Next Article in Journal
Lavandula pedunculata subsp. atlantica: A Multifunctional Essential Oil for Potentially Combating Microbial Infections and Inflammatory Processes
Previous Article in Journal
Harvesting Friction Energy on Zinc Oxide and Zinc Oxide/Europium Oxide Sol-Gel Catalysts for Tribocatalytic Paracetamol Degradation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Influence Mechanism of Dissolved Organic Matter on the Photocatalytic Oxidation of Pharmaceuticals and Personal Care Products

Beijing Key Laboratory of Resource-Oriented Treatment of Industrial Pollutants, School of Energy and Environmental Engineering, University of Science and Technology Beijing, Beijing 100083, China
*
Author to whom correspondence should be addressed.
Molecules 2025, 30(11), 2266; https://doi.org/10.3390/molecules30112266
Submission received: 7 May 2025 / Revised: 21 May 2025 / Accepted: 21 May 2025 / Published: 22 May 2025
(This article belongs to the Special Issue Advanced Oxidation of Emerging Pollutants in Water)

Abstract

With the worsening global water pollution crisis, pharmaceuticals and personal care products (PPCPs) have been increasingly detected in aquatic environments. The effective removal of PPCPs remains challenging for conventional water treatment technologies, whereas photocatalytic technology has shown distinct promise. Dissolved organic matter (DOM), a ubiquitous component of aquatic ecosystems, exerts multifaceted effects on the photocatalytic oxidation of PPCPs. In this article, the influence of DOM on the performance of various photocatalysts in PPCP removal is systematically summarized and analyzed. This review highlights DOM’s role in altering the migration and transformation of PPCPs via processes including adsorption and complexation. The adsorption of PPCPs on photocatalysts is achieved by competitive adsorption or by providing more adsorption sites. DOM modifies the structural properties of photocatalysts through mechanisms such as ligand exchange, intermolecular forces, electrostatic forces, and hydrophobic interactions. DOM inhibits the formation of active species via light attenuation and shielding effects while simultaneously enhancing their generation through photosensitization and electron transfer facilitation. In this review, the interaction mechanism among DOM, PPCPs, and photocatalysts within the PPCP photocatalytic oxidation system is expounded on. These findings provide novel insights into optimizing photocatalytic reaction conditions and enhancing treatment efficiency, while providing a theoretical foundation for advancing efficient, eco-friendly PPCPs remediation technologies.

Graphical Abstract

1. Introduction

Globally, water pollution has emerged as a critical challenge threatening human health and ecological balance. Water pollution not only compromises the security of water for daily use but also profoundly impacts aquatic ecosystems and biodiversity. Pharmaceuticals and personal care products (PPCPs) have emerged as a globally prevalent class of contaminants detected in aquatic systems. In recent years, PPCPs have been frequently detected at elevated concentrations in aquatic environments, posing significant ecological risks to aquatic organisms and human health [1]. For instance, in rural Czech wastewater systems, the median mass concentrations of caffeine, ibuprofen, and paracetamol exceeded 10,000 ng·L−1, with diclofenac, atenolol, and metoprolol concentrations surpassing 1000 ng·L−1 [2]. Similarly, elevated concentrations of N,N-Diethyl-3-methyl benzoyl amide (DEET), sulfamethoxazole, and tramadol in Jakarta Port, Indonesia, were also at a relatively high level [3]. In China’s Yangtze River, DEET concentrations in surface water have reached 166 ng·L−1 [4], while oxytetracycline levels in the Pearl River Basin peaked at 2030 ng·L−1 [5]. The sources of DEET in rivers mainly include the discharge of domestic sewage and industrial wastewater, and the runoff and infiltration of agricultural, forestry, and horticultural wastewater. Oxytetracycline belongs to the tetracycline class of antibiotics, which can be discharged into rivers through medical wastewater, agricultural runoff, and pharmaceutical wastewater. In Tangxun Lake and East Lake, bisphenol A and estrone were identified as high-risk contaminants, whereas triclosan and estriol exhibited moderate risks in certain areas [6].
Given the extensive pollution problems caused by PPCPs, traditional water treatment technologies struggle to completely remove these complex organic pollutants. As an advanced oxidation technology, solar-driven photocatalysis demonstrates unique advantages in the treatment of PPCPs due to its low energy consumption, environmental friendliness [7], and high yield of reactive oxygen species (ROS) [8]. Semiconductor-based photocatalysts leverage their optoelectronic properties to enable energy-efficient pollution control. For instance, S-C3N5 achieved a sulfamethoxazole degradation efficiency exceeding 90% after 30 min of visible light exposure [9], while Bi/W18O49 demonstrated removal rates of 93%, 87%, and 62% for trimethoprim, acetaminophen, and tetracycline, respectively, under 610 nm red light irradiation [10]. Similarly, TiO2 nano-dispersions degraded tetracycline by over 90% within minutes under both visible and ultraviolet light [11], demonstrating the favorable effect of photocatalytic technology in degrading PPCPs.
Dissolved organic matter (DOM), ubiquitous in natural aquatic environments, comprises organic substances that dissolve in water and can pass through a 0.45 μm filtration membrane [12]. The primary components of DOM in surface water include humic acids, polysaccharides, and proteins [13], along with small molecular organic acids and other components [12]. In practical research, DOM concentrations are typically quantified using dissolved organic carbon (DOC). In surface water and groundwater, DOM concentrations typically range from 1 to 10 mg·L−1. In most treated wastewater, concentrations generally range between 5 and 30 mg·L−1 [14], and the overall concentration range in the water environment is 0.5–100 mg·L−1 [12]. Given that DOM concentrations far exceed those of PPCPs in aquatic environments, DOM inevitably mediates the interactions between PPCPs and particulate matter or sediments.
Owing to their abundant functional groups and dual ionization structures, PPCPs exhibit a strong tendency to bind with DOM in aquatic environments, significantly affecting their migration, transformation, and removal during the wastewater treatment process [15]. This interaction not only alters the physicochemical forms and stability of PPCPs but also exerts a multifaceted influence on their degradation efficiency in photocatalytic systems. As a natural organic component in the water, DOM possesses diverse functional groups and reactive sites that promote strong interactions with PPCP molecules, resulting in stable complexes.
In addition to the interaction with PPCPs, some components in DOM, especially organic compounds rich in functional groups such as carboxyl and hydroxyl groups, can bind to photocatalysts’ surfaces via electrostatic interactions, hydrogen bonds, or coordination bonds. Such binding can occupy or block active sites on photocatalyst surfaces, limiting PPCPs’ access to reactive sites and thereby diminishing photocatalytic efficiency. Furthermore, DOM adsorption on photocatalyst surfaces can alter surface charge distribution and energy band structures, impacting light absorption and photogenerated charge carrier separation efficiency, ultimately modulating the kinetics of photocatalytic reactions.
Under illumination, certain DOM fractions exhibit photosensitization, generating reactive species such as excited triplet states (3DOM*), hydroxyl radicals, and superoxide radicals, which promote the degradation of PPCPs. Conversely, studies have demonstrated that DOM can inhibit the photodegradation of organic substances through light screening, scavenging of reactive species, and reaction intermediates [16,17]. For instance, DOM acting as a quencher of photoexcited molecules has been shown to reduce the degradation rates of PPCPs (e.g., cimetidine, carbamazepine, and propranolol) by photocatalysts like TiO2 and ZnO [18,19]. DOM has been observed to inhibit PPCPs’ degradation under visible or UV light irradiation through light shielding and light attenuation mechanisms. For example, DOM reduces the degradation of metronidazole by UV/TiO2 and the photocatalytic degradation process of organic substances such as propranolol, diclofenac, and carbamazepine by TiO2 [20,21].
This paper summarizes the influence of DOM on the photocatalytic oxidation of PPCPs. The impacts of DOM on the migration and transformation of PPCPs, the surface and interface structures of photocatalytic materials, and the reactive species in the reaction system are reviewed. The paper elucidated the mechanism of how DOM influences the photocatalytic oxidation of PPCPs and identified current challenges that require addressing.

2. Effect of DOM on the Photocatalytic Oxidation of PPCPs

Photocatalytic materials are a category of catalytic materials that can be excited to generate electron–hole pairs under light illumination and achieve environmental purification through oxidation-reduction reactions. Based on their composition and structural characteristics, these materials can be classified into metal oxide-based, carbon-based, metal sulfide-based, bismuth-based, and composite photocatalytic materials and newly emerged nanocatalytic materials. The structural characteristics of several common photocatalytic materials are shown in Figure 1.
The effect of DOM on the degradation of PPCPs by photocatalytic materials is complex and diverse. Table 1 lists the degrees of influence of DOM on the degradation of PPCPs by different types of photocatalytic materials.

2.1. Metal Oxide-Based Photocatalysts

In recent years, metal oxide-based photocatalysts have garnered significant attention from researchers in the field of organic compound degradation. This is attributed to their exceptional light absorption capabilities under both visible and ultraviolet light irradiation and their favorable safety and stability properties [70]. However, the photocatalytic performance of metal oxide-based photocatalysts is often limited by factors such as the rapid recombination of electron–hole pairs generated by light, limited light absorption in the visible spectrum, and low surface area. To enhance the performance of metal oxide-based photocatalysts, various modification methods can be adopted. At present, the most common modification methods include doping metals [71] or non-metals [72], sensitizing with dyes [73] and quantum dots [74], constructing heterojunctions [75], and loading co-catalysts (noble [76] or non-noble metals [77]), etc. These modification methods can significantly improve the performance of metal oxide-based photocatalysts.

2.1.1. TiO2

TiO2 is widely recognized as one of the most effective semiconductor photocatalysts, offering advantages such as high chemical stability, strong photocatalytic activity, non-toxicity, and low cost. DOM has a significant influence on the catalytic effect of TiO2-based photocatalysts. For instance, the presence of 10 mg·L−1 humic acid (HA) can decrease the reaction rate constant (k) for the degradation of bupropion (BUP) by TiO2 from 7.5 × 10−2 min−1 to 1.5 × 10−2 min−1, resulting in an inhibitory effect of nearly 80% [31]. This phenomenon occurs because DOM can interfere with the decomposition of TiO2 to target pollutants through occupying surface sites, scavenging free radicals, holes, and absorbing ultraviolet light [78]. However, DOM exhibits minimal impact on the catalytic effect of TiO2 nanotube arrays (TiO2 TNAs). This resistance is attributed to their larger specific surface area, which provides more active sites.
The influence of DOM on the degradation of organic matter by TiO2 is strongly related to the crystal form of TiO2 [32]. There are seven crystal forms of TiO2, but only rutile, anatase, and brookite exist in nature [79]. The most commonly used crystal form of TiO2 is anatase. Compared to rutile, which has four Ti–O bonds, anatase exhibits more pronounced octahedral distortions. Anatase, with two Ti–O bonds, is more prone to defect formation, thus generating more electron holes. Additionally, anatase possesses a larger surface area [80], which enhances the efficiency of the photocatalytic reaction [70]. Lee et al. [32] investigated the influence of HA on anatase and rutile. During the degradation of 4-chlorophenol (4-CP), anatase generates ·OH through the O2 reduction pathway, which is inhibited by the competitive adsorption of HA and H2O2. Rutile primarily generates ·OH through the oxidation of H2O. When the HA concentration is below 20 mg·L−1, HA facilitates greater electron participation in the reduction of O2, enhancing the yield of ·O2 and promoting the degradation of 4-CP.

2.1.2. ZnO

ZnO, a nano-semiconductor material, is extensively utilized in solar cells, lithium-ion batteries, and photocatalysis [81]. It has been widely employed for degrading environmental pollutants in air and water, and converting selective organic pollutants into non-toxic small molecules or even into CO2 or H2O [82]. ZnO is a wide-bandgap semiconductor that can only absorb ultraviolet light to generate photoelectrons and holes [81]. As a member of the II-VI group, ZnO has an optical bandgap of 3.26 eV. Owing to its tunable band structure, it has emerged as a promising alternative and photocatalyst [83].
During the process of ZnO photocatalytic degradation of tetracycline (TC) by ZnO under visible light, the addition of HA resulted in a 19% reduction in TC degradation efficiency [34]. The study by Li et al. [84] demonstrated that a specific concentration of HA (5 mg·L−1) significantly inhibits ·OH generation by ZnO under UV irradiation. Complexation between Zn2+ released into the solution and HA deactivates the triplet state of HA, thus hindering the degradation of organic pollutants by ZnO.
Compared with ZnO, ZnO nanomaterials are more significantly affected by DOM. The introduction of DOM caused the removal rate of methylene blue (MB) by ZnO nanoparticles (ZnO NPs) to decrease by 79.4%, and it also caused the removal rate of cefalexin by ZnO nanowires to drop sharply from 100% to 40.9% [36]. This strong inhibitory effect is attributed not only to the quenching of active free radicals by DOM, the attenuation and consumption of light, and the occupation of the active sites of ZnO nanomaterials [82], but also the ability of HA to promote the dissolution of ZnO nanoparticles, with solubility increasing as the HA concentration rises [85]. Therefore, the catalytic degradation efficiency of ZnO nanomaterials is significantly impacted.

2.1.3. WO3

As an n-type semiconductor, WO3 is extensively employed as a photocatalyst for pollutant degradation owing to its high chemical stability, electron mobility [59], relatively narrow bandgap (2.6–2.8 eV [86]), and strong oxidation-reduction potential [87]. However, WO3 is characterized by a small specific surface area [88], a high carrier recombination rate, and a low quantum yield. Therefore, coupling WO3 with functional materials to form a heterostructure or doping and modifying are effective strategies for enhancing its photocatalytic performance [50,89].
Yazdanbakhsh et al. [37] doped Mn into WO3 and investigated the influence of varying HA concentrations on the degradation of diclofenac (DCF) by Mn-WO3. The results showed that the inhibitory effect of HA on the reaction system became increasingly pronounced as its concentration increased. Xu et al. synthesized a WO3/ZnIn2S4 Z-scheme heterojunction and observed that the inhibitory effect (53.9%) of the addition of HA on the degradation of tetrabromobisphenol A (TBBPA) was significantly higher than that of inorganic anions, including Cl, HCO3, and SO42−. This phenomenon can be attributed to the scavenging effect of HA on free radicals and its light shielding effect, both of which hinder pollutant degradation [59].

