DOI:
10.1039/D5RA08016A
(Review Article)
RSC Adv., 2025,
15, 49852-49874
ZnO based photocatalysts for pesticides degradation
Received
19th October 2025
, Accepted 6th December 2025
First published on 15th December 2025
Abstract
The widespread application of pesticides in modern agriculture has significantly boosted crop production; however, their inherent toxicity, persistence, and resistance to conventional cleanup methods have led to serious environmental and public health concerns. Advanced oxidation processes (AOP), especially those utilizing visible light for photocatalysis, have recently emerged as promising eco-friendly alternatives for the degradation of pesticides. In particular, zinc oxide (ZnO) based nanophotocatalysts have garnered considerable attention due to their wide band gap (∼3.37 eV), strong oxidative capability, high electron mobility, low electron–hole recombination rates, and natural antibacterial properties, which enhance their photocatalytic activity under sunlight. This review provides a comprehensive overview of recent progress in ZnO-mediated photocatalytic degradation of pesticides, focusing on synthesis methods, structural modifications such as doping and defect engineering, and material hybridization aimed at improving photocatalytic efficiency. Furthermore, the study critically examines the influence of key factors, including catalyst concentration, surface morphology, and particle size, on degradation performance. This review aims to offer a thorough understanding of the versatility of ZnO as a tunable photocatalyst for mitigating pesticide contamination in wastewater by combining mechanistic insights with experimental observations. This integration not only highlights the potential of ZnO in this context but also establishes a foundation for creating scalable and eco-friendly remediation approaches.
1. Introduction
Pesticides play a vital role in modern agriculture, serving as indispensable tools for farmers to manage and mitigate weeds and control pests in farming practices. They play a significant role in combating insect-borne diseases and are also believed to greatly enhance agricultural productivity by reducing crop losses, improving yield, and ensuring the quality of food, all while being a cost-effective solution for farmers.1 The application of pesticides on crops is estimated to be 3.5 million tons every year at a global scale. Despite the fact that approximately 1% of pesticides are effectively used to target and manage pests on the intended crops, the vast majority of these chemicals end up affecting non-target plants or dispersing into the environment. This unintended distribution leads to significant contamination of soil, water, and air, posing serious risks to ecosystems and public health.2 Consequently, the widespread use of pesticides disrupts food webs in both terrestrial and aquatic ecosystems, as the chemicals and their byproducts are often resistant to biodegradation, leading to persistent environmental contamination.3 Additionally, these agrochemicals contribute to a reduction in biodiversity, a decrease in pollinator populations, harm to native soil microorganisms, and the disturbance of nesting habitats.4
Moreover, pesticide exposure is linked to a variety of health problems, affecting individuals in environmental, community, and occupational settings. These health issues can manifest as both acute and chronic effects, including cancer, genetic mutations, neurotoxicity,5 and developmental disorders.6 It is estimated that annually, more than a million agricultural workers show signs of pesticide poisoning.7 As a result, the breakdown of pesticides and their residues is crucial. Many research efforts focus on techniques to minimize pesticide levels in the environment, particularly in soil and water, employing methods such as membrane filtration, surface adsorption, and biological degradation. However, these approaches may have limitations when faced with high levels of contamination.8
Nanoparticles have garnered significant attention for their potential in pesticide degradation, with ZnO nanoparticles being particularly notable due to their abundance, stability, high reactivity, large surface area, excellent photosensitivity, and cost-effectiveness.9 Furthermore, zinc oxide (ZnO) photocatalysts can be easily immobilized on various substrates, enhancing their applicability in diverse water treatment processes.10 These advantageous characteristics drive researchers to innovate and create advanced ZnO hybrid photocatalysts with improved photo-efficiency, aimed at effectively breaking down hazardous pollutants such as pesticides.
Therefore, in view of the aforesaid, this review employed a narrative literature approach, primarily sourcing scientific publications, review articles, and reports from online databases such as Google Scholar, ScienceDirect, Springer, PubMed, and Scopus. Key terms included “ZnO photocatalysis”, “pesticide degradation”, “nanoparticle remediation”, and “environmental impact of pesticides”. While studies published in peer-reviewed journals over the past 10 to 15 years were prioritized, the review did not impose strict inclusion or exclusion criteria. It aims to provide a comprehensive understanding of ZnO-mediated photocatalytic pesticide degradation by integrating both theoretical and experimental findings, alongside relevant earlier research that contributes valuable context or mechanistic insights. Only English-language publications specifically addressing ZnO-based photocatalytic pesticide breakdown were considered, while patents, conference abstracts, and non-peer-reviewed materials were excluded. Moreover, this manuscript was proofread and edited for clarity and language enhancement using ChatGPT; OpenAI. The authors assume full responsibility for all content and ensure the accuracy and integrity of the work.
2. Environmental fate and impact of pesticides
Pesticides are man-made chemical substances that include nematicides, herbicides, fungicides, acaricides, insecticides, and molluscicides. These compounds are employed to control, eliminate, reduce, or repel organisms that are detrimental to crops or cause damage. Their absence would result in a startling 78% loss in fruit productivity, a 54% decrease in vegetable production, and a 32% decrease in cereal production.11 Interestingly, pesticide use skyrocketed during World War II as a result of the increased need for food by a growing population. Initially, synthetic pesticides were primarily designed to eliminate mosquitoes, particularly those that carry malaria.12 But after the 1950s, global pesticide production has seen growth at an average annual rate of approximately 11%, escalating from 0.2 million tons (ref. 13) to over 5 million tons by the year 2000.14 Whereas annually, around 3 billion kilograms of pesticides are utilised globally.15
Likewise, synthetic pesticides have been used by humans since 1940; ever since, they vary in chemical, physical, and other characteristics from one category to another. Hence, it is important to classify them according to their properties and study them within their specific category. At present, pesticides can be broadly categorized in two ways: (a) based on the type of pest they target (Fig. 1) and (b) according to their chemical composition (Table 1).
