Next Article in Journal
Production, Recycling and Economy of Palladium: A Critical Review
Previous Article in Journal
High-Performance Electrochromic Devices Based on Size-Controlled 2D WO3 Nanosheets Prepared Using the Intercalation Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Research Progress on Inactivation of Bacteriophages by Visible-Light Photocatalytic Composite Materials: A Mini Review

by
Deqiang Zhao
1,2,3,*,†,
Heng Lu
1,2,†,
Qingkong Cheng
1,2,4,*,
Qi Huang
1,2,
Jing Ai
3,
Zhibo Zhang
3,
Hainan Liu
1,2,
Zongfei He
5 and
Qiuhong Li
1,2
1
School of River and Ocean Engineering, Chongqing Jiaotong University, Chongqing 400074, China
2
National Engineering Research Center for Inland Waterway Regulation, Key Laboratory of Hydraulic and Waterway Engineering of the Ministry of Education, Chongqing 400074, China
3
Department of Materials and Environmental Chemistry, Stockholm University, SE-106 91 Stockholm, Sweden
4
Joint Graduate Training Base for Resources and Environment between Chongqing Jiaotong University and Chongqing Gangli Environmental Protection Co., Ltd., Chongqing Jiaotong University, Chongqing 400074, China
5
School of Environmental and Municipal Engineering, Lanzhou Jiaotong University, Lanzhou 730070, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Materials 2024, 17(1), 44; https://doi.org/10.3390/ma17010044
Submission received: 16 November 2023 / Revised: 16 December 2023 / Accepted: 19 December 2023 / Published: 21 December 2023

Abstract

:
Infectious diseases caused by waterborne viruses have attracted researchers’ great attention. To ensure a safe water environment, it is important to advance water treatment and disinfection technology. Photocatalytic technology offers an efficient and practical approach for achieving this goal. This paper reviews the latest studies on visible-light composite catalysts for bacteriophage inactivation, with a main focus on three distinct categories: modified UV materials, direct visible-light materials and carbon-based materials. This review gives an insight into the progress in photocatalytic material development and offers a promising solution for bacteriophage inactivation.

1. Introduction

Waterborne viruses in drinking water and infectious diseases caused by waterborne viruses have gradually attracted people’s attention [1]. In order to ensure the safety of drinking water, it is of great significance to develop an efficient and low-cost water treatment and disinfection method. Phages are commonly found in sewage and polluted water, and their morphology and structure are similar to those of enteroviruses [2]. Single-stranded RNA phages are thus expected to replace some enteroviruses (such as poliovirus PV-1) as virus indicators for sewage evaluation. With Escherichia coli serving as the specific host, phages are safe as virus indicators. Additionally, phages possess clear nucleic acid sequences and simple structures, and provide for a short experimental period, ease of operation, and low cost with objective evaluation results, aligning with international standard methods suitable for assessing and testing water quality. As a result, bacteriophage MS2 has been recommended by the US EPA as an indicator organism for viruses in water and is now widely used for evaluating water and sewage disinfection efficiency and understanding the mechanism of virus inactivation [3].
Chemical disinfectants commonly used in traditional water treatment include chlorine, chlorine dioxide and ozone. However, studies have found that the use of chemical disinfection processes will form a number of harmful compounds (such as formaldehyde and acetaldehyde produced by using ozone) and toxic disinfection byproducts (DBPs), seriously threatening human health [4,5]. Ultraviolet light (UV) disinfection and filtration technology are also widely used water treatment technologies. Compared with traditional chemical disinfectants, ultraviolet disinfection is more effective without the production of DBPs. However, the biggest disadvantage of ultraviolet disinfection is the lack of continuous disinfection ability, and thus the ultraviolet method is only suitable for primary disinfection [6]. Filtration technology is used to separate pollutants in water using physical barriers, effectively removing particles, microorganisms and pollutants in water. However, filtration technology has high operating costs, restricting its wide application [7]. Given the virus’ small size, unique structure and high pathogenicity, it remains a challenging task to effectively eliminate waterborne viruses using current disinfection technology. Therefore, an efficient and environmentally friendly alternative disinfection method is in demand.
As a new advanced oxidation technology, photocatalysis has attracted extensive attention. Photocatalysis is a versatile technology used to treat heavy metals [8], eliminate organic pollutants [9,10,11,12], and reduce nitrates [13]. Via the photoexcitation of electrons on catalyst surfaces, it facilitates the reduction or oxidation reactions of heavy metal ions, transforming them into stable or precipitated forms, thereby diminishing the toxicity of heavy metals in water or soil. By leveraging the photosensitivity of catalysts to generate active oxides, organic substances undergo oxidation and degradation into water and carbon dioxide, mitigating the organic load in water or soil [14,15,16,17]. In contrast to conventional disinfection methods, photocatalytic disinfection showcases extensive bactericidal effectiveness [18], simultaneously addressing a variety of microorganisms such as bacteria, viruses, and fungi. The notable advantage lies in the lack of chemical residues after treatment, minimizing adverse impacts on both the environment and human health [19]. Moreover, the sustainability of the photocatalytic disinfection process is noteworthy, as it only requires a light source to activate the catalyst, eliminating the necessity for frequent chemical replacements or additions. It uses solar energy directly to drive chemical reactions, facilitating the decomposition of organic substances, thus surpassing the limitations of traditional disinfection methods.
In the photocatalytic process, when light irradiates the photocatalyst with the energy of the radiation equal to or greater than the electron band gap, the surface of the photocatalyst generates electron-hole (e-h+) pairs, and those pairs then separate and migrate to the photocatalyst surface to participate in redox reactions [20]. For instance, photogenerated e reduces O2 in water to •O2, while the photogenerated h+ oxidizes the surface H2O/OH to •OH. Both •O2 and •OH are reactive oxygen species (ROSs), which are highly active oxidants that play an important role in the process of the photocatalytic inactivation of viruses [5].
TiO2 is one of the most common photocatalysts, used as a fungicide for the first time in 1985, successfully inactivating photocatalytic Saccharomyces cerevisiae under UV irradiation [21]. TiO2 inactivates microorganisms by absorbing UV to produce ROSs, and it is widely used in water disinfection due to its low cost, high disinfection efficiency, non-toxicity and good light stability [22,23]. In addition to TiO2, many materials have also been used in the photocatalytic degradation of refractory organics, water disinfection and other fields, such as ZnO, WO3, and g-C3N4 [24,25,26]. In recent years, some studies have reported good performance of photocatalyst materials for the removal of Salmonella, Pseudomonas aeruginosa and viruses such as MS2 phage and ΩX174 phage [27,28]. In spite of the success of UV photocatalytic disinfection [29,30,31,32], solar-induced photocatalytic water disinfection still faces challenges. At present, most semiconductor materials can only achieve light response in the ultraviolet region, which only accounts for 4–5% of the entire solar spectrum, while the visible light accounts for about 44%. Accordingly, how to effectively utilize visible light in photocatalytic disinfection is an urgent problem to be solved, and the development of visible-light-active photocatalysts has become a research hotspot in the development of photocatalytic disinfection.
With continuous research and development, many visible-light photocatalytic composite materials have been developed using metal doping and semiconductor composite methods. Photocatalytic activity may be enhanced by doping TiO2 with non-metals or/and noble and transition metals, which significantly influence the photo-reactivity, charge carrier recombination rates and interfacial electron-transfer rates. These composite catalysts have been shown to have superior antiviral and antibacterial behavior compared to pure titania [33]. They also showcase significant potential in hydrolyzing and degrading organic pollutants [34,35,36]. Despite these advancements, the practical application of photocatalytic water disinfection technology faces challenges associated with the immobilization of photocatalysts and the design of photocatalytic reactors, mainly attributed to the prevalent use of powder-form photocatalysts [5]. Future research endeavors will be directed towards the development of more efficient and easily recyclable photocatalysts, including metal–organic frameworks (MOFs) with semiconductor-like properties [37] and photocatalyst with magnetic properties [38]. These advanced photocatalysts possess notable advantages, including the presence of open metal sites and catalytically active ligands, which empower them to act as charge carriers upon photoexcitation. Their exceptional design flexibility arises from the ease of modifying metal centers or organic ligands, distinguishing them significantly from conventional semiconductors [39].
This review aims to introduce the latest research progress of visible-light composite photocatalysts for bacteriophage inactivation and focuses on the synthesis and modification of visible-light catalysts. The bacteriophage inactivation mechanism and efficiency of several visible-light photocatalysts are also reviewed with perspectives on the development of visible-light photocatalytic disinfection.

