Next Article in Journal
Mesenchymal Stem/Stromal Cells Derived from Dental Tissues Mediate the Immunoregulation of T Cells through the Purinergic Pathway
Previous Article in Journal
The Potential Use of Arsenic Trioxide in the Treatment of Systemic Lupus Erythematosus
Previous Article in Special Issue
Far-Ultraviolet Light at 222 nm Affects Membrane Integrity in Monolayered DLD1 Colon Cancer Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Photodegradation of Amoxicillin in Aqueous Systems: A Review

by
Mohammad Ashraf Ali
and
Ibrahim M. Maafa
*
Department of Chemical Engineering, College of Engineering and Computer Sciences, Jazan University, Jazan 45142, Saudi Arabia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(17), 9575; https://doi.org/10.3390/ijms25179575
Submission received: 24 June 2024 / Revised: 28 August 2024 / Accepted: 28 August 2024 / Published: 4 September 2024
(This article belongs to the Special Issue Recent Advances in Photolysis and Photodegradation)

Abstract

:
Amoxicillin (AMX) is utilized in the treatment of several infectious diseases, and its concentration in wastewater has increased quite significantly over the years, posing high health hazards for humans and other living organisms. Investigations are in progress globally to eliminate AMX and other related pollutants using several methods that include adsorption, photolysis, photocatalytic degradation, photoelectrocatalytic degradation, and electrochemical conversion. AMX can be eliminated efficiently from the environment using photodegradation, either by photolysis or a photocatalytic process. Several types of semiconductor NMs have been used to eliminate AMX and other related drugs present in wastewater. This review spans the photodegradation studies conducted during the years 2018–2024 to degrade and eliminate AMX in aquatic systems. Several studies have been reported to eliminate AMX from different water streams. These studies are categorized into TiO2-containing and non-TiO2-based catalysts for better comparison. A section on photolysis is also included, showing the use of UV alone or with H2O2 or PS without using any nanomaterial. A tabulated summary of both types of catalysts showing the catalysts, reaction conditions, and degradation efficiency is presented. Researchers have used a variety of reaction conditions that include radiation types (UV, solar, and visible), pH of the solution, concentration of AMX, number of nanomaterials, presence of other additives and activators such as H2O2 as oxidant, and the influence of different salts like NaCl and CaCl2 on the photodegradation efficiency. TiO2 was the best nanomaterial found that achieved the highest degradation of AMX in ultraviolet irradiation. TiO2 doped with other nanomaterials showed very good performance under visible light. WO3 was also used by several investigators and found quite effective for AMX degradation. Other metal oxides used for AMX elimination were derived from molybdenum, zinc, manganese, copper, cerium, silver, etc. Some researchers have used UV and/or visible irradiation or sunlight, without using solid catalysts, in the presence of oxidants such as H2O2. A summarized description of earlier published reviews is also presented.

1. Introduction

Researchers worldwide are attempting to remove organic contaminants from wastewater using a variety of strategies, including adsorption, photolysis, photocatalysis, photoelectrocatalysis, and electrochemical conversion. These organic contaminants include dyes, medicines, and surfactants, which have been found in many aquatic environments worldwide. These organic pollutants pose toxicity to aquatic and land-dwelling creatures. The conventional treatment processes for wastewater eliminate these substances only partly, and they generally remain in trace concentrations in treated effluent [1]. Photocatalysis is one of the best techniques for the complete elimination of organic pollutants from wastewater. Organic dyes such as methylene blue, crystal violet, rhodamine B, methyl orange, alizarin red, Eriochrome Black T indicator, reactive red 250, and Congo red have been photodegraded successfully using different types of nanocomposites, which include Co-SnO2-loaded, sulfur-doped graphitic carbon nitride [2]; spinel MOF [3]; ZnO [4]; reactive red 250 [5]; Ag/Mn–ZnO [6]; Ni/g-C3N4/ZnO [7] and Mn/g-C3N4/ZnO [4]; Ba-doped Mg ferrites [8]; and WO3/NiWO4 [9]. Similarly, pharmaceuticals present in wastewater have been photodegraded by several researchers using a variety of nanomaterials. Husain et al. successfully degraded ibuprofen and ofloxacin in wastewater using the photo-Fenton process with MnO2 and ozonation, achieving over 90% elimination [10].

2. Reported AMX Photodegradation Studies

One of the most-used antibiotics is AMX for the treatment of urinary, gastrointestinal, skin, and respiratory bacterial infections [11]. The World Health Organization has ranked AMX as one of the medically important antimicrobial medicines for human use, and this ranking has encouraged higher AMX use that has led to increased exposure of AMX in the environment [12]. Several studies have been reported to eliminate AMX from wastewater streams. We have categorized the studies into TiO2-containing and non-TiO2-based catalysts for a clear understanding.