 |
| | Fig. 1 Classification of pesticides based on the targeted pest. | |
Table 1 Classification of pesticides based on chemical groups and relevance to ZnO-mediated photocatalysisa
| Chemical group |
Examples |
Structure type |
Relevance to ZnO photocatalysis |
Ref. |
| DDT = dichlorodiphenyltrichloroethane. 2,4-D = 2,4-dichlorophenoxyacetic acid. MCPA = 2-methyl-4-chlorophenoxyacetic acid. |
| Organochlorines |
DDT, aldrin, endosulfan, chlordane, heptachlor |
Chlorinated aromatic hydrocarbons, high stability |
Persistent organic pollutants require prolonged photocatalytic exposure to break C–Cl bonds |
16 |
| Organophosphates |
Malathion, parathion, chlorpyrifos, diazinon |
Phosphorothioates or phosphates, aliphatic or aromatic |
Degrade relatively quickly; P–O and P–S bonds cleaved under UV/visible light |
17 |
| Carbamates |
Carbaryl, aldicarb, carbofuran, methomyl |
Carbamate esters of aromatic/aliphatic amines |
ZnO photocatalysis breaks carbamate linkages, reducing toxicity |
18 |
| Pyrethroids |
Permethrin, cypermethrin, deltamethrin |
Cyclopropane carboxylate esters |
Degrade into less toxic acids and alcohols under photocatalytic conditions |
19 |
| Neonicotinoids |
Imidacloprid, thiamethoxam, acetamiprid |
Chloronicotinyl or nitroguanidine derivatives |
Water-soluble; susceptible to oxidative degradation in aqueous ZnO systems |
20 |
| Phenoxy herbicides |
2,4-D, MCPA, mecoprop |
Chlorinated phenoxyacetic acids |
Aromatic rings and –COOH groups are readily attacked by OH˙ |
21 |
| Triazines |
Atrazine, simazine, propazine |
Nitrogen-containing heterocyclic aromatic rings |
Photocatalysis disrupts the triazine ring structure and dechlorinates |
22 |
| Bipyridyl herbicides |
Paraquat, diquat |
Bipyridinium salts |
Strongly absorb visible light; photocatalysis reduces toxicity via demethylation and ring cleavage |
23 |
| Fumigants |
Methyl bromide, aluminium phosphide |
Simple halides or phosphides |
Easily decomposed into non-toxic products under photocatalytic oxidation |
24 |
| Rodenticides |
Warfarin, brodifacoum |
Coumarin derivatives |
Aromatic and lactone structures oxidised under UV/visible photocatalysis |
25 |
However, the overuse of pesticides poses a significant risk to non-target organisms, as nearly 98% of all sprayed pesticides, whether directly or indirectly, impact them. Evidence suggests that soil and water quality deteriorate due to pesticide accumulation. Moreover, pesticide accumulation can lead to a decrease in soil respiration up to 35%. Existing literature indicates that nearly 90% of water sources in agricultural areas are polluted with pesticides.26
What is more, in the presence of various existing persistent pollutants (industrial compounds, natural solvents, and cleansing agents), water-soluble pesticides are causing great trouble. Poor management, particularly in on-farm handling, is the primary contributor to pesticide contamination in the ecosystem.27 There are many different sources of pesticide contamination, such as industrial manufacturing processes, the discharge from agricultural activities like container cleaning, spraying, and washing contaminated crops, soil degradation, atmospheric deposition, and many more, which potentially lead to environmental contamination through processes like bioaccumulation.28 So basically, the persistent pesticides reach the environment by either direct or indirect means. What's more, pesticide pollution originates from two main categories of sources: point sources (specific) and non-point sources, also known as diffuse sources. This includes scenarios like the pesticide's transfer from various surface water sources, leading to water pollution, which affects both aquatic and land ecosystems. Along with other sources, agriculture is a major contributor to diffuse pollution, which is caused by activities without specific release points.29 On a global and regional level, non-point source pollution is a major environmental challenge and is acknowledged as a major contributor to the deterioration of water quality. According to reports, non-point source pollution has affected between 30% and 50% of surface water bodies globally.30
Overview of pesticides and their environmental impact revealed that the excessive use of pesticides ultimately leads to their accumulation in soils, then either eliminates microorganisms that are essential for many biological processes, like nutrient uptake or organic matter breakdown, or inhibits their activity, which leads to lower soil fertility.31 These inadequately and excessively used products can also harm non-targeted organisms because only a small portion (1–3%) of the pesticides being used reach their targets. The dispersal of contaminants to non-target areas occurs irrespective of any natural barriers, and because of multiple transport pathways, it can be difficult to track the contamination.32 Preliminary findings show that aquatic organisms are seriously endangered due to the presence of pesticides in waterbodies. Fish communities experience behavioural and physiological changes, and aquatic plants suffer from lower oxygen levels in the water.33 Additionally, pesticides affect terrestrial fauna populations, including beneficial insects, whose numbers may significantly decrease when broad-spectrum insecticides are applied.34 Alternatively, the accumulation of pesticides in birds' and mammals' tissues has been responsible for a decline in their population because of the adverse effects on the nervous systems. It causes behavioural changes that can lead to death.35 Through contaminated feed, water, or direct touch, farm animals are also exposed to pesticides, which can have detrimental effects. Multiple studies report that acute toxicity, immune system weakness, reproductive issues, and organ damage are all possible outcomes of pesticide exposure in farm animals (Table 2). Even at low concentrations, prolonged exposure might cause long-term health problems such as hormone imbalances and decreased productivity. Pesticide residues that bioaccumulate in animal tissues can also endanger human health by way of the food chain.
Table 2 Pesticide impacts on farm animals
| Pesticide class |
Adverse effects in farm animals |
Ref. |
| Carbamates |
Neuromuscular weakness; potential residues in milk/eggs if misuse |
43 |
| Pyrethroids |
Dermal/respiratory irritation; stress; resistance issues in stable flies affecting cattle |
44 |
| Neonicotinoids |
Potential residue transfer via feed; limited direct farm-animal toxicity at labelled uses, but monitor feed contamination |
45 |
| Phenoxy herbicides |
Indirect effects via forage contamination, and if misused, may lead to milk residue risks (e.g., feed-to-milk pathway) |
46 |
| Organochlorine |
Bioaccumulation in animals and milk, causing potential human exposure |
47 |
In a similar manner, humans can be adversely affected by pesticides through direct exposure to them during agricultural activities (Fig. 2). Because pesticides came into contact with humans via household vegetation, or having occupational-related farming, as well as indirectly through the food chain and environmental contamination.36 Exposure of humans to pesticides conveys a significant health risk, with both acute effects and chronic effects on health. Numerous health consequences result from these effects, including acute and long-term neurotoxicity from fungicides, insecticides, and fumigants; lung damage from the pesticide paraquat; infant methemoglobinemia from nitrate seeping into groundwater; and chemical injuries like burns from pesticide exposure, such as anhydrous ammonia.37 Furthermore, a variety of cancers have been linked to pesticide exposure, including hematopoietic cancers, cancers of the digestive tract, cancers of the reproductive system, bladder cancer, breast cancer, and lung cancer.38 Cholinergic effects include the possibility of immunologic abnormalities39 and adverse effects on reproductive and developmental processes.40 Moreover, pesticides can have an impact on health even if a person is not showing any serious signs of illness. People who get exposure often report difficulty in localizing sensations while their muscle strength is reduced.41 The causes of diseases that may arise from exposure to pesticides can vary depending on a number of factors. These variables include the kind of pesticide applied, the exposure method and duration, and the general health of the person. Pesticides undergo various processes after entering the bodies of animals or humans, which include metabolism, excretion, storage, and accumulation in adipose tissues42 as shown in Table 3.