2. Modified Ultraviolet-Response Composite Materials

The most widely used photocatalytic materials including ZnO and TiO2 are wide band gap semiconductors that can only be activated in the ultraviolet region and cannot directly absorb visible light. To achieve photo-responsiveness in the visible region, photocatalytic materials have been modified and developed using different methods.

2.1. Cu-TiO2

In the photocatalytic process, a major limitation of TiO2 is its large band gap, which can only be activated by light in the near-ultraviolet region. Also, the recombination of photogenerated charge carriers of TiO2 reduces the overall quantum efficiency. The above problems can be solved by metal doping (such as Cu and Ag). Zheng et al. [40] synthesized Cu-TiO2 nanofiber photocatalysts using electrospinning, which confirmed efficient disinfection performance for bacteriophage f2 under visible-light irradiation and this performance was not affected by reaction pH. The removal efficiency of phage f2 by Cu-TiO2 remained high even when the pH was close to the point of zero charge (PZC) (pH = 6, 7 was neutral). Under visible-light irradiation, the removal efficiency increased with the increase in visible-light intensity, catalyst dosage and temperature, but decreased with increasing initial virus concentration. With a temperature of 25 °C, pH = 7 and a concentration of 75 mg/L Cu-TiO2 nanofibers, the virus with an initial concentration of 105 PFU/mL was completely inactivated after 240 min of visible-light irradiation, showing the best disinfection performance (Figure 1a). Irie et al. [41] created a Cu(II)-TiO2 visible-light photocatalytic material by grafting Cu(II) nanoclusters onto TiO2. As can be seen from Figure 1b, visible-light absorption improved in the 410 to 450 nm rangeand over 650 nm for Cu(II)-TiO2. The absorption of 410–450 nm is due to the fact that the e in the valence band of TiO2 are excited into Cu(II) under visible-light irradiation via an interfacial charge transfer process [42], while the absorption at over 650 nm is ascribed to the d-d transition of Cu(II). Liu et al. [43] optimized the synthesis of Cu(II)-TiO2 nanocomposites and prepared thin film samples by coating a Cu(II)-TiO2 suspension onto glass substrates. As shown in Figure 1c,d, the highest photocatalytic activity of the Cu(II)-TiO2 film showed at a pH of 12. After 30 min of visible-light irradiation, the bacteriophage decreased by 6.5log (Figure 1d), and the virus inactivation reached 99% and 99.99% after 1 h and 2 h of visible-light irradiation, respectively. The experimental data show that Cu(II)-TiO2 has efficient inactivation efficiency on bacteriophage Qβ under visible-light irradiation.

2.2. CuxO-TiO2

Copper oxide nanoclusters (CuxO) are a p-type semiconductor with high visible-light absorption and remarkable electrical properties [44]. Miyauchi et al. [45] produced CuxO (1 < x < 2) nanoclusters by combining Cu(I) and Cu(II) into and grafting them onto the surface of TiO2 to obtain a CuxO/TiO2 photocatalyst, which presented high inactivation activity against bacteriophages under visible-light irradiation. The material with a ratio of Cu(I)/Cu(II) of 1.3 showed the best performance in maintaining effective visible-light photocatalytic activity and antiviral activity. As shown in Figure 2, the antiviral activity of CuxO/TiO2 under visible light was further improved compared to that under dark conditions, and bacteriophage Qβ was reduced by 7.5log in 40 min. This is because Cu(I) has good antiviral activity under dark conditions, while Cu(II) can accept e from the valence band of TiO2 to form Cu(I) by photo-induced interfacial charge transfer transition. Due to the presence of both Cu(I) and h+ in the valence band of TiO2, CuxO/TiO2 exhibits more efficient antiviral performance under visible-light irradiation.

2.3. TiON/PdO

It has been reported that the absorption spectrum of TiO2 can be extended to the visible-light region by doping nitrogen [46]. Li et al. [47] synthesized a palladium-modified nitrogen-doped titanium dioxide (TiON/PdO) nanoparticle photocatalyst using a sol-gel process, and the composite was further fixed onto a mesoporous activated carbon fiber template for easy recovery of the photocatalyst from water. As is shown in Figure 3, the modification with PdO enhanced the light absorption in the visible range (>400 nm), and the absorption increased with increasing PdO dose when PdO < 2% [48]. The removal rate of bacteriophage MS2 by TiON/PdO reached 94.5–98.2% after 1 h of visible-light irradiation, which was significantly higher than that under dark conditions [47].

2.4. ZnO-Based Floating Photocatalyst

ZnO is an environmentally friendly material with a band gap (about 3.2 eV) similar to that of TiO2 and similar photocatalytic properties which need to be irradiated by UV to show its activity. Because ZnO is relatively cheap, it is often considered as a substitute for TiO2 in photocatalytic applications [49]. Studies have shown that doping Ni can improve the photocatalytic performance of ZnO [50,51]. In order to further enhance the photocatalytic efficiency, a floating composite catalyst has been developed by immobilizing the photocatalyst on a floating substrate to utilize the maximum light source and retain the pollutant near the surface of the photocatalyst. The floating photocatalyst can achieve the maximum oxidation and the best light radiation, greatly improving the generation of free radicals and photocatalytic efficiency [52]. Urbonavicius et al. [53] prepared ZnO-based floating photocatalysts with high-density polyethylene (HDPE) spheres with and without a Ni underlayer as the floatable substrate, respectively, and the ZnO film was deposited onto the HDPE sphere (denoted as ZnO-HDPE). It was found that the use of the ZnO-HDPE photocatalyst could inactivate the phage mixture under visible-light irradiation. Without a Ni underlayer, ZnO-HDPE reduced infectivity by 38% and 47% after irradiation of phage T4 alone and in the mixture with PRD1 (T1 + PRD1) for 1 h, respectively (Figure 4a), while ZnO-HDPE with a Ni underlayer reduced phage infectivity by 18% and nearly 50%, respectively (Figure 4b). Although there was no significant difference in the inactivation effect of the ZnO-HDPE photocatalyst with or without the Ni underlayer on phages, the sensitivity of phage T4 was greatly increased after mixing with phage PRD1.

3. Direct Visible Light Response Composite Materials

Graphitic carbon nitride (g-C3N4) is a π-conjugated semiconductor that can absorb visible light up to 460 nm. It is considered to be a novel material as a visible-light-responsive photocatalyst due to its excellent visible-light response, high chemical stability and non-metal property. However, the band gap (2.7 eV) of g-C3N4 is relatively large and the recombination rate of photo-excited electrons (e) and holes (h+) is very high, which causes the low effective utilization of visible light [54]. Several approaches have been developed to enhance the photocatalytic activity of g-C3N4.

3.1. Graphitic Carbon Nitride (g-C3N4)

Zhang et al. [55] investigated the photocatalytic inactivation of bacteriophage MS2 using a g-C3N4 photocatalyst by changing the operating parameters, namely light intensity, photocatalyst loading and reaction temperature via a response surface methodology. The study revealed that the order of influence of variables on the light response of g-C3N4 followed light intensity > photocatalyst loading > reaction temperature. The best performance of photocatalytic virus inactivation was shown under the conditions of light intensity of 199.80 mW/cm2, photocatalyst loading of 135.40 mg/L and reaction temperature of 24.05 °C. Under the optimal conditions, the complete photocatalytic inactivation of MS2 at 108 PFU/mL was achieved after 240 min of visible-light irradiation (Figure 5a). The g-C3N4 was further tested for five consecutive cycles for MS2 inactivation experiments under the optimal conditions, and the results showed the inactivation efficiency remained almost unchanged. It concludes that g-C3N4 has a stable inactivation performance for photocatalytic MS2 under optimal conditions.