2.1. TiO2-Containing Catalysts

Several researchers have used TiO2 alone or TiO2 doped with metals and mixed with metal oxides and carbon nitrides. These studies are reported here. At the end of this section, Table 1 presents a summary of the performance of some of the TiO2-based nanomaterials (NMs) and nanocomposites used for the photodegradation of AMX.
Alshandoudi et al. investigated the efficiency of AMX removal through photocatalytic degradation using nanoTiO2 and nanoTiO2/nanohydroxyapatite composites and achieved 85.3 and 99.5% degradation, respectively, within 90 min using catalyst concentrations of 0.9 g/L [13].
Lalliansanga et al. investigated Ce3+/TiO2 thin-film photocatalysts to degrade tetracycline and AMX under UVA irradiation [14]. The presence of Ce3+ was found to be quite significant to enhance photocatalytic degradation of the antibiotics.
Mhemid, et al. performed photodegradation of AMX using TiO2 and N-doped TiO2 in an aqueous solution under solar irradiation [15]. The N-doped TiO2 was synthesized from urea by the sol-gel process. The optimum conditions achieved were pH 5 and H2O2 concentration of 400 mg/L, using 50 mg catalyst and AMX concentration of 25 mg/L, for a photodegradation efficiency of 95.8% with N-TiO2 compared with 89.3% for TiO2.
Ellepola et al. investigated a TiO2 anatase for AMX photodegradation and observed AMX degradation increased by 4.5-fold in the presence of light [16]. Figure 1 shows AMX degradation plots with TiO2 anatase and kaolinite under dark and irradiation, and the pH variation of the AMX, as a function of reaction time. Almost complete degradation of AMX was achieved using TiO2 anatase under UV–Vis irradiation in 300 h. Figure 2 shows how the AMX interacted with the mineral surface in the presence of light and degraded into simpler products. Figure 3 shows the proposed degradation of AMX into different compounds under dark and solar radiation in the absence of minerals.
Wang et al. prepared 2D/2D TiO2/Bi2MoO6 material into nanosheets that showed high effectiveness and steady performance for AMX photodegradation, which was 18.2 and 5.7 times higher than TiO2 and Bi2MoO6, respectively [17]. Figure 4 presents a visualization of the 2D/2D TiO2/Bi2MoO6 photocatalyst and the process of AMX photodegradation using TiO2/Bi2MoO6 nanomaterials.
Zamani et al. described a visible-light-activated Ag/Ag2O/TiO2 heterostructure photocatalyst for the degradation of AMX [18]. The results revealed that the greatest AMX photodegradation achieved was 97.9% within 80 min at pH 6 with an initial AMX content of 20 mg/L.
Aissani et al. developed a supported catalyst TiO2/Mg-Fe-LDH and used it for the photodegradation of AMX under UVA irradiation in basic conditions [19]. The catalyst exhibited excellent photodegradation efficiency, and high recyclability was retained after four cycles.
Hussien et al. synthesized core/shell g-C3N4@TiO2 heterojunction using melamine pyrolysis and used it for AMX photodegradation under visible light irradiation [20]. The photocatalyst g-C3N4@TiO2 exhibited 100% degradation of AMX in 20 min. The AMX photodegradation using g-C3N4@TiO2 was three times higher compared with g-C3N4 under simulated solar irradiation.
Gao et al. synthesized a ternary Ag/TiO2/M-g-C3N4 nanocomposite that showed the highest degradation efficiency for AMX compared with its components [21].
Alkaim et al. [22] performed photodegradation of AMX over TiO2 under UV irradiation and achieved 98.0% removal after one hour using an initial AMX concentration of 10 mg/L at 25 °C and 0.2 g/L of the catalyst. Higher initial AMX concentrations experienced lower degradation efficiencies, as shown in Figure 5.
Salimi et al. investigated the photodegradation of AMX with Pt and Bi co-doped TiO2 photocatalysts. Among the developed photocatalysts, 1Pt-5Bi-TiO2 photodegraded AMX at 87.7% under visible light irradiation [23].
Yılmaz et al. produced co-doped TiO2 and utilized it to photodegrade AMX in an aqueous solution using UVC and visible irradiation. AMX was completely degraded during 4 h of UVC irradiation and 5 h of visible irradiation [24].
Bergamonti et al. evaluated TiO2-supported chitosan scaffolds (TiO2/CS) for AMX photodegradation under UV–Vis irradiation [25]. TiO2 chitosan scaffolds were prepared by 3D printing through dispersion of 6.0% w/v chitosan and 1.0% w/v commercial P25-TiO2 in powder form. The TiO2/CS system showed a high recycling photodegradation efficiency.
Chinnaiyan et al. performed photodegradation of AMX-containing synthetic hospital wastewater using a 200 mL photoreactor, TiO2 as a photocatalyst, and UV irradiation at 365 nm using a 125 W mercury vapor lamp [26]. The maximum AMX degradation achieved was 90.0% at pH 7.6, using a TiO2 dosage of 563 mg/L and an initial concentration of 10 mg/L within 150 min irradiation.
Wahyuni et al. synthesized Cu-TiO2 and used it under visible light irradiation for AMX degradation in water [27]. It was found that 90% of 10 mg/L AMX was degraded using 0.40 g/L of the Cu-TiO2 photocatalyst (having 4.56 mg Cu/g of TiO2) after 24 h at pH 6. Figure 6 shows the AMX photodegradation efficiency under UV and visible light using 100 mL AMX solution having 10 mg/L concentration after 24 h of irradiation at pH 6. In the case of TiO2, UV irradiation produced more degradation than visible radiation, while for Cu-TiO2, visible light irradiation produced more degradation than UV radiation.
Fazilati used TiO2, ZnO, and GO-Fe3O4 for AMX photodegradation under UV irradiation and achieved maximum removal efficiency of 27.6%, 48.6%, and 87.1%, respectively, under the optimum conditions of AMX 15 mg/L at 18 W of UV irradiation [28]. It was concluded that using the catalyst GO-Fe3O4 under UV irradiation, the solution pH did not affect the photodegradation efficiency of AMX (Figure 7). Figure 8 shows that the photodegradation of AMX increased from 6 W to 18 W using TiO2, ZnO, and GO-Fe3O4 catalysts. For TiO2 and GO-Fe3O4, more photodegradation was observed than with ZnO while increasing from 6 W to 18 W.
Zhou et al. carried out photoelectrocatalytic degradation of AMX in water successfully and achieved maximum degradation of 79% using 200 mL AMX solution having 100 mg/L concentration with 0.5 g CaTiO3, 0.058 g NaCl added as the electrolyte at pH 3, and 45 °C for 120 min irradiation [29] (Figure 9). Using the photocatalytic process, the degradation efficiency was 35.8% only. Figure 10 shows the degradation efficiency for a 200 mL AMX solution with a 100 mg/L concentration and 0.5 g CaTiO3 at room temperature at pH 3, 7, and 10. The highest degradation was observed at an acidic pH of 3, which is contrary to the other reported research, and it could be due to the presence of CaTiO3. Figure 11 exhibits the effect of CaTiO3 concentration on the AMX degradation in 200 mL aqueous solution having 100 mg/L AMX at pH 3 and room temperature. A higher concentration of CaTiO3 produced a higher degradation of AMX. Figure 12 shows the effect of temperature on degradation efficiency for 200 mL aqueous solution having 100 mg/L AMX with 0.5 g CaTiO3 and 0.058 g NaCl at pH 3 with a current intensity of 0.03 mA/cm2. Higher degradation was observed by increasing the temperature. The photoelectrocatalytic degradation data fit the first-order kinetics.
Verma et al. investigated AMX degradation using TiO2 photocatalysis in aqueous solution using UVA (365 nm) and sunlight [30]. The optimal conditions to achieve 80% AMX degradation were a TiO2 dosage of 450 mg/L, an AMX concentration of 30 mg/L, an H2O2 concentration of 150 mg/L, and pH 7.0 under UV irradiation (672 W/m2). Substantial enhancement in the degradation rate was observed using H2O2 and sonication (40 KHz), but the maximum degradation remained the same (Figure 13).
Balarak et al. reported performing AMX photocatalytic degradation under UV irradiation using TiO2 NPs loaded on graphene oxide (GO/TiO2) [31]. The AMX degradation efficiency was almost 100% at pH 6, using a GO/TiO2 concentration of 0.4 g/L and AMX concentration of 50 mg/L, at UV irradiation intensity of 36 W. The catalyst showed exceptional recyclability for four repeated cycles. The detection of NH4+, NO3, and SO42− ions confirmed good mineralization of the AMX. Figure 14a illustrates the possible mechanism of AMX degradation at the GO/TiO2 surface.
Huang et al. prepared novel carbon-rich g-C3N4 nanosheets with large surface areas, which showed superior photocatalytic activity for AMX degradation under solar light [32], as shown in Figure 14b–d. The best degradation results were obtained using the catalyst C-CN90, which was prepared by pyrolysis of a mixture of urea and 90 mg of 1,3,5-cyclohexanetriol for the optimum composite.
Table 1 lists the performance of TiO2-based nanomaterials (NMs) and nanocomposites used for photodegradation of AMX and their reaction conditions. The titania-based catalysts consist of TiO2 NPs and TiO2 doped with other active metals or mixed with other metal oxides. In the case of pure TiO2, UV radiation was used in most of the studies. When mixed with other metal oxides and doped with other metals, visible radiation was employed preferably, due to the reduction in the bandgap of the catalysts as a result of interaction between TiO2 and other metals and metal oxides.
Table 1. Performance of TiO2-based nanomaterials (NMs) and nanocomposites used for photodegradation of AMX.
Table 1. Performance of TiO2-based nanomaterials (NMs) and nanocomposites used for photodegradation of AMX.
NoCatalyst TypeProcess ConditionsDegradation Efficiency, %Degradation Time, hRef.
1.TiO2 NPs, 0.2 g/LUV light irradiation, AMX 10 mg/L98.0%1[22]
2.TiO2 NPs, 450 mg/LUV light irradiation (672 W/m2), AMX 30 mg/L, H2O2 (30% w/v) 150 mg/L, pH 7.0804[30]
3.TiO2 NPs, 563 mg/LUV irradiation at 365 nm (125 W Hg vapor lamp), 200 mL photoreactor, pH 7.6, AMX 10 mg/L902.5[26]
4.TiO2 NPs, 0.9 g/LUV light irradiation85.31.5[13]
5.TiO2 NPs, 40 mgVisible light irradiation, AMX 10 mg/L, 100 mL solution, pH 65024[28]
6.N-TiO2, 50 mgSolar light irradiation, H2O2 400 mg/L, pH 5,
AMX 25 mg/L
95.8%1[15]
7.nanoTiO2/nano-hydroxyapatite composite, 0.9 g/LUV light irradiation99.51.5[13]
8.g-C3N4@TiO2 Visible light irradiation 100%0.33[20]
9.Ag/Ag2O/TiO2Visible light irradiation, AMX 20 mg/L, pH 697.9%1.33[18]
10.Cu-TiO2 NPs, 0.40 g/LVisible light irradiation, AMX 10 mg/L, 100 mL solution, pH 69024[28]
11.10%Cu2O/TiO2 nanotubes, 1.5 g/LVisible light irradiation, AZT 100 μg/mL, pH 71001.5[33]
12.Zn-doped TiO2 UV light irradiation, H2O2 550 mL/L, pH 3951.5[34]
13.TiO2,
ZnO,
GO-Fe3O4
UV light irradiation 18 W, AMX 15 mg/L, pH 527.6
48.6
87.1
0.25
0.25
0.25
[29]
14.CaTiO3 (0.5 g) and NaCl (0.06 g) as electrolytesUV light irradiation, intensity 15 W, photoelectrocatalytic, 200 mL AMX 100 mg/L, pH 3, 45 °C792[30]