 |
| | Fig. 2 Pathways of pesticides entry into the human body.15 Reproduced from ref. 15 with permission from MDPI, copyright 2022. | |
Table 3 Main classes of pesticide and their effect on human health
| Pesticide class |
Effect |
Ref. |
| Organophosphate |
Chronic central nervous system disorders, respiratory disorders, cardiovascular diseases, diabetes mellitus, cancer, infertility issues, impaired vision, Alzheimer's disease, cellular oxidative stress, Parkinson's disease, hypotension, etc. |
48–50 |
| Organochlorine |
Cardiovascular disorders, hypertension, neurological disorders, hormone-related cancers (breast, prostate, lung, stomach), obesity, endocrine disruption, alterations in embryonic development, hematologic and hepatic changes, diabetes in overweight individuals, learning disabilities in children, hyperactivity disorder, and Parkinson's disease |
51 and 52 |
| Carbamate |
Immunotoxicity, cholinergic poisoning, male infertility, rndocrine disruption and inhibition of esterases, endoplasmic reticulum stress, etc. |
53 and 54 |
| Pyrethroids |
Respiratory distress, nausea, tachycardia, apathy, metabolic acidosis, convulsions, anaphylactic shock, pulmonary edema, and oxidative stress. Risks to reproductive health and neurobehavioral, cancer, and the development of autism spectrum disorders in infants, etc. |
55–57 |
3. Conventional method of pesticide remediation
Currently, various remediation methods are utilised for addressing the water58 and soil59 contamination with pesticides. If one technology proves ineffective, combining multiple methods becomes necessary to achieve satisfactory outcomes. Therefore, numerous techniques have been suggested and implemented over time to eliminate persistent pollutants. Traditional approaches for eliminating these pollutants include membrane filtration,60 surface trapping,61 ozonolysis, air stripping, skimming, photolysis, and Fenton oxidation processes.62 Biodegradation, particularly through microbial action, is also found to be a versatile and effective strategy for remediating pesticide-contaminated sites due to the capability of microbes to function even in harsh environmental conditions.63 However, this method is both time-consuming and inefficient, with limitations in its applicability. Likewise, processes like sedimentation, membrane technologies, and chemical filtration entail significant operational expenses and result in the generation of highly poisonous secondary pollutants that enter nature.64 In addition, various physicochemical techniques such as reverse osmosis, carbon adsorption, adsorption, nano-filtration, distillation, adsorption, and ion exchange resins have also been utilised recently. Adsorption and coagulation are the processes that primarily concentrate pollutants, changing their phase, rather than eliminating or degrading them completely.65 Therefore, these methods encounter significant challenges, including issues with disposal, membrane distortion, formation of sludge, operational handling, and various technical limitations.66 In summary, the various physical, chemical, and biological methods employed for pesticide removal each possess distinct advantages and disadvantages. These characteristics influence their effectiveness based on factors such as the type of pollutant, its concentration, and the specific goals of the treatment, as illustrated in Fig. 3.
 |
| | Fig. 3 Comparison of treatment methods for pesticide removal: distinct advantages and limitations associated with each approach. | |
Since the majority of persistent pollutants are resistant to breakdown through conventional chemical and biological treatment approaches. The methods described above are efficient, but they have limitations in terms of their effectiveness, applicability, cost, and time. Consequently, the inadequacy of current treatment processes for pesticide remediation has heightened the need for advanced technologies for treatment. So, scientists have explored alternative approaches to degrade persistent pesticides completely into eco-friendly substances, such as the advanced oxidation process (AOPs).
4. Advanced oxidation process
To address the downsides of traditional treatment methods, scientists are exploring AOPs as a more effective way to remove pesticides from wastewater. AOP stands out as an exceptionally effective and advantageous method for purifying polluted water. It efficiently transforms pesticides or any other organic contaminants into water, carbon dioxide, and basic salts.67 It employs a potent oxidant like hydroxyl radical (OH˙), possessing the second-highest oxidising power, approximately 2.8 eV less than fluorine. These radicals can interact with nearly all organic pollutants at rate constants ranging from 106 to 109 mol L−1 s−1.68 This advanced technology is becoming more promising and an increasingly adopted method for addressing recalcitrant wastewater, containing a lot of organic matter and low pH levels. Moreover, this approach is employed to neutralise pathogens after secondary treatment.69
Generally, AOP involves activating semiconductor photocatalysts through light that leads to the formation of OH˙ that oxidizes pollutants. The reduction of adsorbed oxygen molecules generates oxygen radicals that help break down pollutants.70 The indiscriminate characteristic of OH˙ renders them suitable for remediation in the environment.71 Additional oxidative agents employed in AOP include ozone and superoxide radicals (O2−).72 The general classification of AOP can be seen in Fig. 4.
 |
| | Fig. 4 Classification of AOPs. This figure has been reproduced with permission from ref. 73, Elsevier, copyright 2024. | |
5. Photocatalysis
Within this sequence of procedures, photocatalysis, which relies on the use of a photocatalyst to absorb accelerated photons, is one of the most effective methods in the AOP. According to the Photocatalysis Industry Association of Japan, a photocatalyst is a substance that accelerates chemical reactions by utilizing an external energy source. One important application of photocatalysts is the degradation of harmful substances such as pesticides. By generating reactive free radicals, particularly OH˙, photocatalysts can convert highly toxic pesticides into less harmful compounds.74 This process begins when the photocatalyst absorbs radiation with wavelengths longer than 290 nm. The absorbed energy excites electrons in the catalyst material, promoting them to a higher energy state. These excited electrons, along with the resulting positive “holes” left behind, can interact with surrounding molecules. The interaction leads to the formation of reactive species such as superoxide and OH˙. These radicals then react with the active components of pesticides or other pollutants through oxidation–reduction reactions (Fig. 5), ultimately leading to their breakdown and the detoxification of water.75
 |
| | Fig. 5 Photocatalytic processes over a heterogeneous photocatalyst.79 Reproduced from ref. 79 with permission from MDPI, copyright 2021. | |
Photocatalysts offer safety, cost-effectiveness, exceptional durability, chemical and biological inertness, insolubility in most scenarios, and the ability to be recycled and reused. They can be activated by both artificial light and sunlight.76 However, there are several challenges in photocatalysis research that must be addressed before practical application. These include the inability to function effectively in solar radiation, issues with reactor design, limitations on catalyst reuse and recovery, lower quantum efficiency, and the potential production of toxic byproducts.77 Another challenge in photocatalysis is the rapid recombination of photo-generated electron–hole pairs, which can hinder the efficiency of the process. This phenomenon, known as charge recombination, results in the release of electron energy as heat, ultimately terminating the reaction. To address this issue, doping with metal oxide nanomaterials is recommended. Doping helps prevent electron–hole recombination by introducing metal ions, thus extending the lifespan of electron–hole pairs enough to facilitate redox reactions at the catalyst surface.78
The photocatalysis process is of two kinds: homogeneous and heterogeneous. Homogeneous photocatalysis occurs when a soluble catalyst and reactant are present in a single phase. The derived homogeneous solution contains photon promoters and active catalytic sites that function in two different ways. In the first technique, a photosensitizer is used to transfer electrons to the catalyst, creating an active site that triggers a reduction reaction. In the second technique, the catalyst acts as both a catalyst and a substance that absorbs light. When light incidence excites the catalyst electrons, they shift from the highest occupied molecular orbital to the lowest unoccupied molecular orbital, causing the catalyst to act as a good substance for redox processes.80 Transition metal complexes are the most widely used homogeneous photocatalyst due to their advantageous energy band gap properties and stability, while ozone is the second most commonly used catalyst.70,81 The main disadvantage of a homogeneous photocatalyst system is that it is difficult to separate the photocatalyst from the solution since it is completely soluble.82