3.2. g-C3N4/EP Floating Composite

Expanded perlite (EP)-supported g-C3N4 (g-C3N4/EP) was synthesized and studied for photocatalytic disinfection of MS2 phage [56]. The absorption of visible light by the EP carrier is neglectable, and strong visible-light absorption was observed after loading g-C3N4. The absorption edge of g-C3N4/EP shifted to a longer wavelength as the preparation temperature increased. The g-C3N4/EP-520 (synthesized at 520 °C) presented the best photocatalytic performance. After 240 min of visible-light irradiation at 150 mW/cm2, g-C3N4/EP-520 inactivated 8log MS2 phage (Figure 5b). In general, photocatalytic water disinfection is caused by the redox reaction of photogenerated h+ and e on the surface of the photocatalyst and the production of ROS (e.g., •O2 and •OH). However, since e in the conduction band of g-C3N4 can reduce O2 to •O2 and h+ in the valence band cannot oxidize OH/H2O to •OH, •OH is assumed to play the main role in the inactivation of MS2 using g-C3N4/EP.

3.3. Ag3PO4/g-C3N4 Composite Photocatalyst

Ag3PO4 is a new type of visible-light photocatalyst which has a strong absorption capacity for visible light with a wavelength less than 560 nm. The highly dispersed conduction band of Ag3PO4 and the sensing effect of PO43− contribute to the separation of photogenerated e and h+ [58]. Cheng et al. [57] synthesized Ag3PO4/g-C3N4 (AgCN) composites using a hydrothermal method and applied the composites in the inactivation of viruses in water. A strong light absorption edge was shown at 500 nm, which was attributed to the combination of Ag3PO4 particles. Compared with g-C3N4, the visible-light absorption capacity of AgCN composites was enhanced. With visible-light irradiation, 6.5log bacteriophage f2 was inactivated by the AgCN composite after 90 min (Figure 5c), and the kinetic rate constant of inactivation of AgCN was about 1.5 times greater than that of g-C3N4 alone (Figure 5d).

4. Carbon Composite Materials

Carbon-based material photocatalysts have received extensive attention in the past decade. Graphene and fullerene (such as C60 and C70) are commonly used carbon materials in photocatalytic applications at present. Graphene is one of the most studied nano-carbons due to its unique 2D structure and high conductivity. When graphene is combined with semiconductor photocatalysts, it enhances the availability of active sites and augments the adsorption capacity for reagents, consequently bolstering the photocatalytic efficiency of semiconductor materials [59]. Fullerene can strongly absorb UV light and has a moderate absorption of visible light, and is thus expected as a new type of photocatalyst for disinfection. However, fullerenes are usually not used in their original form but are functionalized with different functional groups [44].

4.1. Cationic Amine-Functionalized C60(HC4)

C60 has extremely low water solubility and can form stable colloidal aggregates in water (nC60), resulting in the loss of its original photochemical properties. It was found that the photochemical properties of C60 can be preserved by presenting electron transfer when C60 bonds with hydrophilic functional groups [60], such as an amine-modified cationic hexakis C60 derivative (HC4) synthesized by Cho et al. [61] (Figure 6a), which can produce singlet oxygen (1O2) under visible-light irradiation for inactivating MS2 phages. The inactivation kinetics were only slightly lower than those of direct sunlight irradiation after using an ultraviolet filter to cut off UV light below 400 nm, indicating that the deactivation was mainly driven by visible-light photocatalysis (Figure 6b).

4.2. Immobilized C60 Photocatalyst

Lee et al. [62] proposed a C60 amino fullerene immobilized on surface-functionalized silica gel, amino C60/SiO2 (Figure 7a). The immobilization improved the photochemical production rate of 1O2 in the C60 amino fullerene solution, and the highest production rate was observed with tetrakis-amino C60/SiO2, which also exhibited the best inactivation effect on MS2 phages (Figure 7b). Cutting off the light below 400 nm did not decrease the inactivation kinetics significantly, corroborating that the tetrakis-amino C60/SiO2 inactivation of viruses is mainly induced by visible light. Moor and Kim [63] developed a silica-supported fullerene material (C60/SiO2) that involved the simple nucleophilic addition of a primary amine across a [6,6] fullerene double bond. It was confirmed that C60/SiO2 prepared with the direct addition of amine had inactivation effects on MS2 phages under visible-light irradiation. MS2 showed a 2 to 2.5 log inactivation rate after 40 min of visible-light irradiation. Choi et al. [64] prepared a C60 amino fullerene–magnetite nanocomposite (C60/msu-SiO2/mag) by covalently immobilizing C60 amino fullerene on functionalized mesoporous mesocellular silica (msu-SiO2) with embedded Fe3O4 nanoparticles (msu-SiO2/mag). This magnetic silica-immobilized amino fullerene material can effectively inactivate MS2 phages by producing 1O2 under visible-light irradiation.

4.3. Immobilized C70 Photocatalyst

C70 has similar visible-light absorption characteristics to C60, but with a stronger absorption. C70 also has a large 1O2 quantum yield, making it an effective photosensitizer. An amine-functionalized silica support (MCM-41)-immobilized C70, C70/MCM-41, was developed by Moor et al. [65]. Compared with the previous C60 and its supported derivative materials, the visible-light activity of MCM-41-immobilized C70 showed a significant improvement. About 3log of virus can be inactivated after 60 min of visible-light irradiation by C70/MCM-41 and 5log after 120 min irradiation (Figure 8a).

4.4. Graphene–Tungsten Oxide Composite (G-WO3)

A composite material composed of semiconductor material and graphene can act as an electron acceptor to hinder the recombination of electron–hole pairs, thus improving the photocatalytic activity of the composite material. These composites promote the development of new materials for the photocatalytic oxidation process [66]. Akhavan et al. [67] developed a graphene–tungsten oxide composite (denoted as G-WO3) by incorporating graphene oxide (GO) into tungsten oxide film following film heat treatment. The study shows that G-WO3 reduces 99.999% of phage MS2 after 3 h of visible-light irradiation (Figure 8b), demonstrating an excellent visible-light response to inactivated phage MS2.

4.5. Graphene Oxide–Aptamer Nanosheet (GOAptamer Nanosheet)

Graphene oxide (GO) has attracted extensive attention due to its high yield, unique surface chemical characteristics and high biocompatibility [66]. However, GO has few specific binding sites to the target and is easy to aggregate, which affects the photocatalytic efficiency. Aptamer is a single-stranded nucleic acid that can bind chemical molecules, nucleic acids, proteins and even microparticles. Hu et al. [68] obtained GO–aptamer nanosheets by covalently bonding the aptamer to GO; GO–aptamer exhibited an obvious reactivity to destroy proteins, suggesting that aptamer played a critical role in photocatalysis, thus improving the photocatalytic activity of GO. As shown in Figure 8c, the inactivation effect of GO–aptamer nanosheets on viruses is better than that of GO nanosheets. After 250 min of visible-light irradiation, the GO–aptamer nanosheets can remove 99.999% of MS2 phage, which is 5 orders of magnitude higher than the dark reaction group. As presented in Figure 8d, the adsorption reached equilibrium after 40 min in the GO–aptamer dark group, and the survival of viruses was dramatically reduced (3 orders of magnitude) on GO–aptamer from 60 min to 240 min of irradiation, presenting a powerful photocatalysis. Figure 8c,d support the dominance of photocatalysis in the inactivation of viruses, with the adsorption from aptamer enhancing the reaction.
Figure 8. (a) MS2 inactivation kinetics of C70/MCM-41 under visible-light irradiation. Reprinted with permission from [65]. Copyright 2015 American Chemical Society. (b) PFU ratio of the infective bacteriophage MS2 viruses remaining on the surfaces of the various thin films under visible-light irradiation and in the dark for 3 h. (Note: The control sample is a glass substrate without any coating). Reprinted with permission from [67]. Copyright 2012 American Chemical Society. Inactivation of MS2 by GO–aptamer nanosheets and GO nanosheets under visible-light or dark conditions (c) aqueous phase and (d) solid. Reproduced from [68], Carbon, with permission from Elsevier Ltd. Press 2012.
Figure 8. (a) MS2 inactivation kinetics of C70/MCM-41 under visible-light irradiation. Reprinted with permission from [65]. Copyright 2015 American Chemical Society. (b) PFU ratio of the infective bacteriophage MS2 viruses remaining on the surfaces of the various thin films under visible-light irradiation and in the dark for 3 h. (Note: The control sample is a glass substrate without any coating). Reprinted with permission from [67]. Copyright 2012 American Chemical Society. Inactivation of MS2 by GO–aptamer nanosheets and GO nanosheets under visible-light or dark conditions (c) aqueous phase and (d) solid. Reproduced from [68], Carbon, with permission from Elsevier Ltd. Press 2012.
Materials 17 00044 g008