2.2. Non-TiO2-Based Catalysts

Many researchers have used non-TiO2-based catalysts for the photodegradation of AMX. The studies reported in this section are summarized in Table 2.
Gadore et al. synthesized a novel integrated photocatalytic adsorbent by employing tea leaf extract (biochar) as a stabilizing agent and SnS2 as a photocatalyst for AMX photodegradation [35]. The overall removal efficiency of AMX was 93.7% achieved within 90 min. Ramos et al. synthesized Fe NPs using pumpkin (Tetsukabuto) peel extract and used them for the photocatalytic degradation of AMX under UV light irradiation [36]. The optimum conditions achieved for the 60% AMX degradation were 2.5 g/L of Fe NPs, initial AMX concentration 10 mg/L, pH 5.6, and 60 min irradiation time. Mmelesi et al. synthesized ZnxCo1−xFe2O4 (x = 0.0 to 0.5) NPs by co-precipitation and calcination [37]. The performance of the prepared NPs was 89% for AMX photodegradation using visible light irradiation.
Asmaa et al. synthesized MIL-53(Al)/ZnO NPs for the photocatalytic degradation of AMX and utilized them under visible irradiation using a metal halide lamp [38]. Figure 15 shows the SEM photographs to illustrate the morphology of the MIL-53(Al)/ZnO, MIL-53Al, and pure ZnO. Figure 16 shows the degradation of AMX under different catalysts and confirms that the highest degradation was obtained using MIL-53(Al)/ZnO NPs. The removal of AMX was 100% within 60 min of irradiation at pH 4.5 using the catalyst dose of 1.0 g/L and initial AMX concentration of 10 mg/L. The charge transfer process between ZnO and MIL-53Al, and the degradation mechanism of AMX, are shown in Figure 17. The AMX degradation was decreased after five consecutive cycles from 78.2 to 62.1% using MIL-53(Al)/ZnO, as shown in Figure 18. The recyclability tests were conducted using 0.6 g/L of catalyst at pH 7.0, 50 mg/L of AMX, and an irradiation time of 60 min.
Figure 19 shows the degradation kinetics of AMX at pH 4.0, 5.5, and 7.0. There was no degradation of AMX in the dark at all three pH, while almost complete degradation occurred under simulated sunlight, after 40 h of irradiation. Figure 20 shows AMX degradation under simulated sunlight and in the dark in the presence of humic acids and different inorganic salts. In the presence of FeCl3 at concentrations of 500 mg/L, AMX showed total degradation, both under simulated sunlight and in the dark [39].
Utami et al. synthesized Bi2O3/Fe by a precipitation method using microwaves [40]. The best-performing sample was 3% Fe/Bi2O3, which provided 76.34% degradation efficiency of the AMX.
Le et al. synthesized vanadium pentoxide/graphitic carbon nitride and achieved 91.3% AMX degradation under solar light [41].
Thuan et al. prepared indium vanadate and silver deposited on graphitic carbon nitride (InVO4@Ag@g-C3N4) for photocatalytic degradation of AMX, which exhibited high photocatalytic performance and recyclability for AMX degradation [42].
Palas and Ersöz prepared Ag and Co-doped perylene diimide supramolecular catalysts for the photocatalytic removal of AMX from aqueous solutions [43]. The AMX photodegradation was 51.8% at pH 4.6, 0.52 g/L catalyst loading, and 10.3 mg/L AMX concentration. The AMX degradation fits a second-order kinetic model.
Moradi et al. studied CuO NPs and ZnO nanorods supported on g-C3N4 nanosheets (CZ@T-GCN) as novel photocatalysts for the degradation of AMX [44]. The results demonstrated enhanced degradation of AMX compared with individual components and showed complete degradation of AMX within 120 min of simulated sunlight irradiation using a catalytic dosage of 0.9 g/L at pH 7.0.
Shi et al. prepared a composite by loading carbon dots onto rod-like CoFe2O4 that efficiently degraded AMX by 97.5% within 80 min using photocatalytic activation of peroxymonosulfate under visible light irradiation [45].
Xu et al. reported a 2D/2D Bi2WO6/Ti3C2 MXene heterostructure that achieved complete degradation of AMX within 40 min [46] due to the generation of reactive oxygen species, mainly holes (h+) and superoxide anion radicals (O2).
Le et al. fabricated a CdS/NH4V4O10 composite by loading CdS quantum dots into the ultrathin NH4V4O10 nanobelts [47]. The synthesized composite possesses a large surface area and plentiful active sites for photodegradation. The optimal composite, 6-CdS/NH4V4O10, provided 94.4% degradation performance with 5.0 wt% AMX with high recyclability.
Le et al. synthesized carbon dot-loaded Sn2Ta2O7/SnO2 (C-dots/Sn2Ta2O7/SnO2) heterostructures that exhibited AMX (initial concentration 20 mg/L) photodegradation efficiency of 88.3% within 120 min under simulated solar light irradiation [48].
Dou et al. synthesized mesoporous carbon nitride and used it for the photodegradation of AMX and cefotaxime [49] and achieved excellent performance under visible light.
Samsudin et al. synthesized a g-C3N4/BiVO4 microflower and evaluated it for the photocatalytic degradation of AMX. It was found that 89.5% of the AMX was successfully degraded within 3 h under visible light irradiation [50].
Nguyen et al. reported a 100% photodegradation efficiency of AMX under simulated solar irradiation using WO3 at pH 4, a catalyst dosage of 0.104 g/L, and initial AMX concentration of 1.0 μM [51]. The photodegradation of AMX was found to follow the pseudo-first-order kinetics.
Chahkandi et al. synthesized thin BiVO4 nanorods and deposited them on a webbed stainless-steel surface [52]. The diameter of the BiVO4 nanorods was in the range of 100 to 150 nm, and it exhibited 97.5% AMX photodegradation in 1.5 h under visible light irradiation.
Abazari et al. prepared amine-functionalized Al-based nanocomposites (NH2-MOF@Sm2O3–ZnO) and used them efficiently for the photodegradation of AMX under ultrasound and visible light [53]. These nanocomposites showed stable photodegradation performance even after five recycles.
Gaim et al. reported the synthesis of Mn-doped Cu2O NPs using aloe vera leaf extract and used them for AMX photodegradation [54]. The AMX photodegradation achieved was 92.0% at pH 9, 15 mg/L initial concentration of AMX, and catalyst dosage of 1 g/L under sunlight irradiation. Figure 21 shows the SEM images of pure Cu2O and Mn-doped Cu2O NP, which reveal that their morphologies were octahedral and truncated octahedral, respectively. Figure 22 shows the effect of pH on the photocatalytic degradation of AMX (initial concentration 15 mg/L) using a catalyst dosage of 1.0 g/L. The degradation of AMX was highest at pH 9 due to the formation of large amounts of OH radicals at higher pH. Figure 23 exhibited the effect of the initial concentration of AMX on the photodegradation of AMX at pH 9 using 1.0 g/L catalyst concentration. The degradation was first increased with increasing AMX concentration up to 15 ppm and then decreased as the AMX concentration increased. Figure 24 shows the effect of catalyst concentration on the photocatalytic degradation of AMX (15 mg/L) at pH 8. Different amounts (0.5 to 2.5 g/L) of the catalyst were added to the AMX solution. Maximum degradation of AMX was observed at 1.0 g/L catalyst dosage. Figure 25 shows the photodegradation of AMX using Cu2O and Mn-Cu2O conducted at pH 9, using an AMX concentration of 15 mg/L and 1.0 g/L catalyst, which achieved 65.0% and 92.0% degradation, respectively, within 3 h of irradiation.
The photodegradation of AMX was performed using the photo-Fenton process with an effective irradiation time of 30 min. Nonetheless, AMX photodegradation with photo-Fenton is a costly process [55]. Furthermore, the effect of pH is very substantial in the AMX photodegradation with photo-Fenton [56].
Haddadou et al. synthesized Ba(Ti0.950Sc0.025Nb0.025)O3, termed as BTSN, as a tetragonal perovskite structure with a ferroelectric behavior and photoelectrochemical properties having a band gap of 2.60 eV [57]. The BTSN was tested for the photocatalytic oxidation of AMX under solar light irradiation, exhibited 72% conversion, and followed first-order kinetic. Figure 26 shows the photodegradation of AMX and the spectra showing AMX degradation with irradiation time.
Kattel et al. investigated the degradation of AMX by the UVC (280 nm) or UVC/Fe2+-activated S2O82− processes [58]. The optimal conditions were the concentration of S2O82− and Fe2+, 15400 μM and 40 μM, respectively, and the reaction followed pseudo-first-order kinetics. AMX degradation achieved was more than 99% in 2 h irradiation time. Both radicals, SO4 and HO, were found to be present in the reaction system, but SO4 was in the major amount.
Beshkar et al. reported employing a CuI/FePO4 heterojunction nanocomposite to degrade AMX under simulated solar exposure [59]. The hybrid CuI/FePO4 NPs demonstrated an energy gap of 3.03 eV and a photodegradation efficiency of 90%.
Silva et al. synthesized metal-free polymeric carbon nitrides (PCNs) using melamine and thiourea precursors [60] and evaluated them for AMX photodegradation under visible light irradiation (16 mW/cm2). The thiourea-based material showed better efficiency of 100% removal within 48 h and excellent recyclability.
Mirzaei et al. investigated a fluorinated graphitic carbon nitride photocatalyst having magnetic properties for the degradation of AMX in water [61]. The photocatalyst exhibited easy separation from the aqueous solution by a magnet and improved photocatalytic degradation of AMX. The study showed that a UV lamp (10 W) was more efficient for AMX degradation than a 500 W visible lamp.
Table 2 lists the reaction conditions and performance of non-TiO2-based nanomaterials and nanocomposites employed for AMX photodegradation. These catalysts comprise metal nanoparticles and single and multiple metal and metal oxides derived from molybdenum, zinc, silver, manganese, copper, cerium, bismuth, vanadium, and tungsten. The doping and mixing of the metals were shown to be very effective in the photodegradation of AMX and to have enhanced degradation efficiency.
Table 2. Performance of non-TiO2-based nanomaterials (NMs) and nanocomposites used for photodegradation of AMX.
Table 2. Performance of non-TiO2-based nanomaterials (NMs) and nanocomposites used for photodegradation of AMX.
NoCatalyst TypeProcess ConditionsDegradation Efficiency, %Degradation Time, hRef.
1.MIL-53(Al)/ZnO nanocomposite, 1.0 g/LVisible light irradiation, AMX 10 mg/L, pH 4.51001[39]
2.Iron NPs (Fe-NPs), 2.5 g/LUV light irradiation, AMX 10 mg/L, pH 5.6601[37]
3.C-dots/Sn2Ta2O7/SnO2 nanocompositeSimulated solar light irradiation, AMX 20 mg/L88.3%2[49]
4.g-C3N4/BiVO4
microflower
Visible light irradiation89.53[50]
5.WO3 NPs, 0.104 g/LSunlight irradiation, AMX 1.0 μM, 50 mL solution, pH 41003[51]
6.WO3 NPs, 0.104 g/LSimulated solar irradiation (300 W xenon lamp), AMX 1.0 μM, pH 4, follows pseudo-first-order kinetics, temp. 25 °C1003[52]
7.BiVO4 thin filmVisible light irradiation97.51.5[53]
8.Mn-doped Cu2O NPs,
1.0 g/L
Sunlight irradiation, pH 9, AMX 15 mg/L923[55]
9.Cu2O NPs, 1.0 g/LSunlight irradiation, pH 9, AMX 15 mg/L 653[55]
10.gamma-Fe2O3@SiO2@ZIF8-Ag nanocomposite, catalyst 0.4 g/LVisible light radiation, pH 5, AMX 10 mg/L1001[62]
11.1% Na-doped hydroxyapatite, 0.1 gAMX 20 mg/L, 50 mL solution, pH 7602[63]