Heterogeneous photocatalysis refers to the enhancement of photoreactions in the presence of a catalyst.83 The reaction involves the interaction of substances with multiple states. Typically, a solid photocatalyst is used, and organic pollutants or compounds in the aqueous phase are exposed to the catalyst's surface for photo degradation.84 Commonly used materials in heterogeneous photocatalysis include ZnO, ZnS, MnO2, WO3, MoO3, TiO2, SnO2, Fe2O3, CdS, CeO2, and ZrO2 (Table 4). These materials are favoured due to their chemical and mechanical stability, as well as their inertness towards biological tissues. They have a low energy gap and possess properties like high porosity, large surface area, and both hydrophobic and hydrophilic interactions that make them well-suited for the photodegradation of persistent pesticides.85,86 Heterogeneous photocatalysis is a complex process involving multiple stages to eliminate harmful substances from wastewaters. Generally, the process starts when light of sufficient energy strikes the catalyst's surface, and it is absorbed, causing an electron to move from the valence band (VB) to the conduction band (CB), leaving a hole in the valence band. This process generates an electron–hole pair called an exciton. A critical aspect of heterogeneous photocatalysis is to prevent the recombination of these electron–hole pairs and utilize excitons for a redox reaction to fully degrade toxic compounds into harmless minerals.87
Table 4 Band gap energies of a semiconductor
| Semiconductor |
Band gap (eV) |
Ref. |
| ZnO |
3.37 |
88 |
| TiO |
3.20 |
89 |
| MnO2 |
1.38 or 1.41 |
90 |
| Fe2O3 |
1.77 to 2.25 |
91 |
| SnO2 |
3.66 |
92 |
Nanomaterials, which are known for their small size and unique properties, have emerged as a promising area of interest for effectively addressing various environmental pollutants, including organic pesticides. A diverse array of materials, including metal oxides (such as TiO2, ZnO, CuO, MgO), metal nanoparticles (like Au, Ag), bimetallic nanoparticles, bio-nanopolymers (such as alginate–Ag, ZnO–cellulose), adsorbents (including triggered charcoal, zeolites, calcite, clays, and other carbonaceous materials), as well as nanoparticles and nanocomposites, have been extensively utilized for pesticide remediation over decade.93
Typically, nano-based remediation technology utilizing nanomaterials as photocatalysts stands out as the most effective advanced oxidation method for addressing pesticides and further harmful pollutants. By using photo-excitation, the photocatalyst creates both electron-donating (reducing) and electron-accepting (oxidizing) species (Fig. 6), which have great potential as redox agents. Technology like nano-photocatalysis follows the principles of green chemistry to remove harmful pollutants from the environment and human life. It achieves this by breaking down stubborn compounds (Fig. 6) into intermediary substances and eventually into harmless byproducts.94
 |
| | Fig. 6 Photocatalytic degradation of the health hazardous organophosphorus pesticides using ZnO coupled photocatalysts. This figure has been reproduced with permission from ref. 95, Elsevier, copyright 2024. | |
6. ZnO and its nanomaterials as photocatalysts
Since visible light makes up the bulk of the solar spectrum, there is a rising need to use solar energy to combat water pollution, which has led to an increased focus on Visible Light Active (VLA) photocatalysts. An ideal photocatalyst should function under both UV and visible light, remain chemically stable, resist photo-corrosion, be environmentally safe, and remain cost-effective,96 but to do so, factors such as surface area, porosity, crystalline structure, particle size, and band gap play a critical role.66
ZnO is classified as a II–VI semiconductor, characterized by a direct band gap of approximately 3.37 eV and an exciton binding energy of around 60 meV. It adopts a wurtzite hexagonal crystal structure, where the tetrahedral coordination of Zn2+ and O2− ions leads to significant polarization, influencing its electronic band structure. The conduction band is primarily derived from the Zn 4s orbitals, while the valence band is formed from the O 2p orbitals. When exposed to UV light, ZnO facilitates the generation of electron–hole pairs, as illustrated in eqn (1):
| | |
ZnO + ℏν → eCB− + hVB+
| (1) |
In nanoscale ZnO, the phenomena of quantum confinement, the presence of surface states, and oxygen vacancies contribute to the formation of shallow donor levels. These factors enhance electron density, facilitate stronger adsorption, and promote efficient electron transfer to pesticide molecules. Additionally, when ZnO is synthesized using environmentally friendly or waste-derived methods, it aligns with green chemistry principles by reducing resource consumption and minimizing its environmental impact.
ZnO exhibits band edges of approximately −0.5 eV for the conduction band and +2.7 eV for the valence band relative to the NHE, facilitating the generation of both superoxide (˙O2−) and hydroxyl radicals (˙OH). Under UV light excitation with wavelengths shorter than 380 nm, electrons are able to reduce O2 to ˙O2−, while the holes oxidize water or OH− to yield ˙OH, as depicted in Fig. 7. Reactive oxygen species exhibit a potent oxidative capacity, enabling them to break down pesticides, dyes, pharmaceuticals, and other enduring pollutants into benign products such as CO2 and H2O.97 The photocatalytic performance of ZnO is strongly influenced by band gap energy, charge carrier dynamics, and synthesis-dependent morphological factors.98
 |
| | Fig. 7 Photocatalytic degradation mechanism at the surface of ZnO. This figure has been reprinted with permission from ref. 97, Elsevier, copyright 2020. | |
Similarly, at the surface of ZnO nanoparticles, upon visible light interaction, mechanistic changes start. The photoreaction begins when electron–hole pairs are generated that migrate across the nanoparticle surface. These pairs undergo oxidation and reduction reactions on the surface of the catalyst. Upon contact with water molecules, the holes produce hydroxyl ions (OH−) and convert into hydroxyl radicals OH˙. Electrons react with molecular oxygen to form superoxide ions (O2˙−), which can further react with hydrogen ions to produce hydrogen peroxide (H2O2). Superoxide and H2O2 radicals then combine to create more OH˙ (potent oxidizing agents that break down organic pollutants into harmless products).99 These ROS, along with photogenerated holes, attack the organic pollutants, forming oxidized intermediates that undergo fragmentation toward partial or complete mineralization (CO2, H2O, and inorganic ions). The process of heterogeneous photocatalysis involving ZnO as the catalyst operates as follows (eqn (2)–(12)).70,100
| | |
ZnO + ℏν → ZnO(e−) + ZnO(h+)
| (2) |
| | |
ZnO[h+(VB)] + H2O → ZnO + H+ + OH˙
| (3) |
| | |
ZnO[h+(VB)] + OH− → ZnO + OH˙
| (4) |
| | |
ZnO[e−(CB)] + O2 → ZnO + O2˙−
| (5) |
| |
 | (6) |
| |
 | (7) |
| | |
ZnO[e−(CB)] + H2O2 → OH˙ + OH−
| (8) |
| | |
H2O2 + O2˙− → OH˙ +OH˙ + O2
| (9) |
| | |
Organic pollutants + OH˙ → intermediates
| (11) |
| | |
Intermediates → CO2 + H2O
| (12) |
In case of pesticides remediation using ZnO, the pesticides undergo photolysis, a process in which UV-visible light breaks down their large molecules into smaller and simpler ones. During this process, photons are absorbed by pesticide molecules or surrounding species (such as water or sensitizers), leading to bond cleavage. For example, when water absorbs high-energy photons, it can undergo photolysis: H2O + ℏν → ˙OH + H˙. These highly reactive ˙OH and H˙ radicals can then attack pesticide molecules, leading to their degradation.101 The fundamental idea behind photocatalysts is their ability to generate electron–hole pairs, which is how OH˙ are created.102
Research has increasingly favored ZnO over other photocatalysts for pesticide degradation, particularly when compared to titanium dioxide (TiO2), which has been the most extensively studied photocatalyst for contaminant removal. Despite its popularity, TiO2 has notable drawbacks, such as low quantum efficiency, rapid charge recombination, and a wide band gap of approximately 3.2 eV, which limits its activation to ultraviolet light. These limitations have prompted a shift in focus toward ZnO as a more effective alternative. With comparable photocatalytic activity and improved UV absorption, ZnO also allows easier modification for visible-light activity.103 ZnO further distinguishes itself with superior biological, optical, catalytic, photochemical, and antibacterial properties compared to TiO2, and has shown enhanced photocatalytic efficiency in dye-contaminated aqueous systems.104 Its higher electron mobility (two orders of magnitude greater than TiO2),105 together with versatile synthesis routes enabling diverse morphologies,106 enhances its applicability. The unique properties of ZnO, including its band structure, wide band gap, highly positive valence band, and the presence of abundant defects, contribute to its effectiveness in degrading persistent pesticides. Additionally, its strong ROS generation and high charge mobility, coupled with its low cost and non-toxic nature, position ZnO as a superior photocatalyst compared to many alternatives, as listed in Table 5. Unlike traditional treatment methods, ZnO photocatalysts stand out because they do not produce secondary pollutants, require no external chemicals and operate without high energy inputs. Their capacity to harness natural sunlight renders the process low-carbon and particularly suitable for decentralized or rural wastewater systems, where energy resources may be limited.