5. Photocatalytic Mechanism

Photocatalytic disinfection mainly relies on the oxidizing free radicals produced to inactivate microorganisms. When a photocatalyst is excited by a photon, electron and hole pairs are produced [20]. Holes have oxidizing properties and can directly oxidize microorganisms in contact with photocatalysts, destroying the structure and function of microorganisms [69]. Holes can also react with H2O or OH (at higher pH) to form •OH, thereby destroying the proteins, nucleic acids, lipids and other components of microorganisms via oxidation reactions, and ultimately leading to the death of microorganisms [70,71]. Photogenerated electrons are reducible but can be captured by dissolved O2. During the linking reaction, the trapped electrons produce various oxidizing groups. When the pH value is low, the conduction band electrons meet the energy level requirements of O2 reduction to form HO•. At higher pH (<5), O2 can be reduced to O2•− by conduction band electrons. O2•− react with two dissociated hydrogen ions and a photogenerated electron to form H2O2. H2O2 can react with a dissociated hydrogen ion and a photogenerated electron to form •OH [72]. A schematic diagram of the photocatalytic mechanism is shown in Figure 9a. According to the above discussion, a possible photocatalytic inactivation mechanism of the g-C3N4/Ag3PO4 composite photocatalysts is shown in Figure 9b. Both g-C3N4 and Ag3PO4 can be irradiated to generate electron–hole pairs under visible-light irradiation. The photoexcited electrons from the valence bands (VBs) of g-C3N4 and Ag3PO4 efficiently transfer to their respective conduction bands (CBs). Moreover, electrons from the CB of Ag3PO4 can migrate to the VB of g-C3N4. The electrons stored in the CB of g-C3N4 react with dissolved oxygen, resulting in the formation of reactive O2•−. Simultaneously, the holes (h+) in the VB of Ag3PO4 react with water or OH, generating •OH. These highly active radical species, along with photogenerated holes, contribute to the oxidation of bacteriophage f2.
In summary, photocatalytic disinfection leverages the production of oxidizing free radicals, such as holes and hydroxyl radicals, to disrupt the molecular structure of microorganisms. This mechanism provides photocatalytic disinfection with broad-spectrum antimicrobial efficacy, exhibiting relatively low resistance from various microorganisms. Moreover, it addresses concerns about secondary pollution. Consequently, photocatalytic disinfection has diverse applications in water treatment, environmental sanitation and the healthcare sector.

6. Summary and Discussions

Considering both removal efficiency and time consumption rate, Cu(II) −TiO2, CuxO/TiO2 nanocomposites and TiON/PdO are more recommended. Table 1 summarizes the above visible-light photocatalyst materials for virus disinfection, and the following conclusions can be drawn.
(1) Common semiconductor materials (such as TiO2 and ZnO) have been used for photocatalytic disinfection in the past few years, which, however, can only be activated by UV. Methods have been developed to extend the light-absorption range of ultraviolet materials to achieve visible-light response, thereby performing photocatalytic disinfection in the visible-light region.
(2) Graphitic carbon nitride can be directly activated by visible light to achieve light response, but the utilization rate of visible light is low. The inactivation efficiency can be improved by changing the parameters of the disinfection process and by using a floating base material or compounding with other visible-light catalytic materials to enhance its photocatalytic effect.
(3) Carbon-based materials are promoted as photocatalysts for water disinfection due to their low cost and no risk of metal contamination. Graphene and fullerene materials can achieve light response in visible regions and have good disinfection properties. The functionalization and immobilization of carbon-based materials can further increase the efficiency of photocatalytic virus inactivation and be applied to practical applications.

7. Conclusions and Prospects

This paper reviews the latest developments in the utilization of visible-light composite photocatalysts for phage inactivation, encompassing the synthesis, characteristics, disinfection mechanism and overall efficiency of photocatalysts. Three types of photocatalytic materials have been reviewed. Photocatalysis as a green oxidation technology has opened up a simple and efficient way to inactivate viruses in water. The most widely used photocatalyst is TiO2. However, with continuous exploration and study, many novel photocatalyst materials have been developed, and their performance exceeds that of TiO2, which shows great potential in photocatalytic disinfection. To date, visible-light composite catalysts have demonstrated exceptional efficacy in virus removal. The ongoing research and development of visible-light catalytic materials remain a prevailing trajectory for the future. Efforts to enhance the efficiency and adaptability of photocatalytic processes are driving the exploration and proposition of novel material compositions. This includes the integration of cutting-edge nanomaterials like organic semiconductors, metal–organic frameworks (MOFs), covalent organic frameworks (COFs) and quantum dots. MOFs possess a remarkably adjustable pore structure, leveraging their role in photocatalysis for adsorption, electron transfer and catalytic reactions. Specifically, photoresponsive MOFs exhibit potential for photocatalytic water splitting to generate hydrogen. Meanwhile, COFs boast exceptional electronic conductivity, predominantly applied in photocatalysis for carbon dioxide reduction and water splitting. Carbon quantum dots exhibit remarkable optoelectronic properties and can synergistically combine with other photosensitive materials, such as TiO2 or ZnO, to enhance their photocatalytic activity. However, the widespread practical application of these catalysts encounters obstacles such as high production costs and intricate synthesis methods. Furthermore, most of the photocatalytic materials are produced as powder, leading to a pressing need within the field for the creation of easily recyclable photocatalysts. The evolution of composite photocatalytic materials for disinfection is advancing across various formats such as synthetic sprays, gels, masks and more. The introduction of these composite materials into sprays has proven highly effective in eradicating airborne bacteria and viruses via photocatalytic processes. Gel-based applications provide a convenient and long-lasting solution for personal hygiene in hand-sanitizing gels, ensuring extended disinfection. Incorporating composite photocatalytic materials into mask manufacturing serves to filter airborne microorganisms, curbing the spread of pathogens. Maintaining cost-effectiveness is pivotal for the widespread adoption of these products in large-scale applications. Addressing safety concerns related to potential adverse effects on human health is equally crucial. Rigorous scientific evaluation and stringent quality control protocols are imperative to ensure the reliability of these technologies, providing a dependable means for disease prevention within society. More applicable visible-light composite catalysts need to be further studied and explored in the future.

Funding

This research was funded by by the Graduate Research Innovation Project of Chongqing Jiaotong University (Grant No. 2023S0039), Chongqing Postgraduate Joint Training Base Project (JDLHPYJD2022005), the Science and Technology Research Program of Chongqing Municipal Education Commission (Grant No. KJQN202100739), Natural Science Foundation of Chongqing, China (Grant No. cstc2021jcyj-msxmX0912), the Foundation for High-level Talents of Chongqing Jiaotong University (19JDKJC-C023, 20JDKIC-A042), the China Scholar Council (CSC) and the APC was funded by Stockholm University.

Data Availability Statement

The data are openly available without restrictions.