2.2. Carbon-Based Photocatalysts

Carbon-based materials possess a large surface area, high mechanical strength, excellent electron mobility, thermal conductivity, and carrier mobility, making them essential for the removal of organic pollutants in water [90].

2.2.1. Graphitic Carbon Nitride (g-C3N4)

The polymer semiconductor g-C3N4 is one of the carbon-based photocatalytic materials that have been extensively investigated in recent years. It exhibits a relatively large specific surface area, a stable energy gap (~2.7 eV), excellent electronic properties, and rich functional groups with abundant surface defects. Therefore, it has been widely used in the removal of environmental pollutants [39].
The influence of DOM on the degradation efficiency of g-C3N4-based photocatalytic materials is also dualistic. Due to the reaction between DOM and photogenerated holes (h+) or the competitive adsorption with PPCPs, the addition of DOM can inhibit the degradation of the target pollutants. For instance, the addition of HA resulted in an 18% reduction in the degradation rate of phenanthrene (PHE) by g-C3N4 [39]. Conversely, Meng et al. [38] found that the presence of HA has an obvious promoting effect on the degradation of carbamazepine (CBZ), with the promoting effect being positively correlated with HA concentration. This phenomenon may be attributed to the coadsorption or accumulation of HA and CBZ on the catalyst, promoting its degradation.

2.2.2. Graphene

Graphene, a honeycomb structure of six-membered rings formed by carbon atoms through sp2 hybrid orbitals [91], exhibits exceptional electrical conductivity, thermal conductivity, and a unique quantum tunneling effect [92]. Consequently, the integration of photocatalysts with graphene can further promote the migration of photogenerated electrons and significantly improve the performance of photocatalysts [93].
Zou et al. [40] investigated the influence of fulvic acid (FA) on the degradation of acetaminophen (APAP) by graphene oxide (GO) and found it to be significant. The addition of 20 mg·L−1 FA reduced the reaction rate constant from 0.4547 min−1 to 0.0689 min−1. This reduction is attributed to the consumption of photogenerated holes and radicals during the photodegradation of FA and the competitive adsorption and oxidation of FA and APAP on the surface of reduced graphene oxide (RGO), which collectively inhibit the reaction [94]. However, Chen et al. [52] demonstrated that a low concentration of FA can generate reactive species, such as 1O2 and 3FA*, under visible light irradiation, thereby promoting the degradation of DCF. In contrast, a high concentration of FA produced a light shielding effect, which inhibited DCF degradation.

2.2.3. Other Carbon-Based Photocatalysts

Fullerenes and carbon nanotubes (CNTs) are widely utilized carbon-based photocatalysts. A fullerene is a hollow molecule composed exclusively of carbon atoms. Based on the total carbon atom count, fullerenes are classified as C20 [95], C60 [96], C70 [97], etc. They exhibit strong UV absorption alongside relatively weak visible light absorption [97]. C60 has exceptional redox performance due to its high electronegativity and strong antioxidant capacity, enabling efficient electron acceptance [98]. In addition, fullerenes are often used in combination with various wide bandgap semiconductor photocatalysts to achieve the purpose of efficient degradation of pollutants, such as TiO2 [99] and ZnO [100].
CNTs are tubular nanostructures formed by rolling single- or multi-layer graphene sheets. These carbon allotropes exhibit high aspect ratios, unique electronic properties, and surface functionalization potential. Based on layer configuration, they are categorized into single-walled carbon nanotubes (SWCNTs) [101] and multi-walled carbon nanotubes (MWCNTs) [102]. SWCNTs can exhibit metallic or semiconductor properties, while MWCNTs have a more complex band structure due to their multilayer structure.
Li et al. [103] employed electron spin resonance spectroscopy (EPR) to investigate the effects of HA under light on the induction of singlet oxygen and hydroxyl radicals by typical carbon nanomaterials. In the co-presence with HA, both fullerenes and CNTs significantly induce the production of ·O2, and fullerenes can also photosensitively generate ·OH. The ability to cooperatively produce 1O2 is as follows: SWCNTs > fullerenes > MWCNTs. Another report from Zhang et al. [104] reported that HA adsorption on the C60 surface promotes the dispersion through electrostatic repulsion and steric stabilization. This uniform dispersion enhances catalyst–pollutant contact, thereby accelerating pollutant degradation. These findings suggested that the presence of DOM may promote the oxidation effect of such carbon-based photocatalysts on PPCPs, though the effect varies among composite systems.

2.3. Metal Sulfide-Based Photocatalysts

Compared to metal oxides, metal sulfides possess fewer valence bands in the sp3 orbitals, resulting in a broader light response range and a higher carrier concentration [105].
Cadmium sulfide (CdS), a metal sulfide semiconductor material, exhibits a strong photoelectric effect. Due to its relatively narrow bandgap, CdS can absorb a wider range of the visible light spectrum from solar radiation, enabling its widespread application in optoelectronic devices and related fields. The bandgap of CdS is 2.4 eV [106]. When the CdS semiconductor absorbs the electromagnetic radiation from solar radiation, an excited state is formed, causing valence band electrons to transition to the conduction band. This process generates highly redox-active electron–hole pairs, which facilitate the degradation of organic pollutants in wastewater.
The influence of DOM on metal sulfide-based photocatalytic materials also has dual characteristics. Jiang et al. [41] loaded CdS onto biochar to synthesize CdS@BC for thiamethoxam (THM) degradation. Their study revealed slight HA-induced inhibition of THM degradation. Conversely, Zhang et al. [63] reported that HA at concentrations of 0–10 mg·L−1 has a slight promoting effect on the degradation of tetracycline (TC) by IS-Ni2P/CdS@g-C3N4. Similarly, Du et al. [49] demonstrated that HA enhances the degradation of 17β-estradiol (17β-E2) by Er3+-CdS/MoS2, which is attributed to the photosensitization effect of DOM.

2.4. Bismuth-Based Photocatalysts

Among photocatalytic materials, bismuth-based semiconductors have garnered significant attention due to their excellent biocompatibility, chemical stability, and relatively narrow bandgap (most < 3.0 eV) [107]. The valence bands of compounds containing Bi3+ are formed by the hybridization of O 2p and Bi 6s2 orbitals, leading to an elevation of the valence bands [107]. Bismuth-based photocatalysts primarily include bismuth oxide, bismuth oxyhalide, bimetallic oxides, etc.
For most bismuth-based photocatalytic materials, DOM shows an obvious inhibitory effect, with the inhibition rate ranging from 5% to 45%. For example, the addition of DOM resulted in a 30% reduction in indomethacin (IDM) during SrBiOI photocatalysis [42]. This phenomenon may be attributed to the light shielding effect of DOM and the consumption of active species, which significantly decrease the degradation efficiency of IDM. However, Fan et al. [44] investigated the influence of FA and HA on the photocatalytic degradation of naproxen (NPX) by bismuth titanate nanobulk (Bi-TNB) and found that low-concentration (5 mg·L−1) HA can double the degradation rate of NPX, whereas FA consistently exhibits an inhibitory effect. Liu et al. [43] synthesized molecularly imprinted BiOCl (MI-BiOCl) and observed that the addition of HA did not significantly inhibit the degradation of venlafaxine (VEN). This suggests that MI-BiOCl with imprinting sites can selectively adsorb VEN without adsorbing HA, thus effectively using radicals to degrade VEN.

2.5. Composite Photocatalysts

Composite photocatalysts include semiconductor composites [108], carbon material composites [109], metal composites [110], organic polymer composites [111], sensitizer composites [112], etc. Based on the composite method, they can be divided into binary [113] and ternary [114] photocatalysts, etc.
The influence of DOM on the photocatalytic effect of composite materials is predominantly inhibitory, although it can exhibit a promoting effect in certain cases. For example, most carbon-based composite materials under visible light irradiation will be inhibited by DOM. However, under blue light irradiation, low-concentration (≤10 mg·L−1) FA can promote the degradation of DCF by rGO/TiO2. This phenomenon is speculated to be related to the specific excitation effect of blue light on FA molecules. Additionally, during the photocatalytic degradation of TC by IS-Ni2P/CdS/CN and MIL-88B(Fe)/ZnTi-LDH, HA also plays a promoting role in degradation, which may be due to the photosensitization effect of HA or the indirect generation of ·OH accelerating the degradation of TC.

2.6. Novel Photocatalysts

Currently, some novel nanomaterials show great potential in the field of photocatalysis, primarily including metal–organic frameworks (MOFs) [115], covalent organic frameworks (COFs) [115], MXene [116], etc. These materials are characterized by their large specific surface area, easy recovery, and strong controllability.
For such catalysts, DOM still mainly exhibits an inhibitory effect. For example, Xu et al. [65] investigated the influence of DOM on the photocatalytic degradation of 2,2′,4,4′-tetrahydroxybenzophenone (BP-2) by BMOF-Ti/Zr6%. They observed that the addition of HA and FA inhibited the adsorption of BP-2 on MOF, and the inhibitory effect of FA was more obvious. The same inhibitory phenomenon also appears in the photocatalytic process of PPCPs by materials such as Zr-MOFs [66] and π-COF [67]. Since such catalytic materials have a large specific surface area, they possess strong adsorption capacities. It is hypothesized that the competitive adsorption of DOM on the material surface inhibits the degradation effect of the target pollutants.
However, in the study by Zhang et al. [68], HA promoted the degradation of bisphenol A (BPA) by ZnFe2O4-seed@TpTt-COF. This is attributed to the presence of quinone functional groups in HA, which can act as active sites, and its rich functional groups can serve as electron donors and electron shuttles [117], thereby accelerating the degradation of BPA.

2.7. Chapter Summary

Overall, the influence of DOM on the photocatalytic oxidation of PPCPs is predominantly inhibitory. This phenomenon is particularly evident in metal oxide-based, carbon-based catalysts, and their composite materials, with the highest inhibitory effect reaching up to 84.8%. In contrast, the promoting effect of DOM is mainly observed in novel photocatalytic materials, such as COFs, and modified photocatalysts that have undergone elemental doping or surface modification. These materials typically exhibit an optimized electronic structure and enhanced light absorption capacity, which may facilitate electron transfer processes when interacting with DOM. In particular, for the same catalytic material, the effect of DOM shows an obvious dependency on the concentration. For example, low concentrations of DOM can promote carrier separation through photosensitization, thereby accelerating the degradation of pollutants. Conversely, high concentrations of DOM significantly inhibit the catalytic activity due to the light shielding effect and competitive adsorption of active sites. This concentration-dependent effect reveals the dualistic mechanism of DOM on the photocatalytic oxidation of PPCPs, that is, there are both the promoting effect mediated by electron transfer and the inhibitory effect caused by light absorption competition.

3. Effect of DOM on the Migration and Transformation of PPCPs

In the past few decades, people’s awareness of the ecological risks of PPCPs has been increasingly enhanced. When PPCPs are released into the aquatic environment, it is essential to understand their migration and transformation pathways. Migration, adsorption, bioabsorption, and degradation are the main behaviors of PPCPs following their introduction into the environment, and the role of DOM in these processes cannot be ignored. The interaction mechanism between DOM and PPCPs is shown in Figure 2.

3.1. Adsorption

During the interaction between DOM and PPCPs, the influence of DOM is particularly significant for highly polar PPCPs, such as certain antibiotics like tetracyclines and sulfonamides. The polar functional groups in these antibiotic molecules, such as the amino group and carboxyl group, facilitate interactions with the polar components of DOM, such as HA and FA. The adsorption of DOM can reduce the environmental mobility of these antibiotics, limiting their diffusion in soil and water bodies, thus potentially affecting the environmental fate and ecological risks of antibiotics. The adsorption mechanisms between DOM and PPCPs include π–π electron donor–acceptor (π–π EDA) interaction, hydrogen bonding, electrostatic interaction, hydrophobic interaction, etc.
Aolin et al. [118] showed that hydrophobic interaction plays an important role in the binding of PPCPs and DOM. Compared to FA, HA has a larger molecular weight distribution, a narrower molecular size distribution, and lower polydispersity [119], and the interaction between HA and PPCPs is more obvious.
The adsorption process between DOM and PPCP molecules typically involves multiple mechanisms acting in combination. For instance, Sun et al. [120] investigated the adsorption properties of coal humic acid (CHA) and soil humic acid (SHA) toward TC, indicating that hydrogen bonding and electrostatic interaction are key mechanisms of the interaction between TC and the functional groups in HA. Similarly, Niu et al. [121] showed that the strong hydrophobic interaction and π–π EDA between PHE and straw biochar DOM (BDOM) led to a high adsorption capacity of PHE on the BDOM400 adsorbent.

3.2. Complexation

DOM can form complex structures with PPCPs through chelation and ionic bonds. Yang et al. [122] pointed out that the presence of DOM reduces the overall electrochemical potential of the solution, and the complexation of DOM with triclosan (TCS) increases its molecular weight, thereby inhibiting the migration and transport of TCS across the ion exchange membrane. With rich functional groups and a large specific surface area, humus can bind to PPCP molecules. For example, PPCPs rich in amino acids can form covalent bonds with phenolic humus [123], promoting their adsorption in soil or sediment, thereby reducing their migration. Yao et al. [123] explored the influence of DOM coupling on the migration of 33 kinds of PPCPs, indicating that higher DOM abundance and aromaticity enhance the migration rate of PPCPs at the soil–water interface through co-transport and competitive adsorption. Humus-like DOM negatively impacts PPCPs, whereas proteinaceous DOM exhibits a higher affinity for PPCPs. Proteinaceous DOM contains functional groups such as the amino group and carboxyl group, which can undergo complexation reactions with PPCPs to form larger complexes, affecting the mobility of PPCPs. Studies by Huang et al. [124] and Ma et al. [125] further confirmed that DOM with a higher molecular weight has more obvious aromaticity; contains more acidic groups, fatty acids, long aliphatic side chains, and polysaccharides; and is thus more likely to chelate with PPCP molecules.
The colloidal environment formed by polysaccharides can encapsulate certain PPCPs, thereby hindering their degradation. In addition, in aquatic environments, polysaccharide-type DOM disintegrates under microbial activity, producing acidic or alkaline substances that change the environmental pH value. For example, when the decomposition of polysaccharides acidifies the water body, some alkaline PPCPs such as sulfonamide drugs may be more likely to dissolve and migrate.