Table 5 Comparative analysis of ZnO with other photocatalysts: pros and cons for pesticides degradation
| Photocatalyst |
Advantages of ZnO |
Disadvantages of ZnO |
Ref. |
| ZnO vs. TiO2 |
• Higher quantum efficiency |
• TiO2 is more stable under acidic/basic conditions |
107 |
| • More oxygen vacancies (stronger ROS) |
• ZnO can undergo photocorrosion |
| • Stronger oxidative holes (higher valence band) |
| • Better adsorption of polar pesticides |
| • Faster electron mobility |
| ZnO vs. Fe2O3 |
• Fe2O3 has very poor charge mobility |
• Fe2O3 is magnetically recoverable |
108 |
| • ZnO has faster electron transfer |
| • Fe2O3 absorbs visible light but produces very weak ROS |
| ZnO vs. spinel ferrites (ZnFe2O4, CoFe2O4) |
• Ferrites have extremely high recombination |
• Ferrites are magnetically recoverable |
109 |
| • ZnO produces stronger ˙OH |
| • Ferrites often require composites to work well |
| ZnO vs. g-C3N4 |
• g-C3N4 alone is weak under UV |
• g-C3N4 works better under visible light alone |
110 |
| • ZnO produces stronger oxidative holes |
• ZnO needs UV or doping to shift into the visible region |
| • ZnO adsorbs pesticide molecules more strongly (especially polar pesticides) |
| ZnO vs. WO3 |
• WO3 has a very low conduction band (cannot generate ˙O2− radicals) |
• WO3 is more stable in acidic conditions |
111 |
| • ZnO can produce both ˙OH and ˙O2− which leads to more complete pesticide mineralization |
| • ZnO offers a higher degradation rate for chlorinated pesticides |
| ZnO vs. Ag-based photocatalysts |
• ZnO is cheap, non-toxic |
• Ag-based catalysts can show better visible-light activity |
112 |
| • Ag is expensive and toxic, risk of Ag+ leaching |
| • Ag nanoparticles can deactivate due to oxidation |
7. Methods to enhance photocatalytic properties of ZnO and its nanomaterials
The photocatalytic performance of ZnO nanoparticles is determined by their synthesis technique. The hydrothermal and sol–gel techniques produce extremely crystalline, uniform ZnO with great efficiency but require longer processing durations. Chemical precipitation is less expensive and more environmentally friendly than solvothermal synthesis, which generates high-purity ZnO but requires careful control and a lot of energy. Green synthesis provides an environmentally friendly alternative with middling efficiency (Table 6). Other approaches, including microwave-assisted, electrochemical, sonochemical, spray pyrolysis, and combustion, offer greater versatility by altering particle size, crystallinity, and overall photocatalytic activity.
Table 6 Influence of synthesis approach on the photocatalytic activity of ZnO nanomaterials
| Method |
Photo-catalyst |
Band gap (eV) |
Efficiency |
Advantage/limitation |
Ref. |
| Hydrothermal |
ZnO NPs, nitrogen-doped ZnO |
3.19 |
99.60 |
High crystallinity and purity, but requires extended processing times |
113 and 114 |
| Sol–gel |
Nanosized ZnO |
3.30 |
99.00 |
Simple and adaptable; produces uniform materials but may take longer to dry |
115 and 116 |
| Chemical precipitation |
Nanosized ZnO |
3.20 |
97.36 |
Low-cost, eco-friendly, and efficient; produces highly active ZnO photocatalysts |
117 |
| Solvothermal |
Nanosized ZnO |
2.99 |
92.00 |
The solvothermal method produces high-purity, uniform ZnO photocatalysts but demands high energy, long reaction times, and precise control |
118 and 119 |
| Green synthesis |
Nanosized ZnO |
3.20 |
87.00 |
Environmentally friendly, low-cost, and highly efficient; variability in quality may occur depending on the plant source |
120 |
However, synthesized ZnO still faces several intrinsic limitations that hinder its photocatalytic efficiency. One of the primary challenges is the rapid recombination of photogenerated electron–hole pairs, which significantly reduces the number of charge carriers reaching the catalyst surface. This recombination not only slows down the redox reactions but also leads to energy loss in the form of heat. Additionally, ZnO's wide band gap restricts its light absorption to the UV region, limiting its effectiveness under visible light irradiation.9 Another drawback is the tendency of ZnO nanoparticles to agglomerate, which decreases the effective surface area available for light absorption and reactive species generation. These larger aggregates also reduce the photon flux within the reaction medium, further lowering the photocatalytic activity. Moreover, charge carrier recombination within these aggregates contributes to a significant decline in overall performance.