2.3. Photolysis Studies

Vettorello et al. [64] investigated the degradation of AMX by photolysis and assessed the formation of breakdown organic structures and their metabolite toxicity. Solutions of 100 µg/L of AMX were subjected to different conditions of UV irradiation with different powers in a quartz batch reactor at pH 4, 7, and 9. The HPLC-MS analysis determined the degradation of AMX and the reaction products generated. The best 90% AMX degradation was achieved using 95 W lamp irradiation located 5–10 cm from the AMX solution within 10 min at pH 9. Two products originating under different pH conditions were identified, namely, AMX penicilloic acid and AMX 4-hydroxyphenilglyl. The resulting solution containing these products was found to have no toxicity.
Zhang et al. [65] studied the reaction kinetics, degradation routes, and antibacterial activity of AMX in UV/H2O2 and UV/persulfate (S2O82−, PS) systems. It was found that UV irradiation alone did not affect AMX degradation. However, adding H2O2 or PS considerably improved degradation efficiency by producing HO. and SO4 radicals. At neutral pH, UV, HO., and SO4 contributed 7.3%, 22.8%, and 69.9% to AMX degradation, respectively. The AMX degradation pathways using UV/H2O2 and UV/PS systems included processes such as hydroxylation, hydrolysis, and decarboxylation. UV/H2O2 photolysis proved to be more cost-effective than UV/PS for AMX degradation. The antibacterial activity of the AMX solution dropped significantly after UV/H2O2 and UV/PS treatment, indicating that the AMX breakdown products were not significantly hazardous.
Arsand et al. [66] performed photolysis of AMX and ampicillin surface water samples and analyzed the degradation and the breakdown products by HPLC-MS. The results displayed that the degradation of AMX and ampicillin was almost complete in 48 h using river water, and more than 65 transformation products of amoxicillin and ampicillin were detected and identified. The photolysis kinetics of AMX and ampicillin in river water followed pseudo-first-order kinetics.
Timm et al. [67] focused their study on the photolysis of the four β-lactam antibiotics including AMX, ampicillin, penicillin V, and piperacillin under simulated environmental conditions. It was observed that all the investigated β-lactam antibiotics were degradable by simulated sunlight (1 kW/m2). Structure interpretation of the transformation products achieved using HPLC-MS showed that the hydrolysis of the β-lactam ring was the primary transformation reaction, followed by the elimination of carboxylic and dimethyl thiazolidine carboxylic acid. The loss of bactericide activity of the irradiated solutions of amoxicillin, ampicillin, and piperacillin suggested that the transformation of the β-lactam ring is responsible for the antibiotic effect of these antibiotics.

3. Perspective from Reviews Published Earlier in This Area

Pirsaheb et al. developed a systematic review of the photo-Fenton process for aqueous AMX degradation, revealing that its efficiency depends on factors like pH, light source characteristics, and AMX concentration [68]. Acidic pH values show the highest efficiency. Energy-efficient light sources reduce catalyst and oxidizer use. Combining AOPs like photo-Fenton, electro-Fenton, and photo-electro-Fenton increases mineralization efficiency. Synthetic wastewater shows higher efficiency.
Qutob et al. published a review of radical and non-radical degradation of AMX using different oxidation process systems [69]. The review explores degradation mechanisms, efficiency, catalyst stability, AMX byproduct formation, and toxicity. Pharmaceutical compounds like AMX are increasingly being studied for their potential environmental leakage. Due to its low metabolic rate, 80–90% of AMX remains unmetabolized. Advanced oxidation processes, including photocatalytic, ultrasonic, electro-oxidation, and partials, are effective in degrading AMX. High pH, temperature, concentration, oxidants, catalysts, and doping ratios can inhibit degradation. Hybrid systems like photo-electro, photo-Fenton, and electro-Fenton are recommended for successful oxidation. This review could provide valuable insights for future researchers.
Aryee et al. reviewed the detection and removal of AMX in wastewater, discussing its entry into the environment and potential effects [70]. They discussed detection methods and remediation techniques, with advanced oxidation processes and constructed wetlands being the most and least applied methods. The review also highlighted research gaps and recommendations for further studies on AMX removal, aiming to inspire further research.
Manikanika et al. reviewed the activity of ZnO NPs for dyes and drug photodegradation [71]. Several metal oxides performed as photocatalysts, but ZnO was found to achieve total degradation and mineralization. ZnO NPs have a large band gap that is 3.37 eV and 60 meV excitation binding energy and have performed very well in photocatalysis. The reported maximum degradation efficiency using ZnO NPs has been reported at 100% and 98% for dyes and drugs, respectively.
Chen et al. reported that antibiotics have been extensively detected in the aquatic environment as one of the major pollutants [72]. For the elimination of antibiotics, photocatalysis using sunlight is considered a promising means because it is environmentally friendly and cost-effective. This review presented a summary of the recent progress in the removal of antibiotic pollutants using the photocatalytic oxidation process within the last five years. The review introduced the general characteristics and environmental dangers of common antibiotics, the basic mechanism of photocatalytic degradation, and the degradation of antibiotics by photocatalysts. Finally, opportunities and challenges in the photocatalytic degradation of antibiotics were discussed.
Asih et al. presented a review on the use of TiO2 photocatalysts for the degradation of dyes and drugs as organic pollutants in aqueous solutions [73]. It was reported that increasing the concentration of the TiO2 catalyst and the irradiation time increases the photodegradation of antibiotics.
Sodhi et al. published a review on the resistance, ecotoxicity, and remediation strategies used for AMX [74]. The increasing use of antibiotics, particularly AMX, has led to water contamination and a global medical crisis. AMX, a common antibiotic used in human and veterinary medicine, is refractory to degradation and cannot be completely removed from the environment. Wastewater treatment plants are struggling to efficiently remove AMX, which is present in drinking water and water bodies. Various methods, including physicochemical parameters, nanoparticles, phytoremediation, and the use of bacteria and algae, are being explored for better and sustainable technology.

4. Conclusions

This study presented a review of the published research for the elimination of antibiotic AMX by photodegradation using UV, visible, and solar irradiation in the presence of a variety of NMs. Most of the researchers have used TiO2 as the base material, doped and combined with other metals and metal oxides to achieve degradation using visible light. This review was made using research published during the last five years, from 2018–2024. Researchers have used a variety of reaction conditions that include radiation types (UV, solar, and visible), pH of the solution, the concentration of AMX and NMs, presence of other additives such as H2O2 as oxidants, and different salts like NaCl and CaCl2 to achieve high photodegradation efficiency. TiO2 was the best nanomaterial found that achieved the highest degradation of AMX in ultraviolet irradiation. TiO2 doped with other nanomaterials showed very good performance under visible light. WO3 was also used by several investigators and found quite effective for AMX degradation. Other metal oxides used for AMX elimination contain molybdenum, zinc, manganese, copper, cerium, and silver or a combination of these metals. Some researchers have used photolysis of AMX with UV, solar, or visible irradiation in the absence of solid catalysts with or without adding additives and oxidants such as H2O2.

Funding

The research was funded by the Deanship of Graduate Studies and Scientific Research, Jazan University, Saudi Arabia, through project number: (RG24-M037).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

The data presented in this study are available from the corresponding author upon reasonable request.