Conflicts of Interest

Author Qingkong Cheng was employed by the company Joint Graduate Training Base for Resources and Environment between Chongqing Jiaotong University and Chongqing Gangli Environmental Protection Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Ban, H.; Duan, H.; Bai, X. Progress on Research of Virus Contamination in Drinking Water and Evaluation Indicators. J. Environ. Hyg. 2017, 7, 321–329. [Google Scholar]
  2. Jebri, S.; Hmaied, F.; Jofre, J.; MariemYahya; Mendez, J.; Barkallah, I.; Hamdi, M. Effect of gamma irradiation on bacteriophages used as viral indicators. Water Res. 2013, 47, 3673–3678. [Google Scholar] [CrossRef] [PubMed]
  3. Torrey, J.; von Gunten, U.; Kohn, T. Differences in Viral Disinfection Mechanisms as Revealed by Quantitative Transfection of Echovirus 11 Genomes. Appl. Environ. Microbiol. 2019, 85, e00961-19. [Google Scholar] [CrossRef] [PubMed]
  4. Papageorgiou, A.; Voutsa, D.; Papadakis, N. Occurrence and fate of ozonation by-products at a full-scale drinking water treatment plant. Sci. Total Environ. 2014, 481, 392–400. [Google Scholar] [CrossRef]
  5. Wang, W.; Huang, G.; Yu, J.C.; Wong, P.K. Advances in photocatalytic disinfection of bacteria: Development of photocatalysts and mechanisms. J. Environ. Sci. 2015, 34, 232–247. [Google Scholar] [CrossRef]
  6. Dotson, A.D.; Rodriguez, C.E.; Linden, K.G. UV disinfection implementation status in US water treatment plants. J. Am. Water Work. Assoc. 2012, 104, E318–E324. [Google Scholar] [CrossRef]
  7. Karnik, B.S.; Davies, S.H.; Baumann, M.J.; Masten, S.J. The effects of combined ozonation and filtration on disinfection by-product formation. Water Res. 2005, 39, 2839–2850. [Google Scholar] [CrossRef]
  8. Yu, L.; Wang, H.; Huang, Q.; Liu, H.; Chen, Q.; Yuan, B.; Li, Q.; Zhao, X.; Tang, J.; Zhao, D. One-pot microwave synthesized high-performance BiVO4/InVO4 heterojunction for photocatalytic reduction of Cr6+. Sep. Purif. Technol. 2023, 310, 123143. [Google Scholar] [CrossRef]
  9. Luo, Y.; Zhou, S.; Sun, Y.; Zhao, D.; Gu, X.; Yan, Q.; Ji, F.; Xu, X. Study of the photocatalytic performance and carrier migration of reduced graphene oxide-modified infrared-responsive photocatalyst β-NaYF4:Yb3+, Er3+@BiOCl-rGO. Opt. Mater. 2019, 96, 109312. [Google Scholar] [CrossRef]
  10. Gu, X.; Yan, Q.; Wei, Y.; Luo, Y.; Sun, Y.; Zhao, D.; Ji, F.; Xu, X. Visible-light-responsive photocatalyst with a microsphere structure: Preparation and photocatalytic performance of CQDs@BiOCl. J. Mater. Sci. Mater. Electron. 2019, 30, 16321–16336. [Google Scholar] [CrossRef]
  11. Zhao, D.; Wang, W.; Zong, W.; Xiong, S.; Zhang, Q.; Ji, F.; Xu, X. Synthesis of Bi2S3/BiVO4 Heterojunction with a One-Step Hydrothermal Method Based on pH Control and the Evaluation of Visible-Light Photocatalytic Performance. Materials 2017, 10, 891. [Google Scholar] [CrossRef] [PubMed]
  12. Du, M.; Xiong, S.; Wu, T.; Zhao, D.; Zhang, Q.; Fan, Z.; Zeng, Y.; Ji, F.; He, Q.; Xu, X. Preparation of a Microspherical Silver-Reduced Graphene Oxide-Bismuth Vanadate Composite and Evaluation of Its Photocatalytic Activity. Materials 2016, 9, 160. [Google Scholar] [CrossRef] [PubMed]
  13. Xu, X.; Zhou, B.; Ji, F.; Zou, Q.; Yuan, Y.; Jin, Z.; Zhao, D.; Long, J. Nitrification, Denitrification, and Power Generation Enhanced by Photocatalysis in Microbial Fuel Cells in the Absence of Organic Compounds. Energy Fuels 2015, 29, 1227–1232. [Google Scholar] [CrossRef]
  14. Zhao, D.; Wang, W.; Sun, Y.; Fan, Z.; Du, M.; Zhang, Q.; Ji, F.; Xu, X. One-step synthesis of composite material MWCNT@BiVO4 and its photocatalytic activity. RSC Adv. 2017, 7, 33671–33679. [Google Scholar] [CrossRef]
  15. Zhao, D.; Zong, W.; Fan, Z.; Xiong, S.; Du, M.; Wu, T.; Fang, Y.; Ji, F.; Xu, X. Synthesis of carbon-doped BiVO4@multi-walled carbon nanotubes with high visible-light absorption behavior, and evaluation of their photocatalytic properties. CrystEngComm 2016, 18, 9007–9015. [Google Scholar] [CrossRef]
  16. Han, X.; Wang, L.; Wang, L.; Wang, X.; Zhao, D. Preparation of α-Fe2O3-Modified Hollow Glass Microspheres/ZnIn2S4 Composite Catalyst and Enhanced Photocatalytic Performance. Chin. Ceram. Soc. 2020, 7, 10. [Google Scholar]
  17. Zhou, S.; Zhang, Q.; Zhao, D.; Zong, W.; Fan, Z.; Sun, Y.; Xu, X. Synthesis, properties and mechanism of photodegradation of core–shell structured upconversion luminescent NaYF4: Yb3+, Er3+@BiOCl. Appl. Organomet. Chem. 2018, 32, e4230. [Google Scholar] [CrossRef]
  18. Huang, Q.; Zhao, Z.; Zhao, X.; Li, Q.; Liu, H.; Lu, H.; Li, Q.; Du, M.; Cao, Y.; Wang, Y.; et al. Effective photocatalytic sterilization based on composites of Ag/InVO4/BiOBr: Factors, mechanism and application. Sep. Purif. Technol. 2023, 327, 125011. [Google Scholar] [CrossRef]
  19. Zhang, C.; Li, Y.; Wang, C.; Zheng, X. Different inactivation behaviors and mechanisms of representative pathogens (Escherichia coli bacteria, human adenoviruses and Bacillus subtilis spores) in g-C3N4-based metal-free visible-light-enabled photocatalytic disinfection. Sci. Total Environ. 2021, 755, 142588. [Google Scholar] [CrossRef]
  20. Saravanan, A.; Kumar, P.S.; Jeevanantham, S.; Karishma, S.; Kiruthika, A.R. Photocatalytic disinfection of micro-organisms: Mechanisms and applications. Environ. Technol. Innov. 2021, 24, 101909. [Google Scholar] [CrossRef]
  21. Matsunaga, T.; Tomoda, R.; Nakajima, T.; Wake, H. Photoelectrochemical sterilization of microbial cells by semiconductor powders. FEMS Microbiol. Lett. 1985, 29, 211–214. [Google Scholar] [CrossRef]
  22. Dunlop, P.S.M.; Byrne, J.A.; Manga, N.; Eggins, B.R. The photocatalytic removal of bacterial pollutants from drinking water. J. Photochem. Photobiol. A Chem. 2002, 148, 355–363. [Google Scholar] [CrossRef]
  23. Lee, H.U.; Lee, S.C.; Choi, S.; Son, B.; Lee, S.M.; Kim, H.J.; Lee, J. Efficient visible-light induced photocatalysis on nanoporous nitrogen-doped titanium dioxide catalysts. Chem. Eng. J. 2013, 228, 756–764. [Google Scholar] [CrossRef]
  24. Akhavan, O.; Mehrabian, M.; Mirabbaszadeh, K.; Azimirad, R. Hydrothermal synthesis of ZnO nanorod arrays for photocatalytic inactivation of bacteria. J. Phys. D Appl. Phys. 2009, 42, 225305. [Google Scholar] [CrossRef]
  25. Hunge, Y.; Mahadik, M.; Mohite, V.; Kumbhar, S.; Deshpande, N.; Rajpure, K.; Moholkar, A.; Patil, P.; Bhosale, C. Photoelectrocatalytic degradation of methyl blue using sprayed WO3 thin films. J. Mater. Sci. Mater. Electron. 2016, 27, 1629–1635. [Google Scholar] [CrossRef]
  26. Zhang, C.; Li, Y.; Shuai, D.; Shen, Y.; Xiong, W.; Wang, L. Graphitic carbon nitride (g-C3N4)-based photocatalysts for water disinfection and microbial control: A review. Chemosphere 2019, 214, 462–479. [Google Scholar] [CrossRef]