To overcome these limitations, various strategies have been developed to enhance the photocatalytic performance of ZnO. These include modifying its physical structure to increase surface area, doping with noble or non-noble metals to alter electronic properties, and forming heterojunctions with other semiconductors to promote charge separation. The deliberate introduction of structural defects has also been shown to improve light absorption and trap charge carriers.121 Additionally, coupling ZnO with plasmonic and photothermal materials can extend its light absorption range into the visible spectrum and enhance photocatalytic efficiency through localized surface plasmon resonance.122
7.1. Doping to modify photocatalysts
7.1.1. Doping with metal. Various transition metals such as silver (Ag), manganese (Mn), nickel (Ni), copper (Cu), and iron (Fe) have been utilized as dopants in ZnO; however, studies have shown that only optimal concentrations result can enhanced photocatalytic activity.123 By reducing the band gap, doping with metal enables ZnO to absorb visible light, which makes up 40–45% of sunlight. Furthermore, doping modifies the photocatalyst's optical and electrical characteristics.124 Consequently, ZnO doped with different metals is used in a variety of industries, such as paints, chemicals, tires, ceramics, pharmaceuticals, and agriculture. For example, scientists investigated the removal of chromium(VI) using a biosynthesized Ag-doped ZnO nanocomposite that contained activated carbon. After 60 hours of treatment, they discovered that the optimal adsorption took place at pH 2.5, with a concentration of 40 ppm of heavy metal ions.125 When Fe3+ ions are added to ZnO nanoparticles, more Zn2+ is produced, which increases surface defects and improves degradation performance.126 Ag-doped ZnO nanoparticles also show that noticeable surface defects improve degradation performance.127 Additional finding also suggests that aluminum (Al) increases hydrophilicity and surface defects in ZnO while cathodically shifting the quasi-Fermi level. By creating smaller particles, Al doping also increases specific surface area, which promotes the breakdown of organic pollutants.128 According to a study, hydroxyl ion absorption on nanoparticle surfaces is increased when high concentrations of magnesium (Mg) dopant are added to ZnO. As a result, carriers are efficiently trapped by abundant Mg2+ ions, which lowers recombination and speeds up degradation.129 At room temperature, Cu-doping with ZnO increases the effectiveness of both photocatalytic and antioxidant activities. These nanoparticles are also versatile for wastewater treatment applications due to their pH-dependent photocatalytic properties.130 It has been demonstrated that doping ZnO photocatalyst with rare earth metals, such as lanthanum (La), increases its efficiency. A study using La-doped ZnO photocatalysts to degrade 2-chlorophenol (2-CP) found that the degradation efficiency was higher when the catalyst dose was 10 mg and the irradiation period was 2 hours at an ideal pH of 2.131 Table 7 contains the list of some metal dopants that are used with ZnO for photocatalytic applications.
Table 7 Impact of metal dopants on improving the photodegradation capabilities of ZnO
| Doping metal |
Efficiency of ZnO (%) |
Efficiency of doped-ZnO (%) |
Contaminant degraded |
Ref. |
| Ag |
90.00 |
98.00 |
Carbaryl |
132 |
| La |
75.85 |
83.92 |
2-Chlorophenol |
131 |
| Cu |
58.50 |
96.97 |
Diazinon |
133 |
| Pd |
38.00 |
82.00 |
Acetamiprid |
134 |
| Pd + graphene oxide |
38.00 |
98.00 |
| WO3 |
27.00 |
78.00 |
2,4-Dichlorophenoxyacetic acid |
135 |
7.1.2. Doping with non-metal. Non-metal dopants such as carbon (C), nitrogen (N), fluorine (F), iodine (I), and sulphur (S) improve ZnO composite photocatalytic efficiency under visible light. They modify the bandgap by replacing oxygen vacancies, increasing surface oxygen vacancy defects. Their minute size enables diffusion through lattice interstices, binding to atoms via oxidation to aid the process.136 In contrast to metal ion dopants, non-metal ions have a lower likelihood of forming recombination centres. Therefore, they are more effective in enhancing photocatalytic activities. In the same context,137 explored the impact of N-doping on ZnO nanoparticles for degrading methylene blue dye, and observed superior efficacy compared to pure ZnO under both UV and visible light. The author explains how N-doping lowered excitation energy, creating new energy states near ZnO's VB, thereby increasing electron–hole pair production under visible light. This narrower band gap facilitated more straightforward electron transfer and intensified oxygen vacancies, facilitating rapid carrier separation and enhancing photocatalytic activity over three cycles. In a separate study,138 synthesized N-doped ZnO composites using a sol–gel combustion method, demonstrating excellent performance in degrading Eosin Yellow under visible light. XPS analysis confirmed the formation of N–Zn bonds, while improved photo response in the visible region resulted from N-induced lattice defects. Catalytic activity was increased by the composite's high surface area and porosity, which allowed for improved dye-catalyst contact. Likewise,139 investigated the photocatalytic degradation of p-aminobenzoic acid using C-doped ZnO nanorods as a catalyst. A cost-effective precipitation method was used to produce nanorods. When exposed to sunlight, 97% of the p-aminobenzoic acid was degraded under ideal circumstances (0.5 g per L catalyst dosage). The nanorods maintained their high photodegradation efficiency and showed signs of reusability.In another analysis,140 investigated I-modified ZnO's antibacterial qualities in the presence of light. With its cage-like structure and abundance of oxygen defects, the composite efficiently separated photoexcited charge carriers to produce more free radicals. Furthermore, iodine and the cage structure decreased electron–hole pair recombination, and photocatalytic performance was greatly enhanced by a smaller grain size. Additionally, by using differently shaped S-doped and F-doped ZnO,141 the photocatalytic degradation of methylene blue was compared. Under visible-light photocatalysis, these dopants changed the electron mobility rate and band gap structure. After six hours of exposure to visible light, S-doped ZnO showed better photoactivity (total removal) than F-doped ZnO. This was because S-doped ZnO had a smaller band gap and was more efficient at absorbing solar energy, which inhibited electron–hole recombination and increased photosensitive activity.
7.2. ZnO composites
Pairing ZnO with semiconductors can significantly boost its photocatalytic activity. This approach creates nanocomposites that improve light absorption, reduce recombination of charge carriers, and enhance charge separation. Combining ZnO with semiconductors that have different band gaps has shown great potential for improving photocatalysis.121 Ref. 142 studied that longer-lived charge carriers in these composites lead to more effective photo-degradation. While143 showed that adding CdS to ZnO changes its growth pattern, resulting in flower-like structures. The defects in the ZnO–CdS composite trap charge carriers, which reduce recombination and significantly enhance its ability to degrade rhodamine B (RhB).
Corresponding ref. 144 compared the degradation rate of chlorpyrifos, using MoO3/Al2O3 and ZnO·MoO3/Al2O3. The degradation kinetics showed that ZnO·MoO3/Al2O3 (95%) degraded chlorpyrifos 8 times faster than MoO3/Al2O3 (86%). In addition, intermediate products of chlorpyrifos were not identified by GC/MS analysis, which could be more toxic than chlorpyrifos itself, indicating a thorough mineralization of the pesticide,145 demonstrated that Fe–ZnO nanocomposite achieved 93.5% degradation of chlorpyrifos in 60 minutes under UV light. The enhanced performance is attributed to its reduced optical band gap and magnetic properties, making it effective for visible light photocatalysis. According to ref. 146, the β-CD–CuO/ZnO nanocomposite exhibits improved photocatalytic efficiency, degrading methylene blue by up to 89.15% and malachite green by 79.90% in 180 minutes when exposed to radiation. Its capacity to encapsulate pollutants in its hydrophobic interior and enhanced electron–hole recombination are the reasons for its superior performance.