Acknowledgments

The authors gratefully acknowledge the funding of the Deanship of Graduate Studies and Scientific Research, Jazan University, Saudi Arabia, through project number: (RG24-M037).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Rogowska, J.; Cieszynska-Semenowicz, M.; Ratajczyk, W.; Wolska, L. Micropollutants in Treated Wastewater. Ambio 2020, 49, 487–503. [Google Scholar] [CrossRef]
  2. Javed, M.; Iqbal, S.; Qamar, M.A.; Shariq, M.; Ahmed, I.A.; BaQais, A.; Alzahrani, H.; Ali, S.K.; Masmali, N.A.; Althagafi, T.M.; et al. Fabrication of Effective Co-SnO2/SGCN Photocatalysts for the Removal of Organic Pollutants and Pathogen Inactivation. Crystals 2023, 13, 163. [Google Scholar] [CrossRef]
  3. Fallatah, A.M.; Alahmari, S.D.; Farid, H.M.T. Facile Synthesis of the MOF Derived ZnMn2O4 Nanorods for Dyes Degradation in Water. J. Mater. Sci. Mater. Electron. 2023, 34, 1630. [Google Scholar] [CrossRef]
  4. Ghaffar, S.; Abbas, A.; Naeem-ul-Hassan, M.; Assad, N.; Sher, M.; Ullah, S.; Alhazmi, H.A.; Najmi, A.; Zoghebi, K.; Al Bratty, M.; et al. Improved Photocatalytic and Antioxidant Activity of Olive Fruit Extract-Mediated ZnO Nanoparticles. Antioxidants 2023, 12, 1201. [Google Scholar] [CrossRef] [PubMed]
  5. Helmy, E.T.; Soliman, U.A.; Elbasiony, A.M.; Nguyen, B.-S. CuCe-Ferrite/TiO2 Nanocomposite as an Efficient Magnetically Separable Photocatalyst for Dye Pollutants Decolorization. Top. Catal. 2023, 66, 53–63. [Google Scholar] [CrossRef]
  6. Dilawar, S.; Albalawi, K.; Khan, A.U.; Tahir, K.; Zaki, M.E.A.; Musad Saleh, E.A.; Almarhoon, Z.M.; Althagafi, T.M.; El-Zahhar, A.A.; El-Bialy, E. Rapid Photodegradation of Toxic Organic Compounds and Photo Inhibition of Bacteria in the Presence of Novel Hydrothermally Synthesized Ag/Mn–ZnO Nanomaterial. Environ. Res. 2023, 231, 116093. [Google Scholar] [CrossRef]
  7. Qamar, M.A.; Shahid, S.; Javed, M.; Shariq, M.; Fadhali, M.M.; Madkhali, O.; Ali, S.K.; Syed, I.S.; Awaji, M.Y.; Shakir Khan, M.; et al. Accelerated Decoloration of Organic Dyes from Wastewater Using Ternary Metal/g-C3N4/ZnO Nanocomposites: An Investigation of Impact of g-C3N4 Concentration and Ni and Mn Doping. Catalysts 2022, 12, 1388. [Google Scholar] [CrossRef]
  8. Bessy, T.C.; Sarojoni, V.; El Manna, C.; Sasi, F.S.; Bakri, M.; Johnson, J.; Bindhu, M.R. Optical, Structural, Morphological, Antibacterial, and Photodegradation Characteristics of BaxMg0.8−xFe2O4 (x = 0.2, 0.4, and 0.6) Nanocrystalline Powders Synthesized by Combustion Method. Phys. Status Solidi (a) 2022, 219, 1–8. [Google Scholar] [CrossRef]
  9. Thilagavathi, T.; Venugopal, D.; Thangaraju, D.; Marnadu, R.; Palanivel, B.; Imran, M.; Shkir, M.; Ubaidullah, M.; AlFaify, S. A Facile Co-Precipitation Synthesis of Novel WO3/NiWO4 Nanocomposite with Improved Photocatalytic Activity. Mater. Sci. Semicond. Process. 2021, 133, 105970. [Google Scholar] [CrossRef]
  10. Husain Khan, A.; Abdul Aziz, H.; Khan, N.A.; Ahmed, S.; Mehtab, M.S.; Vambol, S.; Vambol, V.; Changani, F.; Islam, S. Pharmaceuticals of Emerging Concern in Hospital Wastewater: Removal of Ibuprofen and Ofloxacin Drugs Using MBBR Method. Int. J. Environ. Anal. Chem. 2023, 103, 140–154. [Google Scholar] [CrossRef]
  11. Shoeb, M.; Islam, R.; Parvin, N. Antibiotic Resistance. In Transcending Humanitarian Engineering Strategies for Sustainable Futures; IGI Global: Hershey, PA, USA, 2023; pp. 82–105. [Google Scholar]
  12. World Health Organization. Critically Important Antimicrobials for Human Medicine 6th Revision 2018 Ranking of Medically Important Antimicrobials for Risk Management of Antimicrobial Resistance Due to Non-Human Use; World Health Organization: Geneva, Switzerland, 2019. [Google Scholar]
  13. Alshandoudi, L.M.; Al Subhi, A.Y.; Al-Isaee, S.A.; Shaltout, W.A.; Hassan, A.F. Static Adsorption and Photocatalytic Degradation of Amoxicillin Using Titanium Dioxide/Hydroxyapatite Nanoparticles Based on Sea Scallop Shells. Environ. Sci. Pollut. Res. 2023, 30, 88704–88723. [Google Scholar] [CrossRef] [PubMed]
  14. 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]
  15. Mhemid, R.K.S.; Salman, M.S.; Mohammed, N.A. Comparing the Efficiency of N-Doped TiO2 and Commercial TiO2 as Photo Catalysts for Amoxicillin and Ciprofloxacin Photo-Degradation under Solar Irradiation. J. Environ. Sci. Health Part A 2022, 57, 813–829. [Google Scholar] [CrossRef] [PubMed]
  16. Ellepola, N.; Rubasinghege, G. Heterogeneous Photocatalysis of Amoxicillin under Natural Conditions and High-Intensity Light: Fate, Transformation, and Mineralogical Impacts. Environments 2022, 9, 77. [Google Scholar] [CrossRef]
  17. Wang, Y.; Zuo, G.; Kong, J.; Guo, Y.; Xian, Z.; Dai, Y.; Wang, J.; Gong, T.; Sun, C.; Xian, Q. Sheet-on-Sheet TiO2/Bi2MoO6 Heterostructure for Enhanced Photocatalytic Amoxicillin Degradation. J. Hazard. Mater. 2022, 421, 126634. [Google Scholar] [CrossRef]
  18. Zamani, S.; Rahimi, M.R.; Ghaedi, M. Spinning Disc Photoreactor Based Visible-Light-Driven Ag/Ag2O/TiO2 Heterojunction Photocatalyst Film toward the Degradation of Amoxicillin. J. Environ. Manag. 2022, 303, 114216. [Google Scholar] [CrossRef]
  19. Aissani, A.; Kameche, M.; Benabbou, K. Synthesis and Characterization of TiO2/LDH Layered Double Hydroxide Composites: Utilization as Photocatalysts for Amoxicillin Degradation under UVA Irradiation. Inorg. Nano-Met. Chem. 2022, 52, 1197–1207. [Google Scholar] [CrossRef]
  20. Hussien, M.S.A.; Yahia, I.S. Hybrid Multifunctional Core/Shell g-C3N4@TiO2 Heterojunction Nano-Catalytic for Photodegradation of Organic Dye and Pharmaceutical Compounds. Environ. Sci. Pollut. Res. 2021, 28, 29665–29680. [Google Scholar] [CrossRef] [PubMed]
  21. Gao, B.; Wang, J.; Dou, M.; Xu, C.; Huang, X. Enhanced Photocatalytic Removal of Amoxicillin with Ag/TiO2/Mesoporous g-C3N4 under Visible Light: Property and Mechanistic Studies. Environ. Sci. Pollut. Res. 2020, 27, 7025–7039. [Google Scholar] [CrossRef]
  22. Alkaim, A.; Aljobree, A. Adsorption and Photocatalytic Degradation of Pharmaceutical Amoxicillin Using TiO2 Nanoparticles in Aqueous Solutions: Oxidative Coupling as Spectrophotometric Method. Int. J. Adv. Sci. Technol. 2020, 29, 5480–5487. [Google Scholar]
  23. Salimi, M.; Behbahani, M.; Sobhi, H.R.; Gholami, M.; Jonidi Jafari, A.; Rezaei Kalantary, R.; Farzadkia, M.; Esrafili, A. A New Nano-Photocatalyst Based on Pt and Bi Co-Doped TiO2 for Efficient Visible-Light Photo Degradation of Amoxicillin. New J. Chem. 2019, 43, 1562–1568. [Google Scholar] [CrossRef]
  24. Yılmaz, Ç.H.; Akgeyik, E.; Bougarrani, S.; El Azzouzi, M.; Erdemoğlu, S. Photocatalytic Degradation of Amoxicillin Using Co-Doped TiO2 Synthesized by Reflux Method and Monitoring of Degradation Products by LC–MS/MS. J. Dispers. Sci. Technol. 2020, 41, 414–425. [Google Scholar] [CrossRef]
  25. Bergamonti, L.; Bergonzi, C.; Graiff, C.; Lottici, P.P.; Bettini, R.; Elviri, L. 3D Printed Chitosan Scaffolds: A New TiO2 Support for the Photocatalytic Degradation of Amoxicillin in Water. Water Res. 2019, 163, 114841. [Google Scholar] [CrossRef] [PubMed]
  26. Chinnaiyan, P.; Thampi, S.G.; Kumar, M.; Balachandran, M. Photocatalytic Degradation of Metformin and Amoxicillin in Synthetic Hospital Wastewater: Effect of Classical Parameters. Int. J. Environ. Sci. Technol. 2019, 16, 5463–5474. [Google Scholar] [CrossRef]