4. Interference Mechanism of DOM on the Surface and Interface Structure of Photocatalytic Materials

4.1. Effect of DOM on the Adsorption of PPCPs by Photocatalytic Materials

DOM is commonly present in aquatic environments, and its competitive adsorption with PPCPs or photocatalytic materials is a key factor influencing the adsorption of PPCPs on photocatalysts. DOM can undergo competitive adsorption on the catalysts’ surfaces through hydrophobic interactions, π–π EDA interactions, hydrogen bonding, and electrostatic interactions, thereby leading to an enhancement or weakening of the adsorption of PPCPs [126].
The micropore filling effect explains why adsorbents can efficiently adsorb low-molecular-weight compounds. Compared to macromolecular natural DOM, DOM with a smaller molecular size can compete for the adsorption sites within micropores [127], thus affecting the adsorption of PPCPs on the catalysts’ surfaces. Furthermore, more hydrophobic DOM, such as HA, can compete for adsorption sites on the catalysts’ surfaces through hydrophobic interactions [128].
Studies have demonstrated that the competitive adsorption of DOM on catalysts significantly reduces the degradation rates of PPCPs, such as fibrates, carbamazepine, and sulfamethoxazole, by photocatalysts like TiO2 and MCNT-TiO2 [21,129]. Liu et al. [130] found that the presence of HA reduces the adsorption of highly organic compounds, such as ketoprofen, carbamazepine, and BPA, on catalysts. Similarly, Liao et al. [131] pointed out that HA competes with norfloxacin (NOR) and sulfamethoxazole (SMX) for the polar groups and hydrophobic sites on the adsorbent through hydrogen bonding and π–π interactions.
Under different pH conditions, the extent of DOM influence on the adsorption of organic pollutants by catalysts differs significantly. The study by Ye et al. [132] showed that acidic conditions are favorable for the adsorption of DOM [133,134], while alkaline conditions reduce the competitive adsorption of DOM, thereby promoting the degradation of pollutants.
Conversely, DOM complexed with catalysts can also improve the adsorption of PPCPs on the catalysts’ surfaces by providing additional adsorption sites, such as oxygen-containing functional groups and aromatic rings, or by reducing the electrostatic repulsion between PPCPs and catalysts [135]. For example, DOM enhances the adsorption of TC on functionalized oxidized graphene nanoparticles [136] and the adsorption of propranolol (PRO) on MWCNTs [137]. Lin et al. [138] further showed that HA bound to the surface of MWCNTs introduces oxygen-containing functional groups and negative charges, thus significantly increasing the adsorption capacity of MWCNTs for Pb2+.

4.2. Effect of DOM on the Apparent Structure of Photocatalysts

DOM contains a variety of functional groups that can interact with the surface of photocatalysts through mechanisms such as ligand exchange, intermolecular forces, electrostatic force, and hydrophobic interaction (as shown in Figure 3), thereby influencing the photochemical properties of the photocatalysts [139].

4.2.1. Ligand Exchange

The main components of DOM include tyrosine, FA, HA, polysaccharides, lipids, proteins, and other organic substances. [140]. DOM is rich in various functional groups, including aldehyde, amino, carboxyl, ester, hydroxyl, ketone, and phenol groups [141]. The oxidized functional groups in DOM exhibit strong electrophilicity, making them prone to pairing with other functional groups on the surface of the photocatalytic materials (e.g., amino and hydroxyl groups) to form stable compounds. Notably, DOM contains a relatively large number of acidic functional groups, such as carboxyl groups and phenolic hydroxyl groups, which can undergo ligand exchange reactions with the functional groups on the surface of the photocatalytic materials [140]. For example, Li et al. [142] demonstrated that the surface complexation–ligand exchange reaction between the surface of HA and nano-TiO2 resulted in the disappearance of the HA phenolic hydroxyl groups peak at 1245 cm−1, indicating a strong interaction between the phenolic hydroxyl groups and the surface of nano-TiO2.

4.2.2. Intermolecular Forces

When DOM coexists with photocatalysts, various intermolecular forces arise between them, including Coulomb force, van der Waals forces, hydrogen bonds, and covalent bonds [13].
For instance, DOM containing a high proportion of polar functional groups, such as carboxyl and hydroxyl groups, can form strong bonds with the surface of the photocatalytic material through hydrogen bonds. By analyzing Fourier transform infrared (FTIR) spectroscopy, Dong et al. [143] observed that after HA was adsorbed on Mo-Se/OHNT, the formation of hydrogen bonds caused a significant red shift in the O–H/N–H stretching vibrations and a slight blue shift in the bending vibrations. When OHNT was calcined to remove surface hydroxyl groups and the HA adsorption experiment was repeated, no hydrogen bonds were formed. This further confirmed that hydrogen bonds formed between the hydroxyl/amino of HA and the hydroxyl of Mo-Se/OHNT, which promoted the degradation of pollutants.
Through theoretical calculations, Yang et al. [144] found that the stable adsorption of low-molecular-weight organic acids (LOAs) on TiO2 and the orbital overlap between the highest occupied molecular orbital (HOMO) of LOAs and the lowest unoccupied molecular orbital (LUMO) of TiO2 were the key factors affecting the photocatalytic reduction process of Cr (VI). Further analysis revealed that the orbital configuration between LOAs and TiO2 facilitates the formation of covalent bonds between them, enabling direct electron transfer from LOAs to the TiO2 surface, thereby effectively promoting the reduction reaction of Cr (VI).

4.2.3. Electrostatic Force

Since the carboxyl and phenolic hydroxyl groups of DOM can ionize, DOM generally carries a negative charge in natural environments [140], whereas photocatalysts such as TiO2 may carry either a positive or a negative charge in the environment [13]. As the concentration of DOM increases, the interaction between DOM and the catalyst becomes more complex. At this point, electrostatic repulsion caused by the electrostatic double layer formed during DOM adsorption on the catalyst surface becomes dominant. This electrostatic repulsion stabilizes the catalyst, enabling it to maintain relatively consistent catalytic performance despite variations in DOM concentration [145].

4.2.4. Hydrophobic Interaction

DOM containing a relatively high proportion of aromatic functional groups can adsorb onto the surface of photocatalysts through hydrophobic interaction [146]. Luo et al. [147] observed that HA and FA could form different hydrophobic–hydrophilic layers on TiO2 NPs. Furthermore, HA, with its higher molecular weight and greater abundance of hydrophobic groups, exhibits a stronger steric stabilization effect on the catalytic material. In contrast, DOM derived from pig manure, sludge, and sediments has a lower molecular weight and aromaticity, resulting in a limited impact on the catalytic material [148].

5. Effect of DOM on Active Species

Reactive oxygen species (ROS) are a class of chemically active oxygen-containing molecules or free radicals. ROS mainly includes superoxide anions (·O2), hydrogen peroxide (H2O2), hydroxyl radicals (·OH), and singlet oxygen (1O2). These substances are highly reactive due to the presence of unpaired electrons and can attack pollutant molecules. The addition of DOM can affect the generation of ROS in the photocatalytic system. However, the photosensitization effect of DOM or its combination with the photocatalyst can also generate ROS, thereby promoting the degradation of pollutants in water [149]. Figure 4 shows the way DOM affects active species.

5.1. Generation of Active Species

According to the energy band theory, semiconductor photocatalysts possess a unique electronic band structure, comprising a lower energy valence band (VB) filled with electrons and a higher energy empty conduction band (CB) [150]. The schematic diagram of the photocatalytic degradation of pollutants is shown in Figure 5. When the photon energy () is greater than or equal to the bandgap energy (Eg), the electrons (e) in the VB will be excited into the CB, leaving holes (h+) in the VB. The photogenerated holes possess strong oxidizing power and can oxidize the substances adsorbed on the catalyst surface. In an aqueous solution, H2O or OH captures the holes to generate ·OH, while the electrons react with oxygen to form ·O2. The ·O2 radicals are unstable and highly active, and they will be converted into ·OH through a series of reactions [151]. The generation of singlet oxygen can occur in various ways. For example, h+ or e can directly transfer energy to dissolved oxygen, causing it to transition from the ground state (triplet state, 3O2) to the excited state (singlet state, 1O2). Meanwhile, ·O2 generates H2O2 through protonation or disproportionation reactions, and H2O2 can decompose into 1O2.

5.1.1. Light Attenuation/Shielding of DOM

DOM competes with pollutants for photoreactive species, thereby interfering with the degradation of organic pollutants [152]. Functional groups in DOM (e.g., carbonyl, carboxyl, and aromatic rings) can absorb light energy. Moreover, as the DOM concentration increases, a light shielding effect occurs, resulting in light attenuation, which leads to a decrease in the number of photons available for the degradation process of organic pollutants. Wang et al. [34] demonstrated that increasing HA concentrations lead to intermediate accumulation on ZnO surfaces, blocking photons from reaching the catalyst surface, thus inhibiting the degradation of TC. Peng et al. [153] indicated that DOM adsorption onto the surface of TiO2 particles forms a DOM layer that covers the surface, reducing the utilization efficiency of luminous energy.

5.1.2. DOM Facilitated Electron Transfer

The interaction between DOM and the catalyst influences the generation of active species in the system, with the types and quantities of functional groups in DOM being the main factors. Generally, DOM containing more phenolic hydroxyl groups and methoxy groups is more likely to bind to the active sites on the photocatalysts’ surfaces to form a stable complex. This promotes the separation of photogenerated electrons and holes, thereby enhancing the redox capacity of the photocatalyst.
The electrons generated by DOM can be captured by the catalyst, thus facilitating the electron transfer of the photocatalytic semiconductor [154]. Wu et al. [155] grafted citric acid (CA) onto CeO2, which initiated electron transfer, accelerated the regeneration of Ce3+ sites as electron donors, and significantly enhanced the photocatalytic performance through the ligand-to-metal charge transfer. Liu et al. [156] showed that HA contains a relatively large number of fluorophores. The time-resolved fluorescence results revealed that the interaction between DOM and TiO2 facilitates electron or energy transfer from 1DOM* to TiO2. Li et al. [84] further confirmed that the generation of ·O2 in the TiO2/HA mixture was due to the reduction of O2 by the electrons transferred from the photodissociated HA to TiO2. At the same time, DOM increased the generation of 1O2 on TiO2, CuO, CeO2, and SiO2, with FA exhibiting a greater promoting effect than HA. Sun et al. [157] found that low-molecular-weight organic acids (LMWOAs) prevented the recombination of photogenerated electron–hole pairs, creating strong reducing sites ·CO2, thereby enhancing the photocatalytic degradation effect of nitrazole. Similarly, Feng et al. [158]’s research indicated that low concentrations (2–20 mg·L−1) of FA could act as an electron shuttle, thereby promoting the degradation of monochlorobenzene by Fe-Mn bimetallic sulfide. This is because the metal ions on the surface of the catalyst can form complexes with DOM. When DOM rich in quinone functional groups complexes with and adsorbs metal ions on the surface of the catalyst [159], it can exhibit the effect of promoting electron transfer, thereby accelerating the degradation of pollutants.

5.1.3. Photosensitization of DOM

When DOM exists independently in a system, it exhibits photosensitive properties and can participate in complex photochemical reactions [160]. Studies have shown that chromophoric DOM (CDOM) is the main light absorber and photosensitizer in aquatic environments at wavelengths below 500 nm. Under sunlight irradiation, CDOM can generate photoproduction reaction intermediates (PPRIs), including ·OH, 1O2, ·CO3, eaq, ·O2, H2O2, and 3DOM*, all of which contribute to the degradation of PPCPs. In particular, aromatic components in DOM, such as lignin, are relatively sensitive to solar radiation [161].
Upon light absorption, DOM is photochemically excited from the electronic ground state (S0) to the excited singlet state (1DOM*). It releases energy and returns to the ground state through fluorescence emission and non-radiative transitions or evolves into the triplet state via intersystem crossing (ISC) under favorable conditions, thus forming 3DOM* [141]. 3DOM* typically participates in pollutant degradation through electron or hydrogen atom capture or energy transfer, effectively degrading pollutants such as phenols, phenylurea herbicides, and sulfonamide antibiotics. Additionally, the generated 3DOM* can activate oxygen through energy transfer to produce 1O2 [162]. The reaction with 3DOM* represents a primary photodegradation pathway for certain PPCPs, such as sulfonamides, β-blockers, and macrolides [163]. Liu et al. [164] studied Suwannee River natural organic matter (NOM) and found that terrestrial humus in the low-to-medium molecular weight (L-MW DOM) fraction, characterized by higher humification and fluorophore content, generates more 3DOM* and 1O2. These two reactive intermediates (RIs) play a significant role in the photodegradation of quinolones. Furthermore, 3DOM* is also crucial in the photochemical transformation of amoxicillin in natural water [165].
Under visible light irradiation, low concentrations of FA (<10 mg·L−1) generate 1O2 and 3FA*, which strongly promote the degradation of DCF in the vis-RGO/TiO2/persulfate system [52]. Cheng et al. [166] showed that, compared to PPCPs containing electron-withdrawing groups (e.g., -NOR, -COOR, -OCR), PPCPs with multiple electron-donating groups (e.g., -OH, -NH2, -OR) exhibit higher reactivity with photochemical reaction intermediates such as 3DOM* and 1O2.

