A.A.; Abbas, S.H.; Hassan, H.A.; Abuo-Rahma, G.E.-D.A. Recent Updates of Fluoroquinolones as Antibacterial Agents. Arch. Der. Pharm. 2018, 351, 1800141. [Google Scholar] [CrossRef]
  90. Janecko, N.; Pokludova, L.; Blahova, J.; Svobodova, Z.; Literak, I. Implications of Fluoroquinolone Contamination for the Aquatic Environment—A Review. Environ. Toxicol. Chem. 2016, 35, 2647–2656. [Google Scholar] [CrossRef] [Green Version]
  91. Pretali, L.; Fasani, E.; Sturini, M. Current Advances on the Photocatalytic Degradation of Fluoroquinolones: Photoreaction Mechanism and Environmental Application. Photochem. Photobiol. Sci. 2022, 21, 899–912. [Google Scholar] [CrossRef]
  92. Sayed, M.; Shah, L.A.; Khan, J.A.; Shah, N.S.; Nisar, J.; Khan, H.M.; Zhang, P.; Khan, A.R. Efficient Photocatalytic Degradation of Norfloxacin in Aqueous Media by Hydrothermally Synthesized Immobilized TiO2/Ti Films with Exposed {001} Facets. J. Phys. Chem. A 2016, 120, 9916–9931. [Google Scholar] [CrossRef]
  93. Tang, J.; Wang, R.; Liu, M.; Zhang, Z.; Song, Y.; Xue, S.; Zhao, Z.; Dionysiou, D.D. Construction of Novel Z-Scheme Ag/FeTiO3/Ag/BiFeO3 Photocatalyst with Enhanced Visible-Light-Driven Photocatalytic Performance for Degradation of Norfloxacin. Chem. Eng. J. 2018, 351, 1056–1066. [Google Scholar] [CrossRef]
  94. Lv, X.; Yan, D.Y.S.; Lam, F.L.-Y.; Ng, Y.H.; Yin, S.; An, A.K. Solvothermal Synthesis of Copper-Doped BiOBr Microflowers with Enhanced Adsorption and Visible-Light Driven Photocatalytic Degradation of Norfloxacin. Chem. Eng. J. 2020, 401, 126012. [Google Scholar] [CrossRef]
  95. Tang, L.; Wang, J.; Zeng, G.; Liu, Y.; Deng, Y.; Zhou, Y.; Tang, J.; Wang, J.; Guo, Z. Enhanced Photocatalytic Degradation of Norfloxacin in Aqueous Bi2WO6 Dispersions Containing Nonionic Surfactant Under Visible Light Irradiation. J. Hazard. Mater. 2016, 306, 295–304. [Google Scholar] [CrossRef] [PubMed]
  96. Lalliansanga; Tiwari, D.; Lee, S.-M.; Kim, D.-J. Photocatalytic Degradation of Amoxicillin and Tetracycline by Template Synthesized Nano-Structured Ce3+@TiO2 Thin Film Catalyst. Environ. Res. 2022, 210, 112914. [Google Scholar] [CrossRef] [PubMed]
  97. Balarak, D.; Mengelizadeh, N.; Rajiv, P.; Chandrika, K. Photocatalytic Degradation of Amoxicillin from Aqueous Solutions by Titanium Dioxide Nanoparticles Loaded on Graphene Oxide. Environ. Sci. Pollut. Res. 2021, 28, 49743–49754. [Google Scholar] [CrossRef]
  98. Mirzaei, A.; Chen, Z.; Haghighat, F.; Yerushalmi, L. Magnetic Fluorinated Mesoporous g-C3n4 for Photocatalytic Degradation of Amoxicillin: Transformation Mechanism and Toxicity Assessment. Appl. Catal. B Environ. 2019, 242, 337–348. [Google Scholar] [CrossRef]
  99. Huang, D.; Sun, X.; Liu, Y.; Ji, H.; Liu, W.; Wang, C.-C.; Ma, W.; Cai, Z. A Carbon-Rich g-C3n4 with Promoted Charge Separation for Highly Efficient Photocatalytic Degradation of Amoxicillin. Chin. Chem. Lett. 2021, 32, 2787–2791. [Google Scholar] [CrossRef]
Figure 1. Sources of different emerging antibiotics pollutants in daily life. (Reproduced with permission from [19], copyright 2021, Elsevier).
Figure 1. Sources of different emerging antibiotics pollutants in daily life. (Reproduced with permission from [19], copyright 2021, Elsevier).
Ijms 23 08130 g001
Figure 2. General photocatalytic mechanism on the degradation of antibiotics by the formation of photo-induced charge carriers (e/h+) on the photocatalysts’ surface.
Figure 2. General photocatalytic mechanism on the degradation of antibiotics by the formation of photo-induced charge carriers (e/h+) on the photocatalysts’ surface.
Ijms 23 08130 g002
Figure 3. Band energy gaps of selected semiconductors. (Reproduced with permission from [43], copyright 2017, Springer Nature).
Figure 3. Band energy gaps of selected semiconductors. (Reproduced with permission from [43], copyright 2017, Springer Nature).
Ijms 23 08130 g003
Figure 4. The mechanism of coupled processes with bismuth-based compounds for antibiotic degradation. (Reproduced with permission from [26], copyright 2021, Elsevier).