  27. Wahyuni, E.T.; Yulikayani, P.Y.; Aprilita, N.H. Enhancement of Visible-Light Photocatalytic Activity of Cu-Doped TiO2 for Photodegradation of Amoxicillin in Water. J. Mater. Environ. Sci. 2020, 11, 670–683. [Google Scholar]
  28. Fazilati, M. Photocatalytic Degradation of Amoxicillin, Cephalexin, and Tetracycline from Aqueous Solution: Comparison of Efficiency in the Usage of TiO2, ZnO, or GO-Fe3O4 Nanoparticles. Desalin. Water Treat. 2019, 169, 222–231. [Google Scholar] [CrossRef]
  29. Zhou, L.; Guo, X.; Lai, C.; Wang, W. Electro-Photocatalytic Degradation of Amoxicillin Using Calcium Titanate. Open Chem. 2018, 16, 949–955. [Google Scholar] [CrossRef]
  30. Verma, M.; Haritash, A.K. Photocatalytic Degradation of Amoxicillin in Pharmaceutical Wastewater: A Potential Tool to Manage Residual Antibiotics. Environ. Technol. Innov. 2020, 20, 101072. [Google Scholar] [CrossRef]
  31. 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]
  32. 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]
  33. Sharma, M.; Mandal, M.K.; Pandey, S.; Kumar, R.; Dubey, K.K. Visible-Light-Driven Photocatalytic Degradation of Tetracycline Using Heterostructured Cu2O–TiO2 Nanotubes, Kinetics, and Toxicity Evaluation of Degraded Products on Cell Lines. ACS Omega 2022, 7, 33572–33586. [Google Scholar] [CrossRef] [PubMed]
  34. Nema, S.; Sharma, A.; Rathore, V.K.; Chakraborty, M. Removal of Pharmaceuticals from Aqueous Solutions by Photodegradation Using TiO2 and Sn/Zn/Fe-Doped TiO2 as Photocatalyst Under Ultraviolet and Visible Light. Int. J. Environ. Res. 2024, 18, 12. [Google Scholar] [CrossRef]
  35. Gadore, V.; Mishra, S.R.; Ahmaruzzaman, M. Bio-Inspired Sustainable Synthesis of Novel SnS2/Biochar Nanocomposite for Adsorption Coupled Photodegradation of Amoxicillin and Congo Red: Effects of Reaction Parameters, and Water Matrices. J. Environ. Manag. 2023, 334, 117496. [Google Scholar] [CrossRef]
  36. Martins Bernardes Ramos, R.; Paludo, L.C.; Monteiro, P.I.; Maurat da Rocha, L.V.; Veiga de Moraes, C.; Santos, O.O.; Alves, E.R.; Porto Dantas, T.L. Amoxicillin Degradation by Iron Photonanocatalyst Synthetized by Green Route Using Pumpkin (Tetsukabuto) Peel Extract. Talanta 2023, 260, 124658. [Google Scholar] [CrossRef]
  37. Mmelesi, O.K.; Patala, R.; Nkambule, T.T.I.; Mamba, B.B.; Kefeni, K.K.; Kuvarega, A.T. Effect of Zn Doping on Physico-Chemical Properties of Cobalt Ferrite for the Photodegradation of Amoxicillin and Deactivation of E. coli. Colloids Surf. A Physicochem. Eng. Asp. 2022, 649, 129462. [Google Scholar] [CrossRef]
  38. Fawzy, A.; Mahanna, H.; Mossad, M. Effective Photocatalytic Degradation of Amoxicillin Using MIL-53(Al)/ZnO Composite. Environ. Sci. Pollut. Res. 2022, 29, 68532–68546. [Google Scholar] [CrossRef] [PubMed]
  39. Rodríguez-López, L.; Santás-Miguel, V.; Núñez-Delgado, A.; Álvarez-Rodríguez, E.; Pérez-Rodríguez, P.; Arias-Estévez, M. Influence of pH, Humic Acids, and Salts on the Dissipation of Amoxicillin and Azithromycin Under Simulated Sunlight. Span. J. Soil Sci. 2022, 12, 10438. [Google Scholar] [CrossRef]
  40. Utami, B.A.; Sutanto, H.; Alkian, I.; Sa’Adah, F.; Hidayanto, E. Efficient Degradation of Amoxicillin Using Bi2O3/Fe Synthesized by Microwave-Assisted Precipitation Method. Cogent Eng. 2022, 9, 2119534. [Google Scholar] [CrossRef]
  41. Le, S.; Zhu, C.; Cao, Y.; Wang, P.; Liu, Q.; Zhou, H.; Chen, C.; Wang, S.; Duan, X. V2O5 Nanodot-Decorated Laminar C3N4 for Sustainable Photodegradation of Amoxicillin under Solar Light. Appl. Catal. B 2022, 303, 120903. [Google Scholar] [CrossRef]
  42. Van Thuan, D.; Nguyen, T.L.; Pham Thi, H.H.; Thanh, N.T.; Ghotekar, S.; Sharma, A.K.; Binh, M.T.; Nga, T.T.; Pham, T.-D.; Cam, D.P. Development of Indium Vanadate and Silver Deposited on Graphitic Carbon Nitride Ternary Heterojunction for Advanced Photocatalytic Degradation of Residual Antibiotics in Aqueous Environment. Opt. Mater. 2022, 123, 111885. [Google Scholar] [CrossRef]
  43. Palas, B.; Ersöz, G. Photocatalytic activity of silver doped Self-Assembled supramolecular catalysts for amoxicillin Removal: Investigation of optimum conditions and reaction kinetics. J. Mol. Liq. 2024, 394, 123717. [Google Scholar] [CrossRef]
  44. Moradi, M.; Hasanvandian, F.; Isari, A.A.; Hayati, F.; Kakavandi, B.; Setayesh, S.R. CuO and ZnO Co-Anchored on g-C3N4 Nanosheets as an Affordable Double Z-Scheme Nanocomposite for Photocatalytic Decontamination of Amoxicillin. Appl. Catal. B 2021, 285, 119838. [Google Scholar] [CrossRef]
  45. Shi, W.; Liu, Y.; Sun, W.; Hong, Y.; Li, X.; Lin, X.; Guo, F.; Shi, J. Improvement of Synergistic Effect Photocatalytic/Peroxymonosulfate Activation for Degradation of Amoxicillin Using Carbon Dots Anchored on Rod-like CoFe2O4. Chin. J. Chem. Eng. 2022, 52, 136–145. [Google Scholar] [CrossRef]
  46. Xu, D.; Ma, Y.; Wang, J.; Chen, W.; Tang, Y.; Li, X.; Li, L. Interfacial Engineering of 2D/2D MXene Heterostructures: Face-to-Face Contact for Augmented Photodegradation of Amoxicillin. Chem. Eng. J. 2021, 426, 131246. [Google Scholar] [CrossRef]
  47. Le, S.; Ma, Y.; He, D.; Wang, X.; Guo, Y. CdS/NH4V4O10 S-Scheme Photocatalyst for Sustainable Photo-Decomposition of Amoxicillin. Chem. Eng. J. 2021, 426, 130354. [Google Scholar] [CrossRef]
  48. Le, S.; Yang, W.; Chen, G.; Yan, A.; Wang, X. Extensive Solar Light Harvesting by Integrating UPCL C-Dots with Sn2Ta2O7/SnO2: Highly Efficient Photocatalytic Degradation toward Amoxicillin. Environ. Pollut. 2020, 263, 114550. [Google Scholar] [CrossRef]
  49. Dou, M.; Wang, J.; Gao, B.; Xu, C.; Yang, F. Photocatalytic Difference of Amoxicillin and Cefotaxime under Visible Light by Mesoporous G-C3N4: Mechanism, Degradation Pathway and DFT Calculation. Chem. Eng. J. 2020, 383, 123134. [Google Scholar] [CrossRef]
  50. Samsudin, M.F.R.; Dumas, A.; Bashiri, R.; Mohamed, N.M.; Sufian, S. Development of the G-C3N4/BiVO4 Microflower Photocatalyst for Photocatalytic Degradation of Amoxicillin and Hydrogen Production. Malays. J. Microsc. 2020, 16, 180–187. [Google Scholar]
  51. 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]
  52. Chahkandi, M.; Zargazi, M. New Water Based EPD Thin BiVO4 Film: Effective Photocatalytic Degradation of Amoxicillin Antibiotic. J. Hazard. Mater. 2020, 389, 121850. [Google Scholar] [CrossRef]
  53. Abazari, R.; Mahjoub, A.R. Amine-Functionalized Al-MOF#@yxSm2O3–ZnO: A Visible Light-Driven Nanocomposite with Excellent Photocatalytic Activity for the Photo-Degradation of Amoxicillin. Inorg. Chem. 2018, 57, 2529–2545. [Google Scholar] [CrossRef]
  54. Gaim, Y.T.; Yimanuh, S.M.; Kidanu, Z.G. Enhanced Photocatalytic Degradation of Amoxicillin with Mn-Doped Cu2O under Sunlight Irradiation. J. Compos. Sci. 2022, 6, 317. [Google Scholar] [CrossRef]
  55. Homem, V.; Alves, A.; Santos, L. Amoxicillin Degradation at Ppb Levels by Fenton’s Oxidation Using Design of Experiments. Sci. Total Environ. 2010, 408, 6272–6280. [Google Scholar] [CrossRef]
  56. Turkay, G.K.; Kumbur, H. Investigation of Amoxicillin Removal from Aqueous Solution by Fenton and Photocatalytic Oxidation Processes. Kuwait J. Sci. 2019, 46, 85–93. [Google Scholar]