  27. Gomes, J.; Matos, A.; Gmurek, M.; Quinta-Ferreira, R.; Martins, R. Ozone and Photocatalytic Processes for Pathogens Removal from Water: A Review. Catalysts 2019, 9, 46. [Google Scholar] [CrossRef]
  28. Misstear, D.B.; Gill, L.W. The inactivation of phages MS2, ΦX174 and PR772 using UV and solar photocatalysis. J. Photochem. Photobiol. B Biol. 2012, 107, 1–8. [Google Scholar] [CrossRef]
  29. Sunada, K.; Watanabe, T.; Hashimoto, K. Bactericidal Activity of Copper-Deposited TiO2 Thin Film under Weak UV Light Illumination. Environ. Sci. Technol. 2003, 37, 4785–4789. [Google Scholar] [CrossRef]
  30. Erkan, A.; Bakir, U.; Karakas, G. Photocatalytic microbial inactivation over Pd doped SnO2 and TiO2 thin films. J. Photochem. Photobiol. A Chem. 2006, 184, 313–321. [Google Scholar] [CrossRef]
  31. Thongsuriwong, K.; Amornpitoksuk, P.; Suwanboon, S. Photocatalytic and antibacterial activities of Ag-doped ZnO thin films prepared by a sol–gel dip-coating method. J. Sol-Gel Sci. Technol. 2012, 62, 304–312. [Google Scholar] [CrossRef]
  32. Liga, M.V.; Bryant, E.L.; Colvin, V.L.; Li, Q.L. Virus inactivation by silver doped titanium dioxide nanoparticles for drinking water treatment. Water Res. 2011, 45, 535–544. [Google Scholar] [CrossRef] [PubMed]
  33. Venieri, D.; Gounaki, I.; Binas, V.; Zachopoulos, A.; Kiriakidis, G.; Mantzavinos, D. Inactivation of MS2 coliphage in sewage by solar photocatalysis using metal-doped TiO2. Photocatal. Sci. Appl. 2015, 178, 54–64. [Google Scholar] [CrossRef]
  34. Zhong, Q.; Zhang, W.D.; Shi, Q.; Yu, Y. 0D/2D CQDs/Bi7O9I3 composite with high photocatalytic disinfection performance under visible light. J. Solid State Chem. 2021, 302, 122426. [Google Scholar] [CrossRef]
  35. Wang, T.; Zheng, J.; Cai, J.; Liu, Q.; Zhang, X. Visible-light-driven photocatalytic degradation of dye and antibiotics by activated biochar composited with K+ doped g-C3N4: Effects, mechanisms, actual wastewater treatment and disinfection. Sci. Total Environ. 2022, 839, 155955. [Google Scholar] [CrossRef] [PubMed]
  36. Ji, L.; Liu, B.; Qian, Y.; Yang, Q.; Gao, P. Enhanced visible-light-induced photocatalytic disinfection of Escherichia coli by ternary Bi2WO6/TiO2/reduced graphene oxide composite materials: Insight into the underlying mechanism. Adv. Powder Technol. 2020, 31, 128–138. [Google Scholar] [CrossRef]
  37. Hosseini, S.M.; Safarifard, V. MoS2@MOF composites: Design strategies and photocatalytic applications. Mater. Sci. Semicond. Process. 2024, 169, 107892. [Google Scholar] [CrossRef]
  38. Dhiman, P.; Sharma, J.; Kumar, A.; Sharma, G.; Dawi, E. Photocatalytic degradation of malachite green by waste derived bio-char impregnated Bi5O7Br/Fe3O4 magnetic composite. Opt. Mater. 2023, 146, 114568. [Google Scholar] [CrossRef]
  39. Qiu, J.; Zhang, X.; Feng, Y.; Zhang, X.; Wang, H.; Yao, J. Modified metal-organic frameworks as photocatalysts. Appl. Catal. B Environ. 2018, 231, 317–342. [Google Scholar] [CrossRef]
  40. Zheng, X.; Shen, Z.P.; Cheng, C.; Shi, L.; Cheng, R.; Yuan, D.H. Photocatalytic disinfection performance in virus and virus/bacteria system by Cu-TiO2 nanofibers under visible light. Environ. Pollut. 2018, 237, 452–459. [Google Scholar] [CrossRef]
  41. Irie, H.; Miura, S.; Kamiya, K.; Hashimoto, K. Efficient visible light-sensitive photocatalysts: Grafting Cu(II) ions onto TiO2 and WO3 photocatalysts. Chem. Phys. Lett. 2008, 457, 202–205. [Google Scholar] [CrossRef]
  42. Irie, H.; Kamiya, K.; Shibanuma, T.; Miura, S.; Tryk, D.; Yokoyama, T.; Hashimoto, K. Visible Light-Sensitive Cu(II)-Grafted TiO2 Photocatalysts: Activities and X-ray Absorption Fine Structure Analyses. J. Phys. Chem. C 2009, 113, 10761–10766. [Google Scholar] [CrossRef]
  43. Liu, M.; Sunada, K.; Hashimoto, K.; Miyauchi, M. Visible-light sensitive Cu(Ii)–TiO2 with sustained anti-viral activity for efficient indoor environmental remediation. J. Mater. Chem. A 2015, 3, 17312–17319. [Google Scholar] [CrossRef]
  44. Albiter, E.; Barrera-Andrade, J.M.; Rojas-García, E.; Valenzuela, M. 17—Recent advances of nanocarbon-inorganic hybrids in photocatalysis. In Nanocarbon and Its Composites; Khan, A., Jawaid, M., Inamuddin, Asiri, A., Eds.; Woodhead Publishing: Sawston, UK, 2019; pp. 521–588. [Google Scholar]
  45. Miyauchi, M.; Sunada, K.; Hashimoto, K. Antiviral Effect of Visible Light-Sensitive CuxO/TiO2 Photocatalyst. Catalysts 2020, 10, 1093. [Google Scholar] [CrossRef]
  46. Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Visible-Light Photocatalysis in Nitrogen-Doped Titanium Oxides. Science 2001, 293, 269–271. [Google Scholar] [CrossRef] [PubMed]
  47. Li, Q.; Page, M.A.; Mariñas, B.J.; Shang, J.K. Treatment of Coliphage MS2 with Palladium-Modified Nitrogen-Doped Titanium Oxide Photocatalyst Illuminated by Visible Light. Environ. Sci. Technol. 2008, 42, 6148–6153. [Google Scholar] [CrossRef]
  48. Li, Q.; Xie, R.; Mintz, E.A.; Shang, J.K. Enhanced Visible-Light Photocatalytic Degradation of Humic Acid by Palladium-Modified Nitrogen-Doped Titanium Oxide. J. Am. Ceram. Soc. 2007, 90, 3863–3868. [Google Scholar] [CrossRef]
  49. Di Mauro, A.; Fragalà, M.E.; Privitera, V.; Impellizzeri, G. ZnO for application in photocatalysis: From thin films to nanostructures. Mater. Sci. Semicond. Process. 2017, 69, 44–51. [Google Scholar] [CrossRef]
  50. Zhao, J.; Wang, L.; Yan, X.; Yang, Y.; Lei, Y.; Zhou, J.; Huang, Y.; Gu, Y.; Zhang, Y. Structure and photocatalytic activity of Ni-doped ZnO nanorods. Mater. Res. Bull. 2011, 46, 1207–1210. [Google Scholar] [CrossRef]
  51. Elilarassi, R.; Chandrasekaran, G. Synthesis, structural and optical characterization of Ni-doped ZnO nanoparticles. J. Mater. Sci. Mater. Electron. 2011, 22, 751–756. [Google Scholar] [CrossRef]
  52. Nasir, A.M.; Jaafar, J.; Aziz, F.; Yusof, N.; Salleh, W.N.W.; Ismail, A.F.; Aziz, M. A review on floating nanocomposite photocatalyst: Fabrication and applications for wastewater treatment. J. Water Process Eng. 2020, 36, 101300. [Google Scholar] [CrossRef]
  53. Urbonavicius, M.; Varnagiris, S.; Tuckute, S.; Sakalauskaite, S.; Demikyte, E.; Lelis, M. Visible-Light-Driven Photocatalytic Inactivation of Bacteria, Bacteriophages, and Their Mixtures Using ZnO-Coated HDPE Beads as Floating Photocatalyst. Materials 2022, 15, 1318. [Google Scholar] [CrossRef] [PubMed]
  54. Dao, V.D.; Son, L.T.; Nguyen, T.D.; Van Noi, N.; Ngoc, N.M.; Pham, T.; Van Quan, P.; Trang, H. Superior visible light photocatalytic activity of g-C3N4/NiWO4 direct Z system for degradation of gaseous toluene. J. Solid State Chem. 2019, 272, 62–68. [Google Scholar] [CrossRef]