5.2. Quenching of Active Species

The active groups contained in DOM, such as aromatic amines, alkene compounds, and aromatic alcohols, are capable of interacting with ROS, such as 1O2. These chemical constituents may undergo redox reactions with ROS through electron transfer pathways, subsequently mediating ROS scavenging mechanisms [167].
Qiao et al. [39] demonstrated that HA reacted with photogenerated holes, which inhibited the generation of ·OH. Generally speaking, DOM can inhibit the generation of ·OH by photocatalysts such as TiO2, ZnO, and Fe2O3. Comparative analyses reveal that FA exhibits stronger inhibitory effects than HA [84]. This is attributable to the second-order rate constant of the reaction between ·OH and FA (2.7 × 104 s1(mg of C·L−1)−1) is approximately 1.7 times higher than that of the reaction between ·OH and HA (1.9 × 104 s−1(mg of C·L−1)−1) [168]. Therefore, the presence of DOM can reduce the utilization of luminous energy in the photocatalytic system and decrease the generation of ROS and h+. Generally, the light shielding effect of DOM depends on the wavelength of light, and this effect is more obvious in the ultraviolet range with shorter wavelengths [33].

6. Conclusions and Future Perspectives

At present, research efforts on the photocatalytic oxidation of PPCPs primarily focus on the modification or development of photocatalysts and the influence of environmental factors and water matrices, including anions, cations, DOM, and so on. This review synthesized the influence of DOM on the efficiency of PPCPs’ photocatalytic oxidation and reviewed its impact on PPCP migration and transformation, the surface and interface structure of photocatalytic materials, and the active species in the reaction system. The main conclusions are as follows: DOM affects PPCPs migration and transformation through adsorption, complexation, and other effects. DOM affects the adsorption of PPCPs on the photocatalysts’ surfaces through competitive adsorption or by providing more adsorption sites. DOM affects the apparent structure of photocatalysts through ligand exchange, intermolecular forces, electrostatic forces, and hydrophobic interactions. DOM inhibits the generation of active species through light attenuation/shielding and also promotes the generation of active species through photosensitization and the promotion of electron transfer.
The influence of DOM on the photocatalytic degradation of PPCPs is relatively complex, and there are some drawbacks in the current research:
  • Most of the existing studies focus on the influence of DOM on the degradation efficiency of PPCPs, but there are few studies on how DOM regulates key processes such as the generation pathway of reactive oxygen species and the separation efficiency of photogenerated electron–hole pairs.
  • The types of DOM are complex, such as HA, polysaccharides, and proteins, and the influence of different components on photocatalysis varies significantly. However, most of the current studies use a single standard DOM, such as HA, for simulation experiments, which has a large deviation from the DOM composition in the actual environment, resulting in limited universality of the conclusions.
  • DOM participating in the photocatalytic process may generate more toxic intermediate products, such as halogenated by-products, but most of the related studies focus on the short-term degradation effect, and there is insufficient evaluation of the stability of the catalyst and the ecological risk of degradation products during long-term operation.
Future research can be carried out in the following aspects:
  • Future research should combine in situ characterization techniques, such as in situ fluorescence spectroscopy, electron paramagnetic resonance, and theoretical calculations, such as density functional theory, and other methods to systematically explore the influence of DOM on the generation pathway of reactive oxygen species and the mechanism of DOM in the separation of photogenerated electron–hole pairs.
  • Using high-throughput sequencing, mass spectrometry analysis, and other technical means to accurately quantify different components in DOM and evaluate their respective influences on the photocatalytic process. By comparing the results of simulation experiments with a single standard DOM, such as HA, and the actual DOM composition, improve the universality and accuracy of the results.
  • By extending the experimental cycle and using various ecological toxicity testing methods, comprehensively evaluate the potential impact of DOM participating in the photocatalytic process on the aquatic ecosystem.

Author Contributions

Writing—original draft preparation, validation, formal analysis, J.W.; writing—original draft preparation, M.Z. and A.S.; writing—review and editing, supervision, methodology, conceptualization, project administration, R.Y.; writing—review and editing, supervision, H.C. resources, B.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the “Interdisciplinary Program for Young Teachers” of the University of Science and Technology Beijing, grant number FRF-IDRY-23-040.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
APAPAcetaminophen (paracetamol)
BCBenzocaine
BDOMBiochar dissolved organic matter
Bi-TNBBismuth titanate nanobulk
BP-22,2′,4,4′-Tetrehydroxybenzophenone
BPABisphenol A
BUPBupropion
CACitric acid
CBZCarbamazepine
CDOMChromophoric dissolved organic matter
CHACoal humic acid
COFsCovalent organic frameworks
CPXCephalexin
CTDClothianidin
DCFDiclofenac
DEETN,N-Diethyl-3-methyl benzoyl amide
DOCDissolved organic carbon
DOMDissolved organic matter
FAFulvic acid
FQNsFluoroquinolones
GOGraphene oxide
HAHumic acid
IDMIndomethacin
LOAsLow-molecular-weight organic acids
MWCNTsMulti-walled carbon nanotubes
NOMNatural organic matter
NORNorfloxacin
NPXNaproxen
NPsNanoparticles
OFXOfloxacin
PHEPhenanthrene
PPCPsPharmaceuticals and personal care products
PPRIsPhotoproduction reaction intermediates
rGOReduced graphene oxide
RIsReactive intermediates
ROSReactive oxidative species
SHASoil humic acid
SMXSulfamethoxazole
SSZSulfasalazine
STZSulfathiazole
TBBPATetrabromobisphenol A
TCTetracycline
TCSTriclosan
THMTrimethoprim
UVUltraviolet
VENVenlafaxine
visVisible light
17β-E217β-Estradiol
4-CP4-chlorophenol
π-π EDAπ-π electron donor–acceptor