Figure 4. The mechanism of coupled processes with bismuth-based compounds for antibiotic degradation. (Reproduced with permission from [26], copyright 2021, Elsevier).
Ijms 23 08130 g004
Figure 5. Crystal structure and optical properties of graphitic carbon nitride: (a) Schematic diagram of a perfect graphitic carbon nitride sheet constructed from melem units, (b) Experimental XRD pattern of the polymeric carbon nitride, revealing a graphitic structure with an interplanar stacking distance of aromatic units of 0.326 nm and (c) Ultraviolet-visible diffuse reflectance spectrum of the polymeric carbon nitride. Inset: Photograph of the photocatalyst. (Reproduced with permission from [72], copyright 2009, Springer Nature).
Figure 5. Crystal structure and optical properties of graphitic carbon nitride: (a) Schematic diagram of a perfect graphitic carbon nitride sheet constructed from melem units, (b) Experimental XRD pattern of the polymeric carbon nitride, revealing a graphitic structure with an interplanar stacking distance of aromatic units of 0.326 nm and (c) Ultraviolet-visible diffuse reflectance spectrum of the polymeric carbon nitride. Inset: Photograph of the photocatalyst. (Reproduced with permission from [72], copyright 2009, Springer Nature).
Ijms 23 08130 g005
Figure 6. The illustration of commonly investigated antibiotics in photocatalytic processes. (a) Ciprofloxacin. (b) Tetracycline. (c) Norfloxacin. (d) Amoxicillin. (Reproduced with permission from [1], copyright 2021, Elsevier).
Figure 6. The illustration of commonly investigated antibiotics in photocatalytic processes. (a) Ciprofloxacin. (b) Tetracycline. (c) Norfloxacin. (d) Amoxicillin. (Reproduced with permission from [1], copyright 2021, Elsevier).
Ijms 23 08130 g006
Figure 7. (a) The schematic diagram of the photocatalytic mechanism for the Zn-doped Cu2O (b) Photocatalytic activity of R2-Cu2O with different scavenges, inset is degradation rate of ciprofloxacin. (Reproduced with permission from [80], copyright 2019, Elsevier) (c) Diagram of band gap structure of bulk and exfoliated g-C3N4 (d) Photocatalytic degradation of ciprofloxacin over g-C3N4 as well as exfoliated g-C3N4. (Reproduced with permission from [81], copyright 2019, Springer Nature).
Figure 7. (a) The schematic diagram of the photocatalytic mechanism for the Zn-doped Cu2O (b) Photocatalytic activity of R2-Cu2O with different scavenges, inset is degradation rate of ciprofloxacin. (Reproduced with permission from [80], copyright 2019, Elsevier) (c) Diagram of band gap structure of bulk and exfoliated g-C3N4 (d) Photocatalytic degradation of ciprofloxacin over g-C3N4 as well as exfoliated g-C3N4. (Reproduced with permission from [81], copyright 2019, Springer Nature).
Ijms 23 08130 g007
Figure 8. (a) Proposed heterojunction differences between TCN and TCN(mix) (b) Photocatalytic degradation efficiencies of tetracycline by employing TiO2, TCN, TCN (mix), and g-C3N4 as the photocatalysts under the xenon lamp irradiation. (Reproduced with permission from [87], copyright 2017, Elsevier). Schematic Illustration of the Mechanism for the Photocatalytic Degradation of tetracycline under Visible Light Irradiation over AgI/BiVO4 Nanocomposite: (c) Traditional Model and (d) Z-Scheme Heterojunction System. (Reproduced with permission from [88], copyright 2016, American Chemical Society).
Figure 8. (a) Proposed heterojunction differences between TCN and TCN(mix) (b) Photocatalytic degradation efficiencies of tetracycline by employing TiO2, TCN, TCN (mix), and g-C3N4 as the photocatalysts under the xenon lamp irradiation. (Reproduced with permission from [87], copyright 2017, Elsevier). Schematic Illustration of the Mechanism for the Photocatalytic Degradation of tetracycline under Visible Light Irradiation over AgI/BiVO4 Nanocomposite: (c) Traditional Model and (d) Z-Scheme Heterojunction System. (Reproduced with permission from [88], copyright 2016, American Chemical Society).
Ijms 23 08130 g008