  55. Zhang, C.; Li, Y.; Zhang, W.; Wang, P.; Wang, C. Metal-free virucidal effects induced by g-C3N4 under visible light irradiation: Statistical analysis and parameter optimization. Chemosphere 2018, 195, 551–558. [Google Scholar] [CrossRef] [PubMed]
  56. Zhang, C.; Li, Y.; Shuai, D.; Zhang, W.; Niu, L.; Wang, L.; Zhang, H. Visible-light-driven, water-surface-floating antimicrobials developed from graphitic carbon nitride and expanded perlite for water disinfection. Chemosphere 2018, 208, 84–92. [Google Scholar] [CrossRef] [PubMed]
  57. Cheng, R.; Shen, L.J.; Yu, J.H.; Xiang, S.Y.; Zheng, X. Photocatalytic Inactivation of Bacteriophage f2 with Ag3PO4/g-C3N4 Composite under Visible Light Irradiation: Performance and Mechanism. Catalysts 2018, 8, 406. [Google Scholar] [CrossRef]
  58. Ge, M.; Li, Z. Recent progress in Ag3PO4-based all-solid-state Z-scheme photocatalytic systems. Chin. J. Catal. 2017, 38, 1794–1803. [Google Scholar] [CrossRef]
  59. Cao, S.; Yu, J. Carbon-based H2-production photocatalytic materials. J. Photochem. Photobiol. C Photochem. Rev. 2016, 27, 72–99. [Google Scholar] [CrossRef]
  60. Lee, J.; Mackeyev, Y.; Cho, M.; Li, D.; Kim, J.H.; Wilson, L.J.; Alvarez, P.J.J. Photochemical and Antimicrobial Properties of Novel C60 Derivatives in Aqueous Systems. Environ. Sci. Technol. 2009, 43, 6604–6610. [Google Scholar] [CrossRef]
  61. Cho, M.; Lee, J.; Mackeyev, Y.; Wilson, L.J.; Alvarez, P.J.J.; Hughes, J.B.; Kim, J.H. Visible Light Sensitized Inactivation of MS-2 Bacteriophage by a Cationic Amine-Functionalized C60 Derivative. Environ. Sci. Technol. 2010, 44, 6685–6691. [Google Scholar] [CrossRef]
  62. Lee, J.; Mackeyev, Y.; Cho, M.; Wilson, L.J.; Kim, J.H.; Alvarez, P.J.J. C60 Aminofullerene Immobilized on Silica as a Visible-Light-Activated Photocatalyst. Environ. Sci. Technol. 2010, 44, 9488–9495. [Google Scholar] [CrossRef] [PubMed]
  63. Moor, K.J.; Kim, J.H. Simple Synthetic Method Toward Solid Supported C60 Visible Light-Activated Photocatalysts. Environ. Sci. Technol. 2014, 48, 2785–2791. [Google Scholar] [CrossRef] [PubMed]
  64. Choi, Y.; Ye, Y.; Mackeyev, Y.; Cho, M.; Lee, S.; Wilson, L.J.; Lee, J.; Alvarez, P.J.J.; Choi, W.; Lee, J. C60 aminofullerene-magnetite nanocomposite designed for efficient visible light photocatalysis and magnetic recovery. Carbon 2014, 69, 92–100. [Google Scholar] [CrossRef]
  65. Moor, K.J.; Valle, D.C.; Li, C.; Kim, J.H. Improving the Visible Light Photoactivity of Supported Fullerene Photocatalysts through the Use of [C70] Fullerene. Environ. Sci. Technol. 2015, 49, 6190–6197. [Google Scholar] [CrossRef] [PubMed]
  66. Olatunde, O.C.; Onwudiwe, D.C. Chapter 3—Copper-based ternary metal sulfide nanocrystals embedded in graphene oxide as photocatalyst in water treatment. In Nanotechnology in the Beverage Industry; Amrane, A., Rajendran, S., Nguyen, T.A., Assadi, A.A., Sharoba, A.M., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 51–113. [Google Scholar]
  67. Akhavan, O.; Choobtashani, M.; Ghaderi, E. Protein Degradation and RNA Efflux of Viruses Photocatalyzed by Graphene–Tungsten Oxide Composite Under Visible Light Irradiation. J. Phys. Chem. C 2012, 116, 9653–9659. [Google Scholar] [CrossRef]
  68. Hu, X.; Mu, L.; Wen, J.; Zhou, Q. Covalently synthesized graphene oxide-aptamer nanosheets for efficient visible-light photocatalysis of nucleic acids and proteins of viruses. Carbon 2012, 50, 2772–2781. [Google Scholar] [CrossRef]
  69. Xu, X.; Sun, Y.; Fan, Z.; Zhao, D.; Xiong, S.; Zhang, B.; Zhou, S.; Liu, G. Mechanisms for ·O2 and ·OH production on flowerlike BiVO4 photocatalysis based on electron spin resonance. Front. Chem. 2018, 6, 64. [Google Scholar] [CrossRef]
  70. Lebedev, A.; Anariba, F.; Tan, J.C.; Li, X.; Wu, P. A review of physiochemical and photocatalytic properties of metal oxides against Escherichia coli. J. Photochem. Photobiol. A Chem. 2018, 360, 306–315. [Google Scholar] [CrossRef]
  71. Bora, L.V.; Bora, N.V.; Koladia, G. Intensification strategies of solar photocatalytic disinfection—A review. Inorg. Chem. Commun. 2023, 157, 111292. [Google Scholar] [CrossRef]
  72. Yu, H.; Chen, S.; Quan, X.; Zhang, Z. The Mechanism, Materials and Reactors of Photocatalytic Disinfection in Water and Wastewater Treatment. Prog. Chem. 2017, 29, 1030–1041. [Google Scholar]
Figure 1. (a) Photocatalytic disinfection of bacteriophage f2 using different catalyst dosages. Visible-light intensity: 100 mW/cm2, initial virus concentration: 105 PFU/mL, pH = 7, temperature: 25 °C. Reproduced from [40], Environmental Pollution, with permission from Elsevier Ltd. Press 2018. (b) UV−vis diffuse reflectance spectra of rutile TiO2 and Cu (II)/TiO2. Reproduced with permission from [42]. Copyright 2009 American Chemical Society. Inactivation of bacteriophage Qβ by photocatalyst under dark conditions and visible-light irradiation. (c) Cu(II)-TiO2 (pH = 7), (d) Cu(II)-TiO2 (pH = 12). Reproduced from [43], Journal of Materials Chemistry A, with permission from The Royal Society of Chemistry Press 2015.
Figure 1. (a) Photocatalytic disinfection of bacteriophage f2 using different catalyst dosages. Visible-light intensity: 100 mW/cm2, initial virus concentration: 105 PFU/mL, pH = 7, temperature: 25 °C. Reproduced from [40], Environmental Pollution, with permission from Elsevier Ltd. Press 2018. (b) UV−vis diffuse reflectance spectra of rutile TiO2 and Cu (II)/TiO2. Reproduced with permission from [42]. Copyright 2009 American Chemical Society. Inactivation of bacteriophage Qβ by photocatalyst under dark conditions and visible-light irradiation. (c) Cu(II)-TiO2 (pH = 7), (d) Cu(II)-TiO2 (pH = 12). Reproduced from [43], Journal of Materials Chemistry A, with permission from The Royal Society of Chemistry Press 2015.
Materials 17 00044 g001
Figure 2. Inactivation efficiency of bacteriophage Qβ by CuxO/TiO and TiO2−xNx samples under white light and dark conditions (using white fluorescent bulbs with UV-blocking film for light exposure). Reproduced from [45], Catalysts, with permission from MDPI Press 2020.
Figure 2. Inactivation efficiency of bacteriophage Qβ by CuxO/TiO and TiO2−xNx samples under white light and dark conditions (using white fluorescent bulbs with UV-blocking film for light exposure). Reproduced from [45], Catalysts, with permission from MDPI Press 2020.
Materials 17 00044 g002
Figure 3. Inactivation of MS2 under different conditions using only visible light (>400 nm) (⧫), TiON/PdO photocatalyst in a dark environment (○Δ∇), and TiON/PdO photocatalyst with visible light (>400 nm) irradiation after corresponding dark control (●▲▼)). Reprinted with permission from [47]. Copyright 2008 American Chemical Society.