References

  1. Chen, X.; Lu, X.; Zhao, R.; Su, G.; Meng, J.; Li, Q.; Hua, Y.; Shi, B. Occurrence and risks of PPCPs of a typical mountainous region: Implications for sustainable urban water systems. Sci. Total Environ. 2024, 951, 175714. [Google Scholar] [CrossRef] [PubMed]
  2. Chen, Y.; Vymazal, J.; Březinová, T.; Koželuh, M.; Kule, L.; Huang, J.; Chen, Z. Occurrence, removal and environmental risk assessment of pharmaceuticals and personal care products in rural wastewater treatment wetlands. Sci. Total Environ. 2016, 566–567, 1660–1669. [Google Scholar] [CrossRef] [PubMed]
  3. Sudaryanto, A.; Witama, R.O.; Nosaki, K.; Tanoue, R.; Suciati, F.; Sachoemar, S.I.; Hayami, Y.; Morimoto, A.; Nomiyama, K.; Kunisue, T. Occurrence of emerging contaminants in Jakarta Bay, Indonesia: Pharmaceuticals and personal care products. Iop Conf. Series Earth Environ. Sci. 2023, 1137, 12050. [Google Scholar] [CrossRef]
  4. Liu, W.; Zhao, J.; Liu, Y.; Chen, Z.; Yang, Y.; Zhang, Q.; Ying, G. Biocides in the Yangtze River of China: Spatiotemporal distribution, mass load and risk assessment. Environ. Pollut. 2015, 200, 53–63. [Google Scholar] [CrossRef] [PubMed]
  5. Gao, F.; He, L.; Hu, L.; Chen, J.; Yang, Y.; He, L.; Bai, H.; Liu, Y.; Zhao, J.; Ying, G. The variations of antibiotics and antibiotic resistance genes in two subtropical large river basins of south China: Anthropogenic impacts and environmental risks. Environ. Pollut. 2022, 312, 119978. [Google Scholar] [CrossRef]
  6. Yi, J.; Huang, X.; Hou, J.; Xiong, J.; Qian, Z.; Liu, S.; Zhang, J.; Yin, D.; Li, J.; Su, Q.; et al. Occurrence and distribution of PPCPs in water from two largest urban lakes of China: First perspective from DGT in-situ measurement. Sci. Total Environ. 2023, 904, 166656. [Google Scholar] [CrossRef]
  7. Roshan, R.; Nahak, B.K.; Mahata, D.; Yadav, P.; Panda, S.; Patra, S.; Mahato, S.S.; Tiwari, A.; Mahata, S. Photocatalytic waste-to-renewable energy nexus using solar light induced quantum dots. Energy Conv. Manag. 2023, 283, 116917. [Google Scholar] [CrossRef]
  8. Muhammad, E.; Jan, T.; Jalil, A. Experimental and computational study of Bi2O3/WO3 heterostructures as efficient solar driven photocatalyst. Inorg. Chem. Commun. 2025, 171, 113581. [Google Scholar] [CrossRef]
  9. Wang, W.; Zhou, N.; Li, Y.; Zhu, Z.; Chen, X.; Zhang, L. S doped g-C3N5 for highly efficient photocatalytic H2O2 production and PPCPs degradation via directional transfer of electrons. J. Environ. Chem. Eng. 2023, 11, 111491. [Google Scholar] [CrossRef]
  10. Zhao, L.; Liu, Y.; Xi, X.; Nie, Z.; Lim, F.Y.; Ong, S.L.; Hu, J. Design of dual-LSPR Bi/W18O49 composite and its red-light driven photocatalytic property in the photocatalytic degradation of PPCPs. Chem. Eng. J. 2024, 481, 148119. [Google Scholar] [CrossRef]
  11. Feng, Q.; Guo, S.; Bai, Z.; Pu, Y.; Zhang, H.; Wang, J. Highly efficient photocatalytic degradation of tetracycline antibiotic enabled by TiO2 nanodispersion. J. Ind. Eng. Chem. 2024, 145, 755–763. [Google Scholar] [CrossRef]
  12. Guo, Y.; Peng, B.; Liao, J.; Cao, W.; Liu, Y.; Nie, X.; Li, Z.; Ouyang, R. Recent advances in the role of dissolved organic matter during antibiotics photodegradation in the aquatic environment. Sci. Total Environ. 2024, 916, 170101. [Google Scholar] [CrossRef] [PubMed]
  13. Philippe, A.; Schaumann, G.E. Interactions of dissolved organic matter with natural and engineered inorganic colloids: A review. Environ. Sci. Technol. 2014, 48, 8946–8962. [Google Scholar] [CrossRef]
  14. Yang, X.; Rosario-Ortiz, F.L.; Lei, Y.; Pan, Y.; Lei, X.; Westerhoff, P. Multiple Roles of Dissolved Organic Matter in Advanced Oxidation Processes. Environ. Sci. Technol. 2022, 56, 11111–11131. [Google Scholar] [CrossRef] [PubMed]
  15. Yang, B.; Cheng, X.; Zhang, Y.; Li, W.; Wang, J.; Guo, H. Probing the roles of pH and ionic strength on electrostatic binding of tetracycline by dissolved organic matters: Reevaluation of modified fitting model. Env. Sci. Ecotechnol. 2021, 8, 100133. [Google Scholar] [CrossRef]
  16. Canonica, S.; Laubscher, H. Inhibitory effect of dissolved organic matter on triplet-induced oxidation of aquatic contaminants. Photochem. Photobiol. Sci. 2008, 7, 547–551. [Google Scholar] [CrossRef]
  17. Wenk, J.; von Gunten, U.; Canonica, S. Effect of Dissolved Organic Matter on the Transformation of Contaminants Induced by Excited Triplet States and the Hydroxyl Radical. Environ. Sci. Technol. 2011, 45, 1334–1340. [Google Scholar] [CrossRef]
  18. Maeng, S.K.; Cho, K.; Jeong, B.; Lee, J.; Lee, Y.; Lee, C.; Choi, K.J.; Hong, S.W. Substrate-immobilized electrospun TiO2 nanofibers for photocatalytic degradation of pharmaceuticals: The effects of pH and dissolved organic matter characteristics. Water Res. 2015, 86, 25–34. [Google Scholar] [CrossRef]
  19. Haroune, L.; Salaun, M.; Ménard, A.; Legault, C.Y.; Bellenger, J. Photocatalytic degradation of carbamazepine and three derivatives using TiO2 and ZnO: Effect of pH, ionic strength, and natural organic matter. Sci. Total Environ. 2014, 475, 16–22. [Google Scholar] [CrossRef]
  20. He, Y.; Sutton, N.B.; Rijnaarts, H.H.H.; Langenhoff, A.A.M. Degradation of pharmaceuticals in wastewater using immobilized TiO2 photocatalysis under simulated solar irradiation. Appl. Catal. B Environ. 2016, 182, 132–141. [Google Scholar] [CrossRef]
  21. Tran, M.L.; Fu, C.; Juang, R. Effects of water matrix components on degradation efficiency and pathways of antibiotic metronidazole by UV/TiO2 photocatalysis. J. Mol. Liq. 2019, 276, 32–38. [Google Scholar] [CrossRef]
  22. Subramani, K.; Ganeshan, J.K.S.; Srinivasan, S.; Mohanasundaram, N.V.T.; Incharoensakdi, A. Hydrothermally and green synthesized TiO2 nanoparticles with high photocatalytic textile industry dye degradation, antibacterial and antioxidant activities. J. Alloy. Compd. 2025, 1017, 179068. [Google Scholar] [CrossRef]
  23. Yan, C.; Abed, A.M.; Shaban, M.; Li, X.; Zhou, X.; Lei, G.; Abdullaev, S.; Mahariq, I. Superb photocatalytic H2 production/tetracycline pollutant degradation by synthesizing novel and recyclable ternary g-C3N4-based photocatalyst: Characterization/optimization/mechanism/toxicity assessment. J. Water Process. Eng. 2025, 69, 106642. [Google Scholar] [CrossRef]
  24. Liu, D.; Zhu, C.; Yuan, G. Design of CdS/Covalent Organic Framework as an efficient heterojunction photocatalyst for enhanced photocatalytic H2 evolution. Mater. Lett. 2024, 365, 136417. [Google Scholar] [CrossRef]
  25. Eghbali, S.; Boroujerdnia, M.; Haghighatzadeh, A. Visible-light-active Ag-doped BiVO4 nanostructures: Preparation and optical and photocatalytic studies. Phys. B 2025, 703, 416990. [Google Scholar] [CrossRef]
  26. Abdoli, M.; Cigeroglu, Z.; Acikbas, Y. Eco-friendly synthesis of ZnO/g-C3N4 using Salix aegyptiaca extract: A high-performance photocatalyst for tetracycline degradation under visible light, response surface methodology. J. Mol. Struct. 2025, 1332, 141719. [Google Scholar] [CrossRef]
  27. Jiang, S.; Zheng, G.; Sun, J.; Zhang, F.; Yu, H. SPR-promoted Bi/Bi2WO6 heterojunction with effectively enhanced visible light photocatalytic degradation of pollutants. J. Alloy Compd. 2025, 1017, 179057. [Google Scholar] [CrossRef]
  28. Zhou, P.; Han, X.; Wu, L.; Qi, M.; Shen, Y.; Nie, J.; Liu, X.; Wang, F.; Bian, L.; Liang, J. Facile construction of novel BiOBr/black-TiO2/tourmaline composites for the synergistic degradation of tetracycline in aqueous. Sep. Purif. Technol. 2025, 362, 131976. [Google Scholar] [CrossRef]
  29. Karamifar, M.; Sabbaghi, S.; Mohtaram, M.S.; Rasouli, K.; Mohsenzadeh, M.; Kamyab, H.; Derakhshandeh, A.; Dolatshah, L.; Moradi, H.; Chelliapan, S. Ultrasonic-assisted synthesis of TiO2/MWCNT/Pani nanocomposite: Photocatalyst characterization and optimization of efficient variables in the degradation of benzene via RSM-CCD. Powder Technol. 2024, 432, 119176. [Google Scholar] [CrossRef]
  30. Zhang, L.; Zhang, Y.; Jiao, Y.; Yu, J.; Wang, L.; Chen, J. SnO2-x/CD-MOF S-Scheme heterostructure photocatalysts synergized by oxygen vacancies and highly porous for rhodamine B and tetracycline degradation: Process, and mechanism analysis. Colloid. Surf. A-Physicochem. Eng. Asp. 2025, 705, 135625. [Google Scholar] [CrossRef]
  31. Evgenidou, E.; Rapti, A.; Koronaiou, L.; Petromelidou, S.; Anagnostopoulou, K.; Lambropoulou, D. Photocatalytic degradation of the antidepressant drug bupropion. Performance, water matrix effect and identification of transformation products. Sustain. Chem. Environ. 2023, 3, 100028. [Google Scholar] [CrossRef]
  32. Lee, Y.; Jae Jeong, Y.; Sun Cho, I.; Lee, C.; Park, S.; Alvarez, P.J.J. The inhibitory mechanism of humic acids on photocatalytic generation of reactive oxygen species by TiO2 depends on the crystalline phase. Chem. Eng. J. 2023, 476, 146785. [Google Scholar] [CrossRef]
  33. Ye, Y.; Feng, Y.; Bruning, H.; Yntema, D.; Rijnaarts, H.H.M. Photocatalytic degradation of metoprolol by TiO2 nanotube arrays and UV-LED: Effects of catalyst properties, operational parameters, commonly present water constituents, and photo-induced reactive species. Appl. Catal. B Environ. 2018, 220, 171–181. [Google Scholar] [CrossRef]
  34. Wang, L.; Zhu, Y.; Fu, T.; Liu, K.; Sun, J.; Yan, J.; Jiang, L.; Tong, Z.; Zhou, Z.; Zhang, H. Different water matrixes leading to noticeable differences of tetracycline removal by ZnO. Mater. Today Commun. 2023, 37, 106967. [Google Scholar] [CrossRef]
  35. Chandran, P.; Netha, S.; Sudheer, K.S. Effect of humic acid on photocatalytic activity of ZnO nanoparticles. J. Photochem. Photobiol. B 2014, 138, 155–159. [Google Scholar] [CrossRef]
  36. He, J.; Zhang, Y.; Guo, Y.; Rhodes, G.; Yeom, J.; Li, H.; Zhang, W. Photocatalytic degradation of cephalexin by ZnO nanowires under simulated sunlight: Kinetics, influencing factors, and mechanisms. Environ. Int. 2019, 132, 105105. [Google Scholar] [CrossRef]
  37. Yazdanbakhsh, A.; Eslami, A.; Massoudinejad, M.; Gholami, Z.; Sarafraz, M.; Noorimotlagh, Z.; Adiban, M.; Mirzaee, S.A. Photocatalytic degradation and dechlorination mechanism of diclofenac using heterojunction Mn-doped tungsten trioxide (Mn-WO3) nanoparticles under LED visible light from aqueous solutions. Sci. Rep. 2024, 14, 29583. [Google Scholar] [CrossRef]
  38. Meng, Y.; Li, Z.; Tan, J.; Li, J.; Wu, J.; Zhang, T.; Wang, X. Oxygen-doped porous graphitic carbon nitride in photocatalytic peroxymonosulfate activation for enhanced carbamazepine removal: Performance, influence factors and mechanisms. Chem. Eng. J. 2022, 429, 130860. [Google Scholar] [CrossRef]
  39. Qiao, M.; Fu, L.; Barcelo, D. Removal of polycyclic aromatic hydrocarbons by g-C3N4 nanosheets under visible light irradiation and effect of typical co-existence substances in river water. Process Saf. Environ. Protect. 2022, 159, 376–381. [Google Scholar] [CrossRef]
  40. Zou, Y.; Wang, W.; Wang, H.; Pan, C.; Xu, J.; Pozdnyakov, I.P.; Wu, F.; Li, J. Interaction between graphene oxide and acetaminophen in water under simulated sunlight: Implications for environmental photochemistry of PPCPs. Water Res. 2023, 228, 119364. [Google Scholar] [CrossRef]
  41. Jiang, H.; Hu, Z.; Ouyang, Y.; Ji, X.; Hu, X.; Li, T.; Ouyang, K.; Wang, P.; Wang, H.; Hu, X. Multi-phase CdS loaded on biochar for photocatalytic activation of peroxymonosulfate for thiamethoxam degradation: π-conjugation improves PMS adsorption. Sep. Purif. Technol. 2023, 326, 124842. [Google Scholar] [CrossRef]
  42. Huang, J.; Zheng, X.; Liu, Y.; Wang, F.; Li, D.; Liu, H.; Li, R.; Chen, T.; Lv, W.; Liu, G. Integration of oxygen vacancies into BiOI via a facile alkaline earth ion-doping strategy for the enhanced photocatalytic performance toward indometacin remediation. J. Hazard. Mater. 2021, 412, 125147. [Google Scholar] [CrossRef]
  43. Liu, L.; Zhi, S.; Chen, R.; Yang, Y.; Luo, C.; Liang, P.; Liu, Y.; Zhu, G. Highly selective and rapid sequential degradation of venlafaxine-class antidepressants using a novel BiOCl-based two-dimensional molecularly imprinted photocatalyst. Sep. Purif. Technol. 2025, 362, 131777. [Google Scholar] [CrossRef]
  44. Fan, G.; Ning, R.; Luo, J.; Zhang, J.; Hua, P.; Guo, Y.; Li, Z. Visible-light-driven photocatalytic degradation of naproxen by Bi-modified titanate nanobulks: Synthesis, degradation pathway and mechanism. J. Photochem. Photobiol. A Chem. 2020, 386, 112108. [Google Scholar] [CrossRef]
  45. Hojamberdiev, M.; Czech, B.; Göktaş, A.C.; Yubuta, K.; Kadirova, Z.C. SnO2@ZnS photocatalyst with enhanced photocatalytic activity for the degradation of selected pharmaceuticals and personal care products in model wastewater. J. Alloy. Compd. 2020, 827, 154339. [Google Scholar] [CrossRef]
  46. Guo, C.; Gao, S.; Lv, J.; Hou, S.; Zhang, Y.; Xu, J. Assessing the photocatalytic transformation of norfloxacin by BiOBr/iron oxides hybrid photocatalyst: Kinetics, intermediates, and influencing factors. Appl. Catal. B Environ. 2017, 205, 68–77. [Google Scholar] [CrossRef]
  47. Sun, J.; Wang, L.; Huang, T.; Liu, K.; Fu, T.; Xu, Z.; Yang, W.; Tong, Z.; Zhang, H. Floating BiOBr/Ti3C2 aerogel spheres for efficient degradation of quinolone antibiotics: Rapid oxygen transfer via triphase interface and effective charges separation by internal electric field. J. Colloid. Interface. Sci. 2025, 685, 813–825. [Google Scholar] [CrossRef]
  48. Guo, W.; Zhao, S.; Li, G.; Li, J.; Chen, F.; Chen, R. Chlorine-enhanced photocatalytic degradation of PPCPs over Bi2MoO6/(BiO)2CO3 heterostructures. J. Environ. Chem. Eng. 2021, 9, 106597. [Google Scholar] [CrossRef]
  49. Du, B.; Fan, G.; Yang, S.; Luo, J.; Wu, J.; Xu, K. Mechanistic insight into humic acid-enhanced sonophotocatalytic removal of 17β-estradiol: Formation and contribution of reactive intermediates. Environ. Res. 2023, 231, 116249. [Google Scholar] [CrossRef]
  50. Lykos, C.; Bairamis, F.; Efthymiou, C.; Konstantinou, I. Synthesis and Characterization of Composite WO3 Fibers/g-C3N4 Photocatalysts for the Removal of the Insecticide Clothianidin in Aquatic Media. Nanomaterials 2024, 14, 1045. [Google Scholar] [CrossRef]
  51. Chen, R.; Ding, S.; Fu, N.; Ren, X. Preparation of a g-C3N4/Ag3PO4 composite Z-type photocatalyst and photocatalytic degradation of Ofloxacin: Degradation performance, reaction mechanism, degradation pathway and toxicity evaluation. J. Environ. Chem. Eng. 2023, 11, 109440. [Google Scholar] [CrossRef]
  52. Chen, P.; Zhang, Q.; Shen, L.; Li, R.; Tan, C.; Chen, T.; Liu, H.; Liu, Y.; Cai, Z.; Liu, G.; et al. Insights into the synergetic mechanism of a combined vis-RGO/TiO2/peroxodisulfate system for the degradation of PPCPs: Kinetics, environmental factors and products. Chemosphere 2019, 216, 341–351. [Google Scholar] [CrossRef] [PubMed]
  53. Wu, Z.; Liu, J.; Shi, J.; Deng, H. Bi2Fe4O9/rGO nanocomposite with visible light photocatalytic performance for tetracycline degradation. Environ. Res. 2024, 249, 118361. [Google Scholar] [CrossRef]
  54. Gao, P.; Li, Z.; Feng, L.; Liu, Y.; Du, Z.; Zhang, L. Construction of novel MWCNTs/Bi4O5I2 nanosheets with enhanced adsorption and photocatalytic performance for the degradation of tetracycline: Efficiency, mechanism and regeneration. Chem. Eng. J. 2022, 429, 132398. [Google Scholar] [CrossRef]
  55. Hassani, A.; Eghbali, P.; Mahdipour, F.; Wacławek, S.; Lin, K.A.; Ghanbari, F. Insights into the synergistic role of photocatalytic activation of peroxymonosulfate by UVA-LED irradiation over CoFe2O4-rGO nanocomposite towards effective Bisphenol A degradation: Performance, mineralization, and activation mechanism. Chem. Eng. J. 2023, 453, 139556. [Google Scholar] [CrossRef]
  56. Huang, J.; Li, D.; Liu, Y.; Li, R.; Chen, P.; Liu, H.; Lv, W.; Liu, G.; Feng, Y. Ultrathin Ag2WO4-coated P-doped g-C3N4 nanosheets with remarkable photocatalytic performance for indomethacin degradation. J. Hazard. Mater. 2020, 392, 122355. [Google Scholar] [CrossRef]
  57. Gao, X.; Niu, J.; Wang, Y.; Ji, Y.; Zhang, Y. Solar photocatalytic abatement of tetracycline over phosphate oxoanion decorated Bi2WO6/polyimide composites. J. Hazard. Mater. 2021, 403, 123860. [Google Scholar] [CrossRef]