  57. Haddadou, N.; Bensemma, N.; Rekhila, G.; Trari, M.; Taïbi, K. Photoelectrochemical Investigations in Lead-Free Ba (Ti0.950Sc0.025Nb0.025)O3 Ferroelectric Ceramics. Application to Amoxicillin Photodegradation. J. Photochem. Photobiol. A Chem. 2018, 358, 294–299. [Google Scholar] [CrossRef]
  58. Kattel, E.; Kaur, B.; Trapido, M.; Dulova, N. Persulfate-Based Photodegradation of a Beta-Lactam Antibiotic Amoxicillin in Various Water Matrices. Environ. Technol. 2020, 41, 202–210. [Google Scholar] [CrossRef]
  59. Beshkar, F.; Al-Nayili, A.; Amiri, O.; Salavati-Niasari, M.; Mousavi-Kamazani, M. Visible Light-Induced Degradation of Amoxicillin Antibiotic by Novel CuI/FePO4 p-n Heterojunction Photocatalyst and Photodegradation Mechanism. J. Alloys Compd. 2022, 892, 162176. [Google Scholar] [CrossRef]
  60. Silva, I.F.; Teixeira, I.F.; Rios, R.D.F.; do Nascimento, G.M.; Binatti, I.; Victória, H.F.V.; Krambrock, K.; Cury, L.A.; Teixeira, A.P.C.; Stumpf, H.O. Amoxicillin Photodegradation under Visible Light Catalyzed by Metal-Free Carbon Nitride: An Investigation of the Influence of the Structural Defects. J. Hazard. Mater. 2021, 401, 123713. [Google Scholar] [CrossRef]
  61. 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 2019, 242, 337–348. [Google Scholar] [CrossRef]
  62. Al-Musawi, T.J.; Alghamdi, M.I.; Alhachami, F.R.; Zaidan, H.; Mengelizadeh, N.; Asghar, A.; Balarak, D. The application of a new recyclable photocatalyst γ-Fe2O3@SiO2@ZIF8-Ag in the photocatalytic degradation of amoxicillin in aqueous solutions. Environ. Monit. Assess. 2023, 195, 372. [Google Scholar] [CrossRef]
  63. Tarannum, S.; Hossain, M.S.; Bashar, M.S.; Bahadur, N.M.; Ahmed, S. Amplification of photocatalytic degradation of antibiotics (amoxicillin, ciprofloxacin) by sodium doping in nano-crystallite hydroxyapatite. RSC Adv. 2024, 14, 12386–12396. [Google Scholar] [CrossRef]
  64. Vettorello, G.; Kuhn, D.; Dalcorso, A.B.; Weber, A.C.; Costa, B.; Cordeiro, S.G.; Schweizer, Y.A.; de Freitas, E.M.; Rieger, A.; Steffens, C.; et al. Degradation of the micropollutant amoxicillin by photolysis and evaluation of the toxicity of byproducts generated in aquatic environments. Sustain. Water Resour. Manag. 2022, 8, 21. [Google Scholar] [CrossRef]
  65. Zhang, Y.; Xiao, Y.; Zhong, Y.; Lim, T.-T. Comparison of amoxicillin photodegradation in the UV/H2O2 and UV/persulfate systems: Reaction kinetics, degradation pathways, and antibacterial activity. Chem. Eng. J. 2019, 372, 420–428. [Google Scholar] [CrossRef]
  66. Arsand, J.B.; Hoff, R.B.; Jank, L.; Meirelles, L.N.; Díaz-Cruz, M.S.; Pizzolato, T.M.; Barceló, D. Transformation products of amoxicillin and ampicillin after photolysis in aqueous matrices: Identification and kinetics. Sci. Total Environ. 2018, 642, 954–967. [Google Scholar] [CrossRef]
  67. Timm, A.; Borowska, E.; Majewsky, M.; Merel, S.; Zwiener, C.; Bräse, S.; Horn, H. Photolysis of four β-lactam antibiotics under simulated environmental conditions: Degradation, transformation products and antibacterial activity. Sci. Total Environ. 2019, 651, 1605–1612. [Google Scholar] [CrossRef]
  68. Pirsaheb, M.; Hossaini, H.; Raad, N.K.; Kianpour, S.; Hossini, H. A systematic review on photo-Fenton process as an efficient advanced oxidation for degradation of amoxicillin in aqueous environments. Rev. Environ. Health 2023, 38, 313–326. [Google Scholar] [CrossRef]
  69. Qutob, M.; Shakeel, F.; Alam, P.; Alshehri, S.; Ghoneim, M.M.; Rafatullah, M. A review of radical and non-radical degradation of amoxicillin by using different oxidation process systems. Environ. Res. 2022, 214 Pt 1, 113833. [Google Scholar] [CrossRef]
  70. Aryee, A.A.; Han, R.; Qu, L. Occurrence, detection and removal of amoxicillin in wastewater: A review. J. Clean. Prod. 2022, 368, 133140. [Google Scholar] [CrossRef]
  71. Manikanika; Chopra, L. Photocatalytic Activity of Zinc Oxide for Dye and Drug Degradation: A Review. Mater. Today Proc. 2022, 52, 1653–1656. [Google Scholar] [CrossRef]
  72. Chen, Y.; Yang, J.; Zeng, L.; Zhu, M. Recent Progress on the Removal of Antibiotic Pollutants Using Photocatalytic Oxidation Process. Crit. Rev. Environ. Sci. Technol. 2022, 52, 1401–1448. [Google Scholar] [CrossRef]
  73. Asih, A.K.; Yetti, R.D.; Chandra, B. Photodegradation of Antibiotic Using TiO2 as a Catalyst: A Review. Int. J. Pharm. Sci. Med. 2021, 6, 37–43. [Google Scholar] [CrossRef]
  74. Sodhi, K.K.; Kumar, M.; Singh, D.K. Insight into the amoxicillin resistance, ecotoxicity, and remediation strategies. J. Water Process Eng. 2021, 39, 101858. [Google Scholar] [CrossRef]
Figure 1. A comparison of AMX degradation under different experimental conditions. (A) Degradation plots of AMX with TiO2 anatase and kaolinite under light and dark conditions. (B) pH variation of the AMX as a function of reaction time [16].
Figure 1. A comparison of AMX degradation under different experimental conditions. (A) Degradation plots of AMX with TiO2 anatase and kaolinite under light and dark conditions. (B) pH variation of the AMX as a function of reaction time [16].
Ijms 25 09575 g001
Figure 2. Interaction of AMX with the mineral surface and degradation into simpler products [16].
Figure 2. Interaction of AMX with the mineral surface and degradation into simpler products [16].
Ijms 25 09575 g002
Figure 3. Proposed degradation reactions of AMX under dark and solar radiation without minerals [16].
Figure 3. Proposed degradation reactions of AMX under dark and solar radiation without minerals [16].
Ijms 25 09575 g003
Figure 4. Visualization of 2D/2D TiO2/Bi2MoO6 catalyst and AMX photodegradation process [17].
Figure 4. Visualization of 2D/2D TiO2/Bi2MoO6 catalyst and AMX photodegradation process [17].
Ijms 25 09575 g004
Figure 5. Photocatalytic degradation of AMX at different initial concentrations of 10 to 50 ppm using 0.2 g/L of catalyst at 25 °C [22].
Figure 5. Photocatalytic degradation of AMX at different initial concentrations of 10 to 50 ppm using 0.2 g/L of catalyst at 25 °C [22].
Ijms 25 09575 g005
Figure 6. AMX photodegradation efficiency under UV and visible light. (A) No photocatalyst, (B) TiO2, and (C) Cu-TiO2. (AMX solution volume 100 mL, AMX concentration 10 mg/L, photocatalysts 40 mg, irradiation time 24 h, and pH 6) [27].
Figure 6. AMX photodegradation efficiency under UV and visible light. (A) No photocatalyst, (B) TiO2, and (C) Cu-TiO2. (AMX solution volume 100 mL, AMX concentration 10 mg/L, photocatalysts 40 mg, irradiation time 24 h, and pH 6) [27].
Ijms 25 09575 g006
Figure 7. Effect of pH on the degradation efficiency of AMX and cephalexin (CPX) under UV irradiation using ZnO, TiO2, and GO-Fe3O4 catalyst. Irradiation time 20 min, catalyst 1 g/L, antibiotics 5 mg/L, UV radiation intensity 6 W [28].
Figure 7. Effect of pH on the degradation efficiency of AMX and cephalexin (CPX) under UV irradiation using ZnO, TiO2, and GO-Fe3O4 catalyst. Irradiation time 20 min, catalyst 1 g/L, antibiotics 5 mg/L, UV radiation intensity 6 W [28].
Ijms 25 09575 g007