Another nanocomposite, ZnO/BaBi2O6, demonstrated noticeably greater rates of methyl orange, RhB, and eosin degradation than pure ZnO, and it was able to degrade methyl orange by over 95% in 40 minutes. With a band gap energy of 2.89 eV and stability after five reuse cycles, h+ and radical O2− were found to be important components in the degradation process.147 A sustainable biogenic process was used to create the ZnO/multi-walled carbon nanotubes composite, which exhibits remarkable photocatalytic efficiency in eliminating hazardous pollutants and has the potential to be used for environmental remediation with high reusability.148 In another study,149 chemical precipitation was used, which creates free radicals at the conduction band (superoxide ion) and valence band (hydroxyl group) levels in Ag/Fe2O3/ZnO heterostructure, producing effective photocatalysis and superhydrophobic qualities. In another study, Cu-MOF/ZnO nanocomposite was found to exhibit high photocatalytic efficiency, degrading rose bengal by 97.4% in 45 minutes when exposed to natural sunlight. It is a promising material for environmental remediation because it retains stability and effective performance for up to five reuse cycles.150
7.3. Impacts of structural defect
The structure and form of ZnO, or morphology, have a significant impact on its photocatalytic capabilities. For example, although rods have a larger surface area, researchers have found that hexagonal or spherical ZnO nanoparticles often perform better than rod-shaped ones.151 In other cases, researchers have also emphasized the enhancement of the performance of needle-shaped152 or nanowire ZnO.153 Moreover, studies have also been examining structural flaws in addition to the morphology of ZnO nanoparticles. In particular, oxygen defects have emerged as a major focus of current photocatalysis research.154
In this context,155 investigated ZnO nanoparticles and how structural defects influence their photocatalytic efficiency in degrading RhB dye under UV light. Their findings revealed that the catalyst with the highest number of defects exhibited the greatest degradation performance. This improvement was attributed to an increased specific surface area resulting from the presence of more defects. Research suggests that zinc and oxygen vacancies can act as traps for charge carriers, which helps prevent the recombination of electron–hole pairs and thus enhances photocatalytic activity. For instance, ultrathin ZnO/Al2O3 nanosheets with a high density of oxygen defects were shown to rapidly degrade tetracycline and RhB dye within 150 minutes, achieving degradation rates of 88.4% and 76.9%, respectively.156 Similarly,157 synthesized ZnO/NiO nanoparticles with oxygen vacancies using solvothermal and high-temperature reduction methods to enhance their photocatalytic performance. The best solar-driven photocatalytic performance was shown by OZN-10, which almost degraded methylene blue due to its small size and large surface area. Because of its distinct structure and surface flaws, its efficiency was roughly double that of pure ZnO.
Ar–ZnO, which had the highest concentration of oxygen vacancies, showed a reduced bandgap of 3.03 eV, according to an experiment conducted by ref. 158 in which flower-like ZnO photocatalysts with porous nanosheets were prepared under various calcination atmospheres. This enhancement resulted in a 94.5% degradation of methyl orange under UV light in 30 minutes, with a degradation rate constant 3.2 times higher than ZnO calcined in air. Moreover, defect engineering in ZnO ceramics improved electrical conductivity and reduced thermal conductivity, enhancing their thermoelectric performance by increasing oxygen vacancies.159 Oxygen defects are known to boost photocatalytic efficiency and indicate that more defects lead to a larger surface area. However, some research suggests that oxygen and zinc defects can sometimes negatively affect ZnO's photocatalytic performance by serving as recombination centres, which reduces its effectiveness.98
8. Remediation of pesticides by ZnO photocatalysts
Over the years, there has been significant focus on utilizing ZnO and its nanoparticles as photocatalysts to break down pesticides. The majority of research on ZnO as a catalyst indicates that its photocatalytic efficacy in aqueous environments is heavily influenced by factors such as light source and intensity, reaction conditions, catalyst type, presence of oxidizing agents, solution pH, temperature, and pesticide concentration.160 Ref. 161 investigated the impact of the synthesis medium (ethanol and water) on the efficiency of ZnO/carbon xerogel photocatalysts to degrade 4-chlorophenol and bisphenol A. Carbon xerogel was chosen due to its electrical conductivity, surface area, and porosity. The maximum degradation rates achieved were 88% for 4-chlorophenol and 78% for bisphenol A after 5 hours. The photocatalytic mechanism relies heavily on the generation of OH˙, and the materials remained stable for up to three reuse cycles. The result shows that hybrid systems have shown greater efficiency in recent years when compared to pure systems. Another study162 shows that the GO–ZnO nanocomposite is excellent for breaking down the organophosphate pesticide quinalphos in water when exposed to UV light. This approach performs better than graphene GO nanosheets and ZnO nanoflowers. The nanocomposite exhibits pseudo-first-order kinetics and reaches a 98% degradation rate in 45 minutes at pH 6. Researchers found that OH˙ were the main active species in the degradation process after identifying smaller, innocuous byproducts through LC-MS analysis. The produced nanocomposite offers a workable way to degrade pesticides without requiring neutralization before being released into water bodies because it is stable and reusable for at least five cycles. Similarly,163 found that fungicide residues (difenoconazole and thifluzamide) in soil samples can be efficiently removed by chitosan–ZnO nanoparticles, which remain for days in the absence of these nanoparticles. In contrast, no activity was seen with chitosan–ZnO nanoparticle; photocatalytic studies demonstrated a significant increase in activity over a predetermined period. This indicates promising environmental remediation solutions and emphasizes the critical role that chitosan–ZnO nanoparticles play in promoting pesticide degradation.
Ref. 164 discovered that the pesticide chlorpyrifos is efficiently degraded by Ni-doped ZnO–TiO2 nanocomposites. These nanocomposites have a large surface area, distinct crystallinity, and good optical qualities. They convert chlorpyrifos into innocuous byproducts and function well in both visible and ultraviolet light. When exposed to UV light instead of darkness, the electrochemical analysis performed better. Degradation proceeds according to pseudo-first-order kinetics, with UV light causing higher rates (0.0221 min−1) than visible light (0.0088). In related work165 using a hydrothermal process, synthesize a SWAC/ZrO2–ZnO nanocomposite that efficiently breaks down 100 ppm of chlorpyrifos under UV light in 50 minutes at pH 6. A crystal size of 39.41 nm was confirmed by characterization techniques, and LC-MS analysis revealed that chlorpyrifos was fragmenting into smaller pieces. DFT simulations indicated the formation of reactive hydrogen bonds, and the degradation proceeded according to pseudo-first-order kinetics. According to166 study the NiO–ZnO nanocomposite's photocatalytic performance for breaking down the herbicide bentazon under UV light after 100 minutes of exposure. The study found that bentazon had a 70% degradation efficiency. This suggests that the nanocomposite may be useful in breaking down pesticides in water.167 Investigated the photocatalytic degradation of the pesticide lambda-cyhalothrin (LCY) using cerium-doped ZnO nanocomposites in the presence of natural sunlight. The coprecipitation method was used to synthesize Ce–ZnO, and methods like PXRD, SEM, FTIR, and EDAX were used to characterize its properties. The nanocomposites demonstrated high photocatalytic efficiency, degrading approximately 92% of LCY under ideal conditions (100 ppm initial concentration, 20 mg catalyst dose, and a UV index of 10–11) after three hours of exposure to sunlight. The average crystallite size was 31.42 nm. A pseudo-first-order kinetic model described the degradation.