Figure 3. Inactivation of MS2 under different conditions using only visible light (>400 nm) (⧫), TiON/PdO photocatalyst in a dark environment (○Δ∇), and TiON/PdO photocatalyst with visible light (>400 nm) irradiation after corresponding dark control (●▲▼)). Reprinted with permission from [47]. Copyright 2008 American Chemical Society.
Materials 17 00044 g003
Figure 4. Infectivity of separate PRD1 and T4 phages and their mixture under visible-light irradiation for 1 h using ZnO-HDPE (a) without Ni underlayer and (b) with Ni underlayer. Reproduced from [53], Materials, with permission from MDPI Press 2022.
Figure 4. Infectivity of separate PRD1 and T4 phages and their mixture under visible-light irradiation for 1 h using ZnO-HDPE (a) without Ni underlayer and (b) with Ni underlayer. Reproduced from [53], Materials, with permission from MDPI Press 2022.
Materials 17 00044 g004
Figure 5. (a) Inactivation efficiency against MS2 (∼1 × 108 PFU/mL, 100 mL) by g-C3N4 under visible-light irradiation under optimal conditions (199.80 mW/cm2, 135.40 mg/L and 24.05 °C for light intensity, photocatalyst loading and reaction temperature, respectively). Reproduced from [55], Chemosphere, with permission from Elsevier Ltd. Press 2018. (b) The inactivation of MS2 phage by g-C3N4/EP under visible-light irradiation. Reproduced from [56], Chemosphere, with permission from Elsevier Ltd. Press 2018. (c) Inactivation efficiency toward bacteriophage f2 by different materials under visible-light irradiation; (d) linear fitting of the kinetic curves of inactivation. Reproduced from [57], Catalysts, with permission from MDPI Press 2018.
Figure 5. (a) Inactivation efficiency against MS2 (∼1 × 108 PFU/mL, 100 mL) by g-C3N4 under visible-light irradiation under optimal conditions (199.80 mW/cm2, 135.40 mg/L and 24.05 °C for light intensity, photocatalyst loading and reaction temperature, respectively). Reproduced from [55], Chemosphere, with permission from Elsevier Ltd. Press 2018. (b) The inactivation of MS2 phage by g-C3N4/EP under visible-light irradiation. Reproduced from [56], Chemosphere, with permission from Elsevier Ltd. Press 2018. (c) Inactivation efficiency toward bacteriophage f2 by different materials under visible-light irradiation; (d) linear fitting of the kinetic curves of inactivation. Reproduced from [57], Catalysts, with permission from MDPI Press 2018.
Materials 17 00044 g005
Figure 6. (a) Synthesis of cationic hexakis C60 derivative (called HC4). (b) Inactivation of MS2 phage by HC4 under fluorescent-light irradiation (22 °C, pH = 7.0, [MS2] = 4−6 × 104 pfu/mL, [HC4] = 15 μM, light intensity at 365 nm = 165 μW/cm2). Reprinted with permission from [61]. Copyright 2010 American Chemical Society.
Figure 6. (a) Synthesis of cationic hexakis C60 derivative (called HC4). (b) Inactivation of MS2 phage by HC4 under fluorescent-light irradiation (22 °C, pH = 7.0, [MS2] = 4−6 × 104 pfu/mL, [HC4] = 15 μM, light intensity at 365 nm = 165 μW/cm2). Reprinted with permission from [61]. Copyright 2010 American Chemical Society.
Materials 17 00044 g006
Figure 7. (a) Strategy for immobilization of water-soluble C60 amino fullerenes on functionalized silica gel (n = 2, 4, and 6). (b) MS2 phage inactivation by immobilized bis-, tetrakis-, and hexakis-amino C60/silica under visible light and under commercial fluorescence lamp irradiation. Reprinted with permission from [62]. Copyright 2010 American Chemical Society.
Figure 7. (a) Strategy for immobilization of water-soluble C60 amino fullerenes on functionalized silica gel (n = 2, 4, and 6). (b) MS2 phage inactivation by immobilized bis-, tetrakis-, and hexakis-amino C60/silica under visible light and under commercial fluorescence lamp irradiation. Reprinted with permission from [62]. Copyright 2010 American Chemical Society.
Materials 17 00044 g007
Figure 9. (a) Schematic diagram of photocatalytic disinfection mechanism. (b) The possible mechanism of photocatalytic inactivation of bacteriophage f2 by the g−C3N4/Ag3PO4 photocatalysts under visible−light irradiation [57]. Catalysts, with permission from MDPI Press 2018.
Figure 9. (a) Schematic diagram of photocatalytic disinfection mechanism. (b) The possible mechanism of photocatalytic inactivation of bacteriophage f2 by the g−C3N4/Ag3PO4 photocatalysts under visible−light irradiation [57]. Catalysts, with permission from MDPI Press 2018.
Materials 17 00044 g009
Table 1. Summary of Visible-Light Photocatalytic Materials for Virus Inactivation.
Table 1. Summary of Visible-Light Photocatalytic Materials for Virus Inactivation.
Type of CompositesPhotocatalystTarget MicrobeInitial Concentration (PFU/mL)Light Source and WavelengthLight Intensity (mW/cm2)EfficiencyTime (min)Reference
Modified ultraviolet-response materialCu-TiO2 nanofibersBacteriophage f21 × 105300 W Xe lamp70−2.5log240[40]
Cu(II)-TiO2Bacteriophage Qβ1 × 108Fluorescent lampNA−6.5log30[43]
CuxO/TiO2 nanocompositesBacteriophage Qβ1 × 108Fluorescent lampNA−7.5log40[45]
TiON/PdOBacteriophage MS23 × 1081000 W xenon-arc lamp4094.5–98.2%60[47]
ZnO-based floating photocatalystPhage T41 × 108High power led6562–82%60[53]
Phage T4 + Phage PRD11 × 108High power led6550–53%60[53]
Direct visible-light response materialsg-C3N4Bacteriophage MS21 × 108Xe lamp199.8−8log240[55]
g-C3N4/EPBacteriophage MS21 × 108300 W Xe lamp150−8log240[56]
Ag3PO4/g-C3N4Bacteriophage f21 × 106Xe lamp300−6.5log80[57]
Carbon materialsCationic hexakis C60Bacteriophage MS24–6 × 104Fluorescent lamp165−2.1log60[61]
tetrakis-amino C60/SiO2Bacteriophage MS22 × 105Fluorescent lamp165−2.4log75[62]
C60/SiO2Bacteriophage MS25 × 104Fluorescent lampNA−2–−2.5log40[63]
C60 Amino fullerene-magnetiteBacteriophage MS21 × 106300 W Xe lampNA−2.7log75[64]
C70/SiO2Bacteriophage MS23 × 108Fluorescent lamp4.23−4.35log90[65]
G-WO3Bacteriophage MS22 × 106Mercury lamp110−5.9log180[67]
GO-aptamer nanosheetBacteriophage MS26 × 106300 W tungsten lampNA−5.9log240[68]
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

Zhao, D.; Lu, H.; Cheng, Q.; Huang, Q.; Ai, J.; Zhang, Z.; Liu, H.; He, Z.; Li, Q. Research Progress on Inactivation of Bacteriophages by Visible-Light Photocatalytic Composite Materials: A Mini Review. Materials 2024, 17, 44. https://doi.org/10.3390/ma17010044

AMA Style

Zhao D, Lu H, Cheng Q, Huang Q, Ai J, Zhang Z, Liu H, He Z, Li Q. Research Progress on Inactivation of Bacteriophages by Visible-Light Photocatalytic Composite Materials: A Mini Review. Materials. 2024; 17(1):44. https://doi.org/10.3390/ma17010044

Chicago/Turabian Style

Zhao, Deqiang, Heng Lu, Qingkong Cheng, Qi Huang, Jing Ai, Zhibo Zhang, Hainan Liu, Zongfei He, and Qiuhong Li. 2024. "Research Progress on Inactivation of Bacteriophages by Visible-Light Photocatalytic Composite Materials: A Mini Review" Materials 17, no. 1: 44. https://doi.org/10.3390/ma17010044

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