Figure 8. Effect of radiation intensity on the degradation efficiency of AMX, cephalexin (CPX), and tetracycline (TCN) under UV irradiation using ZnO, TiO2, and GO-Fe3O4 catalysts: pH (AMX, CPX, TCN with GO-Fe3O4) unset; pH (AMX, CPX, TCN with TiO2 or ZnO) 5, 7, 5; time (AMX, CPX, and TCN/MGO) 45, 45, 15 min; time (AMX, CPX, and TCN/TiO2 or ZnO) 15, 45, 45 min; MGO quantity (with AMX, CPX, or TCN) 4, 4, 2 g/L; TiO2 or ZnO quantity (with AMX, CPX, or TCN) 2, 2, 2 g/L; antibiotics concentration 15 mg/L [28].
Figure 8. Effect of radiation intensity on the degradation efficiency of AMX, cephalexin (CPX), and tetracycline (TCN) under UV irradiation using ZnO, TiO2, and GO-Fe3O4 catalysts: pH (AMX, CPX, TCN with GO-Fe3O4) unset; pH (AMX, CPX, TCN with TiO2 or ZnO) 5, 7, 5; time (AMX, CPX, and TCN/MGO) 45, 45, 15 min; time (AMX, CPX, and TCN/TiO2 or ZnO) 15, 45, 45 min; MGO quantity (with AMX, CPX, or TCN) 4, 4, 2 g/L; TiO2 or ZnO quantity (with AMX, CPX, or TCN) 2, 2, 2 g/L; antibiotics concentration 15 mg/L [28].
Ijms 25 09575 g008
Figure 9. Effect of initial AMX concentrations on the degradation efficiency for 200 mL solution containing 100 mg/L and 300 mg/L AMX with 0.5 g CaTiO3 at room temperature at different irradiation times [29].
Figure 9. Effect of initial AMX concentrations on the degradation efficiency for 200 mL solution containing 100 mg/L and 300 mg/L AMX with 0.5 g CaTiO3 at room temperature at different irradiation times [29].
Ijms 25 09575 g009
Figure 10. AMX degradation efficiency of 200 mL solution containing 100 mg/L AMX with 0.5 g CaTiO3 at room temperature at different irradiation times at pH 3, 7, and 10 [29].
Figure 10. AMX degradation efficiency of 200 mL solution containing 100 mg/L AMX with 0.5 g CaTiO3 at room temperature at different irradiation times at pH 3, 7, and 10 [29].
Ijms 25 09575 g010
Figure 11. Effect of CaTiO3 amount on the AMX degradation efficiency for 200 mL aqueous solution having 100 mg/L AMX at pH 3 and room temperature [29].
Figure 11. Effect of CaTiO3 amount on the AMX degradation efficiency for 200 mL aqueous solution having 100 mg/L AMX at pH 3 and room temperature [29].
Ijms 25 09575 g011
Figure 12. Effect of temperature on the AMX degradation efficiency for 200 mL solution having 100 mg/L AMX with 0.5 g CaTiO3 and 0.058 g NaCl at pH 3 [29].
Figure 12. Effect of temperature on the AMX degradation efficiency for 200 mL solution having 100 mg/L AMX with 0.5 g CaTiO3 and 0.058 g NaCl at pH 3 [29].
Ijms 25 09575 g012
Figure 13. Process of AMX degradation using TiO2 under UV irradiation (UVA 365 nm, 672 W/m2) for 30 mg/L AMX concentration, TiO2 dosage of 450 mg/L, H2O2 concentration of 150 mg/L at pH 7.0 [30].
Figure 13. Process of AMX degradation using TiO2 under UV irradiation (UVA 365 nm, 672 W/m2) for 30 mg/L AMX concentration, TiO2 dosage of 450 mg/L, H2O2 concentration of 150 mg/L at pH 7.0 [30].
Ijms 25 09575 g013
Figure 14. (a) Possible mechanism of AMX degradation at the GO/TiO2 surface [31]. Photocatalytic degradation kinetics of AMX by the synthesized materials under (b) visible light and (c) simulated solar light. (d) AMX degradation rate constants under solar and visible light [32].
Figure 14. (a) Possible mechanism of AMX degradation at the GO/TiO2 surface [31]. Photocatalytic degradation kinetics of AMX by the synthesized materials under (b) visible light and (c) simulated solar light. (d) AMX degradation rate constants under solar and visible light [32].
Ijms 25 09575 g014
Figure 15. SEM images of photocatalysts (a) MIL-53(Al)/ZnO, (b) MIL-53Al, and (c) ZnO [38].
Figure 15. SEM images of photocatalysts (a) MIL-53(Al)/ZnO, (b) MIL-53Al, and (c) ZnO [38].
Ijms 25 09575 g015
Figure 16. Photodegradation results of AMX using different catalysts, using 50 mg/L of AMX and 0.6 g/L of catalyst at pH 7.0 [38].
Figure 16. Photodegradation results of AMX using different catalysts, using 50 mg/L of AMX and 0.6 g/L of catalyst at pH 7.0 [38].
Ijms 25 09575 g016
Figure 17. The charge transfer process between ZnO and MIL-53Al and the degradation routes of AMX [38]. AMX molecule has Red: Oxygen, Blue: Nitrogen, Yellow: Sulfur, White: Hydrogen, Black: Carbon.
Figure 17. The charge transfer process between ZnO and MIL-53Al and the degradation routes of AMX [38]. AMX molecule has Red: Oxygen, Blue: Nitrogen, Yellow: Sulfur, White: Hydrogen, Black: Carbon.
Ijms 25 09575 g017
Figure 18. Recyclability results of AMX in five runs using 0.6 g/L of MIL-53(Al)/ZnO catalyst at pH 7.0, 50 mg/L of AMX, and irradiation time of 60 min [38].
Figure 18. Recyclability results of AMX in five runs using 0.6 g/L of MIL-53(Al)/ZnO catalyst at pH 7.0, 50 mg/L of AMX, and irradiation time of 60 min [38].
Ijms 25 09575 g018
Figure 19. Kinetics plots of AMX degradation at (A) pH 4.0, (B) pH 5.5, and (C) pH 7.0 [39].
Figure 19. Kinetics plots of AMX degradation at (A) pH 4.0, (B) pH 5.5, and (C) pH 7.0 [39].
Ijms 25 09575 g019
Figure 20. AMX degradation under simulated sunlight and in the dark using inorganic salts: (A) NaCl; (B) CaCl2; (C) NaNO3; (D) Na2HPO4; (E) FeCl3; and (F) humic acids [39].
Figure 20. AMX degradation under simulated sunlight and in the dark using inorganic salts: (A) NaCl; (B) CaCl2; (C) NaNO3; (D) Na2HPO4; (E) FeCl3; and (F) humic acids [39].
Ijms 25 09575 g020
Figure 21. SEM images of (a) Cu2O and (b) Mn-doped Cu2O [54].
Figure 21. SEM images of (a) Cu2O and (b) Mn-doped Cu2O [54].
Ijms 25 09575 g021
Figure 22. Effect of pH on photocatalytic degradation of 15 mg/L AMX using 1 g/L of Mn-doped Cu2O catalyst [54].
Figure 22. Effect of pH on photocatalytic degradation of 15 mg/L AMX using 1 g/L of Mn-doped Cu2O catalyst [54].
Ijms 25 09575 g022
Figure 23. Effect of initial concentration of AMX on photocatalytic degradation using 1 g/L Mn-doped Cu2O catalyst at pH 9 [54].
Figure 23. Effect of initial concentration of AMX on photocatalytic degradation using 1 g/L Mn-doped Cu2O catalyst at pH 9 [54].
Ijms 25 09575 g023
Figure 24. Effect of Mn-doped Cu2O catalyst dose on photocatalytic degradation of 15 mg/L AMX at pH 8 [54].
Figure 24. Effect of Mn-doped Cu2O catalyst dose on photocatalytic degradation of 15 mg/L AMX at pH 8 [54].
Ijms 25 09575 g024
Figure 25. Photodegradation of AMX using Cu2O and Mn-doped Cu2O under the optimal conditions of 15 mg/L of AMX and 1 g/L of catalyst at pH 9 [54].
Figure 25. Photodegradation of AMX using Cu2O and Mn-doped Cu2O under the optimal conditions of 15 mg/L of AMX and 1 g/L of catalyst at pH 9 [54].
Ijms 25 09575 g025
Figure 26. Photodegradation of AMX using (a) Ba(Ti0.950Sc0.025Nb0.025)O3 and (b) the spectra showing degradation with irradiation time [57].
Figure 26. Photodegradation of AMX using (a) Ba(Ti0.950Sc0.025Nb0.025)O3 and (b) the spectra showing degradation with irradiation time [57].
Ijms 25 09575 g026
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ali, M.A.; Maafa, I.M. Photodegradation of Amoxicillin in Aqueous Systems: A Review. Int. J. Mol. Sci. 2024, 25, 9575. https://doi.org/10.3390/ijms25179575

AMA Style

Ali MA, Maafa IM. Photodegradation of Amoxicillin in Aqueous Systems: A Review. International Journal of Molecular Sciences. 2024; 25(17):9575. https://doi.org/10.3390/ijms25179575

Chicago/Turabian Style

Ali, Mohammad Ashraf, and Ibrahim M. Maafa. 2024. "Photodegradation of Amoxicillin in Aqueous Systems: A Review" International Journal of Molecular Sciences 25, no. 17: 9575. https://doi.org/10.3390/ijms25179575

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