Figure 9. (a) Photocatalysis Mechanism of {001} Faceted TiO2/Ti Film. (Reproduced with permission from [92], copyright 2016, American Chemical Society) (b) Possible mechanism diagram on Z-scheme Ag/FeTiO3/Ag/BiFeO3 system (c) Effect of mass ratio of FeTiO3 and BiFeO3 and (d) degradation reaction kinetics on photocatalytic activity (2.0 wt.% Ag; 5.0 mg/L norfloxacin; 1.0 g/L catalyst). (Reproduced with permission from [93], copyright 2018, Elsevier).
Figure 9. (a) Photocatalysis Mechanism of {001} Faceted TiO2/Ti Film. (Reproduced with permission from [92], copyright 2016, American Chemical Society) (b) Possible mechanism diagram on Z-scheme Ag/FeTiO3/Ag/BiFeO3 system (c) Effect of mass ratio of FeTiO3 and BiFeO3 and (d) degradation reaction kinetics on photocatalytic activity (2.0 wt.% Ag; 5.0 mg/L norfloxacin; 1.0 g/L catalyst). (Reproduced with permission from [93], copyright 2018, Elsevier).
Ijms 23 08130 g009
Figure 10. (a) Possible mechanism of amoxicillin degradation at GO/TiO2 surface. (Reproduced with permission from [97], copyright 2021, Springer Nature). Photocatalytic degradation kinetics of amoxicillin by the synthesized materials under (b) visible light and (c) simulate solar light. (d) amoxicillin degradation rate constants under solar and visible light. (Reproduced with permission from [99], copyright 2021, Elsevier).
Figure 10. (a) Possible mechanism of amoxicillin degradation at GO/TiO2 surface. (Reproduced with permission from [97], copyright 2021, Springer Nature). Photocatalytic degradation kinetics of amoxicillin by the synthesized materials under (b) visible light and (c) simulate solar light. (d) amoxicillin degradation rate constants under solar and visible light. (Reproduced with permission from [99], copyright 2021, Elsevier).
Ijms 23 08130 g010
Table 1. Comparison of the photocatalytic activity of different photocatalysts for antibiotic degradation.
Table 1. Comparison of the photocatalytic activity of different photocatalysts for antibiotic degradation.
AntibioticCatalystResultsDegradation MechanismRef.
CiprofloxacinZn-doped Cu2O94.6% Ciprofloxacindegraded in 240 minMechanism by the •OH radical and h+[80]
CiprofloxacinExfoliated g-C3N478% Ciprofloxacindegraded in 60 minMechanism by the •O2 radical and h+[81]
TetracyclineHeterostructured AgI/BiVO494.91% Tetracycline degraded in 60 minMechanism by the •OH, •O2-radical and h+[88]
TetracyclineHeterostructured TiO2/g-C3N4100 mg TiO2/g-C3N4can decompose 2 mg Tetracycline in 9 min (2.2 mg/min)Mechanism by the •O2 radical and h+[87]
TetracyclineC–N–S-TiO2>99% Tetracycline degraded in 360 minMechanism by the •O2 radical and h+[86]
NorfloxacinCu-doped BiOBr96.5% Norfloxacin degraded in 150 minMechanism by the h+[94]
NorfloxacinZ-scheme Ag/FeTiO3/Ag/BiFeO3Photocatalytic degradation rate of Norfloxacin is 0.64 × 10−2 min−1Mechanism by the •OH radical and h+[93]
AmoxicillinGraphene Oxide/TiO291.25% Amoxicillin degraded in 60 minMechanism by the h+[97]
AmoxicillinCarbon-rich g-C3N4 nanosheetsPhotocatalytic degradation rate of Amoxicillin is 0.47 × 10−2 min−1Mechanism by the •O2 radical[99]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bai, X.; Chen, W.; Wang, B.; Sun, T.; Wu, B.; Wang, Y. Photocatalytic Degradation of Some Typical Antibiotics: Recent Advances and Future Outlooks. Int. J. Mol. Sci. 2022, 23, 8130. https://doi.org/10.3390/ijms23158130

AMA Style

Bai X, Chen W, Wang B, Sun T, Wu B, Wang Y. Photocatalytic Degradation of Some Typical Antibiotics: Recent Advances and Future Outlooks. International Journal of Molecular Sciences. 2022; 23(15):8130. https://doi.org/10.3390/ijms23158130

Chicago/Turabian Style

Bai, Xue, Wanyu Chen, Bao Wang, Tianxiao Sun, Bin Wu, and Yuheng Wang. 2022. "Photocatalytic Degradation of Some Typical Antibiotics: Recent Advances and Future Outlooks" International Journal of Molecular Sciences 23, no. 15: 8130. https://doi.org/10.3390/ijms23158130

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

Article Metrics

Back to TopTop