  58. Zhong, J.; Huang, J.; Liu, Y.; Li, D.; Tan, C.; Chen, P.; Liu, H.; Zheng, X.; Wen, C.; Lv, W.; et al. Construction of double-functionalized g-C3N4 heterojunction structure via optimized charge transfer for the synergistically enhanced photocatalytic degradation of sulfonamides and H2O2 production. J. Hazard. Mater. 2022, 422, 126868. [Google Scholar] [CrossRef]
  59. Xu, B.; An, Z.; Li, M.; Zhang, B.; Guo, L. Photocatalytic degradation of tetrabromobisphenol A by Z-scheme WO3/ZnIn2S4 heterojunction: Efficiency, mechanism and toxicity evaluation. J. Environ. Chem. Eng. 2023, 11, 111522. [Google Scholar] [CrossRef]
  60. Meng, J.; Zhang, X.; Yang, G.; Qin, L.; Pan, Y.; Guo, Y. Porous cyclopentadiene unit-incorporated graphitic carbon nitride nanosheets for efficient photocatalytic oxidation of recalcitrant organic micropollutants in wastewater. J. Hazard. Mater. 2023, 460, 132365. [Google Scholar] [CrossRef]
  61. Jiang, R.; Lu, G.; Yan, Z.; Wu, D.; Zhou, R.; Bao, X. Insights into a CQD-SnNb2O6/BiOCl Z-scheme system for the degradation of benzocaine: Influence factors, intermediate toxicity and photocatalytic mechanism. Chem. Eng. J. 2019, 374, 79–90. [Google Scholar] [CrossRef]
  62. Cai, M.; Liu, Y.; Dong, K.; Wang, C.; Li, S. A novel S-scheme heterojunction of Cd0.5Zn0.5S/BiOCl with oxygen defects for antibiotic norfloxacin photodegradation: Performance, mechanism, and intermediates toxicity evaluation. J. Colloid. Interface. Sci. 2023, 629, 276–286. [Google Scholar] [CrossRef] [PubMed]
  63. Zhang, Q.; Wu, T.; Che, H.; Tang, C.; Liu, B.; Ao, Y. In-situ growing Ni2P on CdS@g-C3N4 composites for highly efficient synergistically photocatalytic H2 evolution and antibiotic degradation. Surf. Interfaces 2024, 47, 104205. [Google Scholar] [CrossRef]
  64. Qiu, P.; Liu, H.; Wang, G.; Wang, C.; Huang, Z.; Huang, C.; Lai, Y.; Xing, C.; Liang, K. Dual vacancies and S-scheme BiOBr/Bi2WO6 heterojunction synergistically boost the directional transfer of photogenerated electrons for efficient photocatalytic degradation of norfloxacin. J. Water Process. Eng. 2024, 68, 106372. [Google Scholar] [CrossRef]
  65. Xu, J.; Zhu, C.; Liu, R.; Fang, Q.; Zheng, Y.; He, Z.; Song, S.; Shen, Y. Elucidating the charge redistribution in bimetallic metal–organic frameworks: High-electron-density reactive center for removal of organic micropollutants in water. Chem. Eng. J. 2023, 468, 143723. [Google Scholar] [CrossRef]
  66. Yang, P.; Wu, C.; Zhang, X.; Sun, Q.; Hou, X.; Wang, T. Enhanced photocatalytic degradation of sulfathiazole via dual-ligand Zr-MOFs: A porphyrin and pyrene-based energy transfer. Sep. Purif. Technol. 2025, 354, 128727. [Google Scholar] [CrossRef]
  67. Hu, Z.; Wang, Y.; Wang, L.; Wang, Q.; Zhang, Q.; Cui, F.; Jiang, G. Synthesis of S-type heterostructure π-COF for photocatalytic tetracycline degradation. Chem. Eng. J. 2024, 479, 147534. [Google Scholar] [CrossRef]
  68. Zhang, M.; Cao, H.; Wu, M.; Zhang, Y.; Zhu, U.; Xu, J.; Deng, J.; Zhou, T. Rational fabrication of Z-scheme heterojunction ZnFe2O4-seed@TpTt-COF for the visible-light-driven photocatalytic degradation of bisphenol A in food waste leachate boosted by primitive humic acid. Chem. Eng. J. 2024, 487, 150684. [Google Scholar] [CrossRef]
  69. Luo, Y.; Shi, G.; Yu, S.; Liu, Z.; Yin, J.; Xue, M.; Sun, Q.; Shen, F.; Li, X.; Yin, Z.; et al. Novel MIL-88B(Fe)/ZnTi-LDH high-low junctions for adsorption and photodegradation of tetracycline: Characteristics, performance, and mechanisms. Chem. Eng. J. 2023, 473, 145198. [Google Scholar] [CrossRef]
  70. Jiangbo, Y.; Jing, Y.; Jie, L.; Zhanchao, W.; Shaoping, K. Photocatalytic Removal of Antibiotics from Water. Prog. Chem. 2024, 36, 95–105. [Google Scholar]
  71. Hua, Q.; Cheng, L.; Qu, W.; Wu, D.; Li, Y.; Xu, Z.; Cao, X.; Xie, M.; Yan, X.; Dong, C.; et al. Tailoring reactant adsorption on WO3 via surface bismuth incorporation for enhanced photocatalytic degradation of acetaldehyde. Appl. Catal. A Gen. 2025, 700, 120303. [Google Scholar] [CrossRef]
  72. Parchami, K.; Habibzadehnesami, N.; Pirkarami, A.; Haghighat-Nezhad, S.; Gheitasi, N.; Nabavi, M.S.; Ghasemi, E. Enhancing NiAl-LDH@TiO2-Tiv-S photoelectrocatalyst performance for textile dye degradation and oxygen production: S-doped Ti vacancies. Environ. Res. 2025, 279, 121741. [Google Scholar] [CrossRef] [PubMed]
  73. Safaralizadeh, E.; Mahjoub, A.; Janitabardarzi, S. Visible light-induced degradation of phenolic contaminants utilizing nanoscale TiO2 and ZnO impregnated with SR 7B (SR) dye as advanced photocatalytic systems. Ceram. Int. 2025, 51, 1958–1969. [Google Scholar] [CrossRef]
  74. Chang, L.; Ahmad, N.; Zeng, G.; Ray, A.; Zhang, Y. N, S co-doped carbon quantum dots/TiO2 composite for visible-light-driven photocatalytic reduction of Cr (VI). J. Environ. Chem. Eng. 2022, 10, 108742. [Google Scholar] [CrossRef]
  75. Hu, Z.; Zhang, W.; Liu, Z.; Zhang, X.; Wang, X. Construction of CoFe2O4/TiO2@C S-scheme heterojunction and removal of phenol by activated peroxymonosulfate. J. Alloy Compd. 2025, 1022, 179992. [Google Scholar] [CrossRef]
  76. Hernandez-Laverde, M.; Morante, N.; Gutierrez, B.L.; Murcia, J.J.; Monzillo, K.; Sannino, D.; Vaiano, V. Solar Light Elimination of Bacteria, Yeast and Organic Pollutants by Effective Photocatalysts Based on Ag/Cr-TiO2 and Pd/Cr-TiO2. Nanomaterials 2024, 14, 1730. [Google Scholar] [CrossRef]
  77. Rath, P.C.; Mishra, M.; Saikia, D.; Chang, J.K.; Perng, T.; Kao, H. Facile fabrication of titania-ordered cubic mesoporous carbon composite: Effect of Ni doping on photocatalytic hydrogen generation. Int. J. Hydrogen Energy 2019, 44, 19255–19266. [Google Scholar] [CrossRef]
  78. Katz, A.; Mcdonagh, A.; Tijing, L.; Shon, H.K. Fouling and Inactivation of Titanium Dioxide-Based Photocatalytic Systems. Crit. Rev. Environ. Sci. Technol. 2015, 45, 1880–1915. [Google Scholar] [CrossRef]
  79. Singh, R.; Dutta, S. A review on H2 production through photocatalytic reactions using TiO2/TiO2 -assisted catalysts. Fuel 2018, 220, 607–620. [Google Scholar] [CrossRef]
  80. Khan, H.; Shah, M.U.H. Modification strategies of TiO2 based photocatalysts for enhanced visible light activity and energy storage ability: A review. J. Environ. Chem. Eng. 2023, 11, 111532. [Google Scholar] [CrossRef]
  81. Wang, X.; Yang, S.; Jin, P.; Peng, A.; Zhao, X.; Yang, K.; He, J. Highly efficient photocatalytic degradation of microcystin-LR with permonosulfate activated by visible-light over Z-scheme ZnO/g-C3N4: Degradation pathways and mechanism. J. Environ. Chem. Eng. 2023, 11, 111088. [Google Scholar] [CrossRef]
  82. Bai, X.; Sun, C.; Liu, D.; Luo, X.; Li, D.; Wang, J.; Wang, N.; Chang, X.; Zong, R.; Zhu, Y. Photocatalytic degradation of deoxynivalenol using graphene/ZnO hybrids in aqueous suspension. Appl. Catal. B Environ. 2017, 204, 11–20. [Google Scholar] [CrossRef]
  83. Mahesha, A.; Nagaraja, M.; Madhu, A.; Suriyamurthy, N.; Satyanarayana Reddy, S.; Al-Dossari, M.; El-Gawaad, N.S.A.; Manjunatha, S.O.; Gurushantha, K.; Srinatha, N. Chromium-doped ZnO nanoparticles synthesized via auto-combustion: Evaluation of concentration-dependent structural, band gap-narrowing effect, luminescence properties and photocatalytic activity. Ceram. Int. 2023, 49, 22890–22901. [Google Scholar] [CrossRef]
  84. Li, Y.; Niu, J.; Shang, E.; Crittenden, J.C. Influence of dissolved organic matter on photogenerated reactive oxygen species and metal-oxide nanoparticle toxicity. Water Res. 2016, 98, 9–18. [Google Scholar] [CrossRef]
  85. Jiang, C.; Aiken, G.R.; Hsu-Kim, H. Effects of Natural Organic Matter Properties on the Dissolution Kinetics of Zinc Oxide Nanoparticles. Environ. Sci. Technol. 2015, 49, 11476–11484. [Google Scholar] [CrossRef]
  86. Ghattavi, S.; Nezamzadeh-Ejhieh, A. Nanoscale AgI-WO3 binary photocatalyst: Synthesis, brief characterization, and investigation of its photocatalytic activity. Mater. Res. Bull. 2023, 158, 112085. [Google Scholar] [CrossRef]
  87. Zhang, Y.; Liu, D.; Xiong, B.; Li, J.; Li, Y.; Zhou, Y.; Yang, A.; Zhang, Q. Constructing WO3/TiO2 heterojunction with solvothermal-sintering for enhanced photocatalytic activity under visible light irradiation. Solid State Sci. 2022, 131, 106963. [Google Scholar] [CrossRef]
  88. Singla, S.; Sharma, S.; Basu, S. MoS2/WO3 heterojunction with the intensified photocatalytic performance for decomposition of organic pollutants under the broad array of solar light. J. Clean. Prod. 2021, 324, 129290. [Google Scholar] [CrossRef]
  89. Chai, Q.; Dong, J.; Yu, X.; Zhang, X.; Li, J.; Guo, S.; Yang, Y. Photocatalytic activation of oxalic acid over FeOOH loaded FeWO4/WO3 heterojunction for high-efficient degradation of tetracycline. J. Environ. Chem. Eng. 2024, 12, 111728. [Google Scholar] [CrossRef]
  90. Javaid, A.; Latif, S.; Imran, M.; Hussain, N.; Rajoka, M.S.R.; Iqbal, H.M.N.; Bilal, M. Nanohybrids-assisted photocatalytic removal of pharmaceutical pollutants to abate their toxicological effects—A review. Chemosphere 2022, 291, 133056. [Google Scholar] [CrossRef]
  91. Liu, Z.; Wang, G.; Li, Y.; Li, H.; Deng, N. Carboxymethyl-β-cyclodextrin functionalized TiO2@Fe3O4@RGO magnetic photocatalyst for efficient photocatalytic degradation of tetracycline under visible light irradiation. J. Environ. Chem. Eng. 2024, 12, 113303. [Google Scholar] [CrossRef]
  92. Mcallister, M.J.; Li, J.; Adamson, D.H.; Schniepp, H.C.; Abdala, A.A.; Liu, J.; Herrera-Alonso, M.; Milius, D.L.; Car, R.; Prud’Homme, R.K.; et al. Single Sheet Functionalized Graphene by Oxidation and Thermal Expansion of Graphite. Chem. Mat. 2007, 19, 4396–4404. [Google Scholar] [CrossRef]
  93. Ma, Y.; Wang, S.; Zheng, W.; Xue, X.; Liu, H.; Chen, S.; Zhu, Y. Preparation of N-TiO2/RGO nanocomposites through sol-gel method. Korean J. Chem. Eng. 2021, 38, 1913–1922. [Google Scholar] [CrossRef]
  94. Duan, M.; Jiang, L.; Shao, B.; Feng, C.; Yu, H.; Guo, H.; Chen, H.; Tang, W. Enhanced visible-light photocatalytic degradation activity of Ti3C2/PDIsm via π–π interaction and interfacial charge separation: Experimental and theoretical investigations. Appl. Catal. B Environ. 2021, 297, 120439. [Google Scholar] [CrossRef]
  95. Alhowyan, A.; Mahdi, W.A.; Obaidullah, A. DFT-based study of doped C20 nanostructures for efficient detection of the narcotic drug ecstasy. Diam. Relat. Mat. 2025, 152, 111873. [Google Scholar] [CrossRef]
  96. Ray, A.; Passiu, C.; Nasuda, M.; Ramakrishna, S.N.; Rossi, A.; Kuzuya, A.; Spencer, N.D.; Yamakoshi, Y. Reactive-Oxygen-Species-Mediated Surface Oxidation of Single-Molecule DNA Origami by an Atomic Force Microscope Tip-Mounted C60 Photocatalyst. ACS Nano 2021, 15, 19256–19265. [Google Scholar] [CrossRef]
  97. Cho, E.; Ciou, J.; Zheng, J.; Pan, J.; Hsiao, Y.; Lee, K.; Huang, J. Fullerene C70 decorated TiO2 nanowires for visible-light-responsive photocatalyst. Appl. Surf. Sci. 2015, 355, 536–546. [Google Scholar] [CrossRef]
  98. Pan, Y.; Liu, X.; Zhang, W.; Liu, Z.; Zeng, G.; Shao, B.; Liang, Q.; He, Q.; Yuan, X.; Huang, D.; et al. Advances in photocatalysis based on fullerene C60 and its derivatives: Properties, mechanism, synthesis, and applications. Appl. Catal. B Environ. 2020, 265, 118579. [Google Scholar] [CrossRef]
  99. Hou, J.; Lan, X.; Shi, J.; Yu, S.; Zhang, Y.; Wang, H.; Ren, C. A mild and simple method to fabricate commercial TiO2 (P25) and C60 composite for highly enhancing H2 generation. Int. J. Hydrogen Energy 2020, 45, 2852–2861. [Google Scholar] [CrossRef]
  100. Sardar, S.; Munawar, T.; Mukhtar, F.; Nadeem, M.S.; Khan, S.A.; Koc, M.; Manzoor, S.; Ashiq, M.N.; Iqbal, F. Fullerene trigged energy storage and photocatalytic ability of La2O3-ZnO@C60 core-shell nanocomposite. Mater. Sci. Eng. B 2023, 288, 116151. [Google Scholar] [CrossRef]
  101. Murakami, N.; Tango, Y.; Miyake, H.; Tajima, T.; Nishina, Y.; Kurashige, W.; Negishi, Y.; Takaguchi, Y. SWCNT Photocatalyst for Hydrogen Production from Water upon Photoexcitation of (8, 3) SWCNT at 680-nm Light. Sci. Rep. 2017, 7, 43445. [Google Scholar] [CrossRef] [PubMed]
  102. Sobahi, T.R.; Amin, M.S. Photocatalytic oxidation of atrazine using BaTiO3 -MWCNT nanocomposites under visible light. Ceram. Int. 2021, 47, 14366–14374. [Google Scholar] [CrossRef]
  103. Xun, L.; Xuehua, L.; Jingwen, C. Synergistic Photogeneration of 1O2 and ·OH by Humid Acid and Carbon Nanomaterial in Aqueous Phase. Asian J. Ecotoxicol. 2016, 11, 49–56. (In Chinese) [Google Scholar]
  104. Zhang, L.; Zhang, Y.; Lin, X.; Yang, K.; Lin, D. The role of humic acid in stabilizing fullerene (C60) suspensions. J. Zhejiang Univ. Sci. A 2014, 15, 634–642. [Google Scholar] [CrossRef]
  105. Akram, M.Y.; Hu, B.; Jia, J.; Li, C.; Dong, H.; Lu, H. The science behind the scene: Theoretical and experimental foundations of metal sulfide photocatalyst modification. Int. J. Hydrogen Energy 2024, 93, 21–42. [Google Scholar] [CrossRef]
  106. Abdullah, U.; Ali, M.; Pervaiz, E. An Inclusive Review on Recent Advancements of Cadmium Sulfide Nanostructures and its Hybrids for Photocatalytic and Electrocatalytic Applications. Mol. Catal. 2021, 508, 111575. [Google Scholar] [CrossRef]
  107. Chen, Q.; Cheng, X.; Long, H.; Rao, Y. A short review on recent progress of Bi/semiconductor photocatalysts: The role of Bi metal. Chin. Chem. Lett. 2020, 31, 2583–2590. [Google Scholar] [CrossRef]
  108. Ghosh, S.; Sahu, M. Biochar Supported Oxygen Rich TiO2-CuO Heterojunction Photocatalyst for the Degradation of Dimethyl Phthalate: Optimization, Degradation Kinetics and Mechanisms. Int. J. Environ. Res. 2025, 19, 58. [Google Scholar] [CrossRef]
  109. Mahmoudi, K.; Farzadkia, M.; Rezaei, K.R.; Sobhi, H.R.; Yeganeh, M.; Esrafili, A. Efficient removal of oxytetracycline antibiotic from aqueous media using UV/g-C3N4/Fe3O4 photocatalytic process. Heliyon 2024, 10, e30604. [Google Scholar] [CrossRef]
  110. Liu, Y.; Zheng, T.; Xu, Y.; Li, A.; Xu, H.; Gao, Y.; Wang, X.; Fujii, R.; Sasaki, S. β-carotene assisted chlorin-modified Pt/TiO2 photocatalysts for enhanced hydrogen evolution. Appl. Surf. Sci. 2025, 684, 161943. [Google Scholar] [CrossRef]
  111. Ahmad, I.; Athar, M.S.; Muneer, M.; Altass, H.M.; Felemban, R.; Ahmed, S.A. Synergistic design of a graphene oxide-mediated polyaniline/α-Fe2O3 ternary heterostructure: Advancing photocatalytic degradation and adsorption efficiency. Nanoscale 2024, 17, 3822–3836. [Google Scholar] [CrossRef] [PubMed]
  112. Awang, H.; Peppel, T.; Strunk, J. Photocatalytic Degradation of Diclofenac by Nitrogen-Doped Carbon Quantum Dot-Graphitic Carbon Nitride (CNQD). Catalysts 2023, 13, 735. [Google Scholar] [CrossRef]
  113. Dong, W.; Qi, K.; Zhang, L.; Xie, X.; Hu, J.; Mao, M.; Wang, Z.; Yang, B. Synthesis of hydrochar-based AgBr/Bi2WO6 heterojunction photocatalyst using dopant element and its loading on PVDF membrane: Study on degradation of sulfadiazine. Sep. Purif. Technol. 2025, 360, 131132. [Google Scholar] [CrossRef]
  114. Zhan, M.; Hong, Y.; Fang, Z.; Qiu, D. Magnetic recyclable visible light-driven Bi2WO6/Fe3O4/RGO for photocatalytic degradation of Microcystin-LR: Mechanism, pathway, and influencing factors. Environ. Res. 2024, 252, 118885. [Google Scholar] [CrossRef] [PubMed]
  115. Hayat, A.; Rauf, S.; Al Alwan, B.; El Jery, A.; Almuqati, N.; Melhi, S.; Amin, M.A.; Al-Hadeethi, Y.; Sohail, M.; Orooji, Y.; et al. Recent advance in MOFs and MOF-based composites: Synthesis, properties, and applications. Mater. Today Energy 2024, 41, 101542. [Google Scholar] [CrossRef]
  116. Ghamarpoor, R.; Fallah, A.; Eghbali, T. Design of bifunctional sandwich-like Co@Si/Ox-MXene nanocomposite to increase the supercapacitor properties and removal of pollutants from wastewater. J. Alloy Compd. 2024, 983, 173920. [Google Scholar] [CrossRef]
  117. Wang, X.; Muhmood, A.; Lyu, T.; Dong, R.; Liu, H.; Wu, S. Mechanisms of genuine humic acid evolution and its dynamic interaction with methane production in anaerobic digestion processes. Chem. Eng. J. 2021, 408, 127322. [Google Scholar] [CrossRef]
  118. Aolin, H.; Qin, L.; Zhu, S.; Hu, X.; Yin, D. Combined effects of pH and dissolved organic matter on the availability of pharmaceuticals and personal care products in aqueous environment. Sci. Total Environ. 2024, 929, 172637. [Google Scholar] [CrossRef]
  119. Yu, M.; He, X.; Liu, J.; Wang, Y.; Xi, B.; Li, D.; Zhang, H.; Yang, C. Characterization of isolated fractions of dissolved organic matter derived from municipal solid waste compost. Sci. Total Environ. 2018, 635, 275–283. [Google Scholar] [CrossRef]
  120. Sun, H.; Shi, X.; Mao, J.; Zhu, D. Tetracycline sorption to coal and soil humic acids: An examination of humic structural heterogeneity. Environ. Toxicol. Chem. 2010, 29, 1934–1942. [Google Scholar] [CrossRef]
  121. Niu, Y.; Wang, S.; Gao, P.; Ren, X.; Li, F.; Liu, Z.; Wang, L.; Peng, H.; Ju, S. Photo-transformation of biochar-derived dissolved organic matter and its binding with phenanthrene/9-phenanthrol: The role of functional group and pyrolysis temperature. Bioresour. Technol. 2024, 413, 131547. [Google Scholar] [CrossRef] [PubMed]
  122. Yang, F.; Liu, J.; Xie, Y.; Ren, M. Effects of DOM and cations on the diffusion migration of PPCPs through ion exchange membranes. Sep. Purif. Technol. 2025, 359, 130508. [Google Scholar] [CrossRef]
  123. Yao, W.; Qi, Y.; Han, Y.; Ge, J.; Dong, Y.; Wang, J.; Yi, Y.; Volmer, D.A.; Li, S.; Fu, P. Seasonal variation and dissolved organic matter influence on the distribution, transformation, and environmental risk of pharmaceuticals and personal care products in coastal zone: A case study of Tianjin, China. Water Res. 2024, 249, 120881. [Google Scholar] [CrossRef] [PubMed]
  124. Huang, X.; Ye, Z.; Cai, J.; Lin, L. Quantification of DOM effects on tetracyclines transport during struvite recovery from swine wastewater. Water Res. 2021, 206, 117756. [Google Scholar] [CrossRef]
  125. Ma, D.; Gao, B.; Sun, S.; Wang, Y.; Yue, Q.; Li, Q. Effects of dissolved organic matter size fractions on trihalomethanes formation in MBR effluents during chlorine disinfection. Bioresour. Technol. 2013, 136, 535–541. [Google Scholar] [CrossRef]
  126. Zhang, M.; Shen, J.; Zhong, Y.; Ding, T.; Dissanayake, P.D.; Yang, Y.; Tsang, Y.F.; Ok, Y.S. Sorption of pharmaceuticals and personal care products (PPCPs) from water and wastewater by carbonaceous materials: A review. Crit. Rev. Environ. Sci. Technol. 2022, 52, 727–766. [Google Scholar] [CrossRef]
  127. Kun, Y.; Lizhong, Z.; Xing, B. Sorption of sodium dodecylbenzene sulfonate by montmorillonite. Environ. Pollut. 2007, 145, 571–576. [Google Scholar] [CrossRef]
  128. Zhao, J.; Wang, Z.; Ghosh, S.; Xing, B. Phenanthrene binding by humic acid-protein complexes as studied by passive dosing technique. Environ. Pollut. 2014, 184, 145–153. [Google Scholar] [CrossRef]
  129. Awfa, D.; Ateia, M.; Fujii, M.; Yoshimura, C. Novel Magnetic Carbon Nanotube-TiO2 Composites for Solar Light Photocatalytic Degradation of Pharmaceuticals in the Presence of Natural Organic Matter. J. Water Process. Eng. 2019, 31, 100836. [Google Scholar] [CrossRef]
  130. Liu, F.; Zhao, J.; Wang, S.; Du, P.; Xing, B. Effects of Solution Chemistry on Adsorption of Selected Pharmaceuticals and Personal Care Products (PPCPs) by Graphenes and Carbon Nanotubes. Environ. Sci. Technol. 2014, 48, 13197–13206. [Google Scholar] [CrossRef]
  131. Liao, X.; Chen, C.; Liang, Z.; Zhao, Z.; Cui, F. Selective adsorption of antibiotics on manganese oxide-loaded biochar and mechanism based on quantitative structure–property relationship model. Bioresour. Technol. 2023, 367, 128262. [Google Scholar] [CrossRef] [PubMed]
  132. Ye, Y.; Bruning, H.; Liu, W.; Rijnaarts, H.; Yntema, D. Effect of dissolved natural organic matter on the photocatalytic micropollutant removal performance of TiO2 nanotube array. J. Photochem. Photobiol. A Chem. 2019, 371, 216–222. [Google Scholar] [CrossRef]
  133. Gora, S.L.; Andrews, S.A. Adsorption of natural organic matter and disinfection byproduct precursors from surface water onto TiO2 nanoparticles: pH effects, isotherm modelling and implications for using TiO2 for drinking water treatment. Chemosphere 2017, 174, 363–370. [Google Scholar] [CrossRef]
  134. Mwaanga, P.; Carraway, E.R.; Schlautman, M.A. Preferential sorption of some natural organic matter fractions to titanium dioxide nanoparticles: Influence of pH and ionic strength. Environ. Monit. Assess. 2014, 186, 8833–8844. [Google Scholar] [CrossRef]
  135. Li, X.; Li, L.; Tang, L.; Mei, J.; Fu, J. Unveiling combined ecotoxicity: Interactions and impacts of engineered nanoparticles and PPCPs. Sci. Total Environ. 2024, 921, 170746. [Google Scholar] [CrossRef]
  136. Prarat, P.; Hongsawat, P.; Punyapalakul, P. Amino-functionalized mesoporous silica-magnetic graphene oxide nanocomposites as water-dispersible adsorbents for the removal of the oxytetracycline antibiotic from aqueous solutions: Adsorption performance, effects of coexisting ions, and natural organic matter. Environ. Sci. Pollut. Res. Int. 2020, 27, 6560–6576. [Google Scholar] [CrossRef] [PubMed]
  137. Nie, W.; Li, Y.; Chen, L.; Zhao, Z.; Zuo, X.; Wang, D.; Zhao, L.; Feng, X. Interaction between multi-walled carbon nanotubes and propranolol. Sci. Rep. 2020, 10, 10259. [Google Scholar] [CrossRef]
  138. Lin, D.; Tian, X.; Li, T.; Zhang, Z.; He, X.; Xing, B. Surface-bound humic acid increased Pb2+ sorption on carbon nanotubes. Environ. Pollut. 2012, 167, 138–147. [Google Scholar] [CrossRef]
  139. Aiken, G.R.; Hsu-Kim, H.; Ryan, J.N. Influence of Dissolved Organic Matter on the Environmental Fate of Metals, Nanoparticles, and Colloids. Environ. Sci. Technol. 2011, 45, 3196–3201. [Google Scholar] [CrossRef]
  140. Huixian, Z.; Ping, J.; Qiurong, Z. Influence of Dissolved Organic Matter on Environmental Behavior and Biological Effects of Titanium Dioxide Nanoparticles. Asian J. Ecotoxicol. 2024, 19, 195–205. [Google Scholar]
  141. Lu, L.; Rijian, L.; Qiucen, S.; Jing, Z.; Guipeng, Y. Research progress on photodegradation behavior of emerging pollutants in water containing dissolved organic matter. Environ. Chem. 2024, 43, 1429–1444. [Google Scholar]
  142. Li, Y.; Yang, C.; Guo, X.; Dang, Z.; Li, X.; Zhang, Q. Effects of humic acids on the aggregation and sorption of nano-TiO2. Chemosphere 2015, 119, 171–176. [Google Scholar] [CrossRef] [PubMed]
  143. Dong, S.; Gong, Y.; Zeng, Z.; Chen, S.; Ye, J.; Wang, Z.; Dionysiou, D.D. Dissolved organic matter promotes photocatalytic degradation of refractory organic pollutants in water by forming hydrogen bonding with photocatalyst. Water Res. 2023, 242, 120297. [Google Scholar] [CrossRef]
  144. Yang, L.; Liu, M.; Liu, Y.; Luo, S.; Luo, Y.; Luo, X.; Li, G.; Peng, P. Theoretical analyses of organic acids assisted surface-catalyzed reduction of Cr VI on TiO2 nanowire arrays. Appl. Catal. B Environ. 2016, 198, 508–515. [Google Scholar] [CrossRef]
  145. Loosli, F.; Le Coustumer, P.; Stoll, S. TiO2 nanoparticles aggregation and disaggregation in presence of alginate and Suwannee River humic acids. pH and concentration effects on nanoparticle stability. Water Res. 2013, 47, 6052–6063. [Google Scholar] [CrossRef]
  146. Sun, D.D.; Lee, P.F. TiO2 microsphere for the removal of humic acid from water: Complex surface adsorption mechanisms. Sep. Purif. Technol. 2012, 91, 30–37. [Google Scholar] [CrossRef]
  147. Luo, M.; Huang, Y.; Zhu, M.; Tang, Y.; Ren, T.; Ren, J.; Wang, H.; Li, F. Properties of different natural organic matter influence the adsorption and aggregation behavior of TiO2 nanoparticles. J. Saudi Chem. Soc. 2018, 22, 146–154. [Google Scholar] [CrossRef]
  148. Zhang, R.; Zhang, H.; Tu, C.; Luo, Y. The limited facilitating effect of dissolved organic matter extracted from organic wastes on the transport of titanium dioxide nanoparticles in acidic saturated porous media. Chemosphere 2019, 237, 124529. [Google Scholar] [CrossRef]
  149. Carlos, L.; Cipollone, M.; Soria, D.B.; Sergio Moreno, M.; Ogilby, P.R.; García Einschlag, F.S.; Mártire, D.O. The effect of humic acid binding to magnetite nanoparticles on the photogeneration of reactive oxygen species. Sep. Purif. Technol. 2012, 91, 23–29. [Google Scholar] [CrossRef]
  150. Wang, H.; Zhang, L.; Chen, Z.; Hu, J.; Li, S.; Wang, Z.; Liu, J.; Wang, X. Semiconductor heterojunction photocatalysts: Design, construction, and photocatalytic performances. Chem. Soc. Rev. 2014, 43, 5234. [Google Scholar] [CrossRef]
  151. Zhang, Y.; Li, K.; Zang, M.; Cheng, Y.; Qi, H. Graphene-based photocatalysts for degradation of organic pollution. Chemosphere 2023, 341, 140038. [Google Scholar] [CrossRef] [PubMed]
  152. Brame, J.; Long, M.; Li, Q.; Alvarez, P. Inhibitory effect of natural organic matter or other background constituents on photocatalytic advanced oxidation processes: Mechanistic model development and validation. Water Res. 2015, 84, 362–371. [Google Scholar] [CrossRef] [PubMed]
  153. Peng, H.; Chen, Y.; Mao, L.; Zhang, X. Significant changes in the photo-reactivity of TiO2 in the presence of a capped natural dissolved organic matter layer. Water Res. 2017, 110, 233–240. [Google Scholar] [CrossRef]
  154. Li, Z.; Wang, F.; Zhang, M.; Lian, F.; Wang, M.; Sun, B.; Li, Y.; Li, M.; Lu, Y.; Sun, H. Citric acid improved photocatalytic performance of iron-enriched sludge hydrochar towards organic Pollutants: Roles of iron species and dissolved organic matter. Sep. Purif. Technol. 2025, 354, 128661. [Google Scholar] [CrossRef]
  155. Wu, H.; Sun, Q.; Chen, J.; Wang, G.; Wang, D.; Zeng, X.; Wang, J. Citric acid-assisted ultrasmall CeO2 nanoparticles for efficient photocatalytic degradation of glyphosate. Chem. Eng. J. 2021, 425, 130640. [Google Scholar] [CrossRef]
  156. Liu, S.; Edara, P.C.; Schäfer, A.I. Influence of organic matter on the photocatalytic degradation of steroid hormones by TiO2-coated polyethersulfone microfiltration membrane. Water Res. 2023, 245, 120438. [Google Scholar] [CrossRef] [PubMed]
  157. Sun, L.; Chen, D.; Wan, S.; Yu, Z.; Li, M. Role of small molecular weight organic acids with different chemical structures as electron donors in the photocatalytic degradation of ronidazole: Synergistic performance and mechanism. Chem. Eng. J. 2017, 326, 1030–1039. [Google Scholar] [CrossRef]
  158. Feng, Z.; Yu, J.; Yang, Z.; Liu, D.; Xu, J.; Ning, Y.; Jiang, F.; Yang, S.; Li, Y. Fe-Mn bimetallic sulfide enhanced persulfate activation for monochlorobenzene degradation in groundwater: Performance, mechanism and applicability. Chem. Eng. J. 2025, 513, 162790. [Google Scholar] [CrossRef]
  159. Cheng, X.; Liang, L.; Ye, J.; Li, N.; Yan, B.; Chen, G. Influence and mechanism of water matrices on H2O2-based Fenton-like oxidation processes: A review. Sci. Total Environ. 2023, 888, 164086. [Google Scholar] [CrossRef]
  160. Cao, R.; Liu, X.; Duan, J.; Gao, B.; He, X.; Nanthi, B.; Li, Y. Opposite impact of DOM on ROS generation and photoaging of aromatic and aliphatic nano- and micro-plastic particles. Environ. Pollut. 2022, 315, 120304. [Google Scholar] [CrossRef]
  161. Coppola, A.I.; Seidel, M.; Ward, N.D.; Viviroli, D.; Nascimento, G.S.; Haghipour, N.; Revels, B.N.; Abiven, S.; Jones, M.W.; Richey, J.E.; et al. Marked isotopic variability within and between the Amazon River and marine dissolved black carbon pools. Nat. Commun. 2019, 10, 4018. [Google Scholar] [CrossRef] [PubMed]
  162. Du, Z.; He, Y.; Fan, J.; Fu, H.; Zheng, S.; Xu, Z.; Qu, X.; Kong, A.; Zhu, D. Predicting apparent singlet oxygen quantum yields of dissolved black carbon and humic substances using spectroscopic indices. Chemosphere 2018, 194, 405–413. [Google Scholar] [CrossRef] [PubMed]
  163. Du, R.; Zhang, Q.; Wang, B.; Huang, J.; Deng, S.; Yu, G. Quantitative structure-activity relationship models for the reaction rate coefficients between dissolved organic matter and PPCPs. J. Hazard. Mater. 2023, 458, 131845. [Google Scholar] [CrossRef] [PubMed]
  164. Liu, S.; Cui, Z.; Ding, D.; Bai, Y.; Chen, J.; Cui, H.; Su, R.; Qu, K. Effect of the molecular weight of DOM on the indirect photodegradation of fluoroquinolone antibiotics. J. Environ. Manag. 2023, 348, 119192. [Google Scholar] [CrossRef] [PubMed]
  165. Xu, H.; Cooper, W.J.; Jung, J.; Song, W. Photosensitized degradation of amoxicillin in natural organic matter isolate solutions. Water Res. 2011, 45, 632–638. [Google Scholar] [CrossRef]
  166. Cheng, F.; Zhang, T.; Yang, H.; Liu, Y.; Qu, J.; Zhang, Y.; Peijnenburg, W.J.G.M. Effects of dissolved organic matter and halogen ions on phototransformation of pharmaceuticals and personal care products in aquatic environments. J. Hazard. Mater. 2024, 469, 134033. [Google Scholar] [CrossRef]
  167. Shang, E.; Li, Y.; Niu, J.; Zhou, Y.; Wang, T.; Crittenden, J.C. Relative importance of humic and fulvic acid on ROS generation, dissolution, and toxicity of sulfide nanoparticles. Water Res. 2017, 124, 595–604. [Google Scholar] [CrossRef]
  168. Goldstone, J.V.; Pullin, M.J.; Bertilsson, S.; Voelker, B.M. Reactions of hydroxyl radical with humic substances: Bleaching, mineralization, and production of bioavailable carbon substrates. Environ. Sci. Technol. 2002, 36, 364–372. [Google Scholar] [CrossRef]
Figure 1. SEM of different types of photocatalytic materials. (a) TiO2 NPs [22]; (b) g-C3N4 [23]; (c) CdS [24]; (d) BiVO4 [25]; (e) ZnO/g-C3N4 [26]; (f) Bi/Bi2WO6 [27]; (g) BiOBr/black-TiO2/tourmaline composites [28]; (h) TiO2/MWCNT/PANI [29]; (i) SnO2-x/CD-MOF S-Scheme [30].
Figure 1. SEM of different types of photocatalytic materials. (a) TiO2 NPs [22]; (b) g-C3N4 [23]; (c) CdS [24]; (d) BiVO4 [25]; (e) ZnO/g-C3N4 [26]; (f) Bi/Bi2WO6 [27]; (g) BiOBr/black-TiO2/tourmaline composites [28]; (h) TiO2/MWCNT/PANI [29]; (i) SnO2-x/CD-MOF S-Scheme [30].
Molecules 30 02266 g001
Figure 2. Interaction mechanism between dissolved organic matter (DOM) and pharmaceuticals and personal care products (PPCPs).
Figure 2. Interaction mechanism between dissolved organic matter (DOM) and pharmaceuticals and personal care products (PPCPs).
Molecules 30 02266 g002
Figure 3. Effect of DOM on the apparent structure of photocatalysts.
Figure 3. Effect of DOM on the apparent structure of photocatalysts.
Molecules 30 02266 g003
Figure 4. Effect of DOM on active species.
Figure 4. Effect of DOM on active species.
Molecules 30 02266 g004
Figure 5. Schematic diagram of the photocatalytic degradation of PPCPs.
Figure 5. Schematic diagram of the photocatalytic degradation of PPCPs.
Molecules 30 02266 g005
Table 1. Effect of DOM on the photocatalytic oxidation of PPCPs.
Table 1. Effect of DOM on the photocatalytic oxidation of PPCPs.
PhotocatalystsTypes of Light SourcesCatalyst Dosage (g·L−1)PPCPsDOM and Its ConcentrationLevel of InfluenceReference
TiO2UV1.0BUPHA, 10 mg·L−1−, 80%[31]
TiO2 (anatase)UV0.24-CPHA, 1–30 mg·L−1[32]
TiO2 (rutile)HA, <20 mg·L−1+
HA, 30 mg·L−1
TiO2 TNAsUV/MTLNOM, 15 mg·L−1−, 48.16%[33]
ZnOvis0.8TCHA, 5 mg·L−1−, 19%[34]
ZnO NPsvis0.01MBHA, 10 mg·L−1−, 79.4%[35]
ZnO nanowiresvis0.02CPXSRNOM, 10 mg·L−1−, 59.1%[36]
Mn-WO3LED2.2DCFHA[37]
Oxygen-doped porous g-C3N4vis1.0CBZHA, 20 mM+[38]
g-C3N4 nanosheetsvis1.0PHEHA, 10 mg·L−1−, 18%[39]
GOvis0.1APAPFA, 20 mg·L−1−, 84.8%[40]
CdS@BCvis1.0THMHA[41]
SrBiOIvis0.4IDMDOM, 10 mg·L−1−, 30%[42]
MI-BiOClvis0.4VENHA, 20 mg·L−1o[43]
Bi-TNBvis0.5NPXHA, 5 mg·L−1+, 2 times[44]
FA, 10 mg·L−1
SnO2@ZnSvis/MTLDOM−, 52%[45]
BiOBr/Fe3O4vis0.5NORHA, 10 mM−, 44.49%[46]
BiOBr/Ti3C2vis2.4FQNsHAo[47]
Bi2MoO6/(BiO)2CO3vis0.5APAPNOM, 10 mg·L−1−, 21%[48]
Er3+-CdS/MoS2vis0.12517β-E2HA+[49]
WO3 Fibers/g-C3N4vis0.1CTDHA, 20 mg·L−1−, 12.9%[50]
g-C3N4/Ag3PO4vis0.5OFXHA[51]
rGO/TiO2Blue light0.3DCFFA, ≤10 mg·L−1+[52]
FA > 10 mg·L−1
Bi2Fe4O9/rGOvis0.2TCHA[53]
MWCNTs/Bi4O5I2 nanosheetsvis0.2TCHA, 8 mg·L−1−, 5%[54]
CoFe2O4-rGOUVA-
LED
0.4BPAHA, 10 mg·L−1[55]
Ag2WO4/PCNvis0.2IDMDOM, 10 mg·L−1−, 26%[56]
PO43−-Bi2WO6/PIvis1.0TCHA[57]
Benzene-ring doped CN/Phosphorus-doped CNBlue light (LED)0.2SSZDOM, 10 mg·L−1−, 74.8%[58]
WO3/ZnIn2S4-3LED1.6TBBPAHA, 1 mmol·L−1−, 53.9%[59]
CCPD-g-C3N4vis1.0MBPSRHA, 50 mg·L−1−, 20%[60]
CQD-SnNb2O6/BiOClvis0.5BCHA, 10 mg·L−1−, 18%[61]
Cd0.5Zn0.5S/BiOClvis0.2NORHA, 20 mg·L−1−, 35%[62]
IS-Ni2P/CdS/CNvis0.1TCHA+[63]
VW+Br-BiOBr/Bi2WO6vis0.3NORHA, 10 mg·L−1−, 21.9%[64]
BMOF-Ti/Zr6%UV/BP-2DOM, 20 μg·L−1[65]
Zr-MOFsvis/STZHA[66]
π-COFvis0.2TCHA[67]
ZnFe2O4-seed@TpTt-COFvis0.1BPAHA+[68]
MIL-88B(Fe)/ZnTi-LDH high-low junctionvis0.2TCHA+[69]
Promotion (+); inhibitory (−); no noteworthy influence (o).
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

Wang, J.; Zhu, M.; Sun, A.; Yuan, R.; Chen, H.; Zhou, B. The Influence Mechanism of Dissolved Organic Matter on the Photocatalytic Oxidation of Pharmaceuticals and Personal Care Products. Molecules 2025, 30, 2266. https://doi.org/10.3390/molecules30112266

AMA Style

Wang J, Zhu M, Sun A, Yuan R, Chen H, Zhou B. The Influence Mechanism of Dissolved Organic Matter on the Photocatalytic Oxidation of Pharmaceuticals and Personal Care Products. Molecules. 2025; 30(11):2266. https://doi.org/10.3390/molecules30112266

Chicago/Turabian Style

Wang, Jie, Minyi Zhu, Anli Sun, Rongfang Yuan, Huilun Chen, and Beihai Zhou. 2025. "The Influence Mechanism of Dissolved Organic Matter on the Photocatalytic Oxidation of Pharmaceuticals and Personal Care Products" Molecules 30, no. 11: 2266. https://doi.org/10.3390/molecules30112266

APA Style

Wang, J., Zhu, M., Sun, A., Yuan, R., Chen, H., & Zhou, B. (2025). The Influence Mechanism of Dissolved Organic Matter on the Photocatalytic Oxidation of Pharmaceuticals and Personal Care Products. Molecules, 30(11), 2266. https://doi.org/10.3390/molecules30112266

Article Metrics

Back to TopTop