The efficiency of photocatalytic degradation, as outlined in Table 8, is influenced by several factors, including the type of catalyst, the structure of the pesticide, the source of irradiation, and the conditions of the reaction. Modified ZnO systems, which may be doped, defect engineered, or formed as composites, demonstrate enhanced activity due to better charge separation and an optimized band structure. When parameters such as catalyst dosage, pH, and exposure time are carefully controlled, these well-designed ZnO catalysts can achieve rapid pesticide removal, typically exceeding 80–95% efficiency under UV or visible light. Ultimately, the interplay between the light source and the design of the catalyst plays a crucial role in determining the degradation performance across various pesticide classes.
Table 8 Degradation of different pesticides using ZnO nano-photocatalyst
| Photocatalyst |
Pesticide |
Efficiency (%) |
Light source |
Reaction condition |
Ref. |
| ZnO/αFe2O3 |
Carbamate |
89 |
Solar light |
Pesticide dose = 5 g L−1, catalyst dose = 1 g L−1, time = 3 h, pH = 8.5 |
18 |
| La-doped ZnO |
2-Chlorophenol |
83.92 |
Visible light |
Pesticide dose = 10 ppm, catalyst dose = 10 mg, time = 2 h |
131 |
| Ag-doped ZnO |
Carbaryl |
98 |
UV light |
Pesticide dose = 5 ppm, photocatalyst = 5 mg L−1, time = 60 min |
132 |
| ZnO NPs |
p-Nitrophenol |
92 |
UV light |
Pesticide dose = 20 mg L−1, catalyst dose = 1.5 g L−1, time = 180 min |
118 |
| ZnO |
Lambda-cyhalothrin |
87 |
Solar light |
Pesticide dose = 20 mL of 100 ppm, ZnO = 50 ppm, temperature = 31 °C, time = 30 min (dark), 60 min (light) |
168 |
| Fe–ZnO |
Chlorpyrifos |
67 |
Solar light |
Pesticide dose = 5 mg L−1, pH = 8, time = 140 min |
169 |
| Cu–ZnO heterostructure |
Chlorpyrifos |
95 |
Solar light |
Pesticide dose = 200 mg L−1, catalyst dose = 3 g L−1, pH = 6.0, time = 240 min |
170 |
| Ce–ZnO nanocomposites |
Lambda-cyhalothrin |
92 |
Solar light |
Pesticide dose = 100 ppm, catalyst dose = 20 mg, time = 3 h |
167 |
| Pbi–ZnO–g-C3N4 |
Atrazine |
85.3 |
Visible light |
Catalyst dose = 216.40 g L−1, time = 260 min |
171 |
| ZnO/rGO |
Metalaxyl |
90.25 |
UV light |
Pesticide dose = 10 mg L−1, catalyst dose = 0.75 g L−1, pH = 7, time = 120 min, UV intensity = 220 MW cm−2 |
172 |
| ZnO/rGO |
Metalaxyl (real agricultural runoff) |
51.17 |
UV light |
Pesticide dose = 10 mg L−1, catalyst dose = 0.75 g L−1, pH = 7, time = 120 min, UV intensity = 220 MW cm−2 |
| rGO/Fe3O4/ZnO |
Metalaxyl |
92.11 |
Visible light |
Time = 120 min, order of reaction = 1st order kinetic model |
173 |
| ZnO·WO3 composite |
Paraquat dichloride |
88.3 |
UV light |
Pesticide dose = 35 mg L−1, catalyst dose = 0.04 g, pH = 9, temperature = 40 °C, cycles of reaction = 3 |
174 |
| ZnO |
Methamidophos |
86.66 |
UV light |
Pesticide dose = 50 mg L−1, catalyst dose = 3 g L−1 with ultra-pure water |
175 |
| ZnO |
Methamidophos |
57.95 |
UV light |
Pesticide dose = 50 mg L−1, catalyst dose = 3 g L−1 with river water |
| Fe2O3–ZnO |
Profenofos |
100 |
Dark |
Pesticide dose = 1825 mg L−1, time = 60 min |
176 |
| CuO–ZnO nanocomposite |
Profenofos |
100 |
UV light |
Pesticide dose = 1215 mg L−1, time = 80 min |
177 |
| ZnO/Cu/GO |
Quinalphos |
99 |
Visible light |
Pesticide dose = 40 ppm, catalyst dose = 3 mg L−1, time = 20 min, pH = neutral |
178 |
| PANI/ZnO–CoMoO4 |
Imidacloprid |
97 |
Visible light |
Pesticide dose = 4.5 ppm, catalyst dose = 163.5 mg, time = 180 min, pH = 4 |
179 |
| ZnO/rGO |
Dimethoate |
99 |
UV light |
Pesticide dose = 5 mg L−1, catalyst dose = 50 mg, light intensity = 2.45 mW cm−2, time = 180 min |
180 |
| ZnO/CoFe2O4 |
Imidacloprid |
98.1 |
Visible light |
Pesticide dose = 5 ppm, catalyst dose = 0.05 g, pH = 10 |
181 |
9. Conclusion
This document reveals the distinctive electrical and structural properties of ZnO-based photocatalysts, which include a direct band gap, favorable band-edge positions, and defect-induced donor states. These characteristics contribute to their exceptional ability to degrade various pesticide contaminants. Although other photocatalytic materials offer certain benefits, ZnO surpasses them in oxidative strength, charge mobility, efficiency of electron–hole separation, and surface reactivity, especially when exposed to UV light. Additionally, recent advancements in ZnO engineering such as metal and non-metal doping, defect manipulation, heterojunction formation, plasmonic coupling, and photothermal integration have substantially enhanced its light-harvesting efficiency, charge transfer dynamics, and ROS generation. These enhancements effectively mitigate ZnO's inherent challenges, including photocorrosion and limited sensitivity to visible light, thereby improving its applicability in real-world wastewater treatment scenarios.
This document specifically pinpoints the potential environmental benefits of ZnO based nanomaterials in agriculture, particularly through photocatalytic processes that can eliminate pesticide-laden runoff, thereby minimizing groundwater contamination and protecting irrigation sources and aquatic ecosystems. This not only supports safe food production but also alleviates ecological pressures on soil microorganisms, aquatic flora, and beneficial insect populations. As the global community strives for low-impact and energy-efficient water treatment solutions, ZnO-based photocatalytic systems emerge as a promising avenue for developing sustainable, climate-resilient, and environmentally responsible remediation technologies.
Despite these promising results, current research on ZnO nanomaterials for wastewater treatment, particularly in pesticide degradation, reveals several critical areas or limitations for further exploration to improve their effectiveness and practical application. One key challenge is to enhance the long-term stability and reusability of ZnO photocatalysts, which are crucial for achieving sustainable and cost-effective solutions. While ZnO has shown potential in laboratory settings, there is a significant gap in studies that combine ZnO photocatalysis with other wastewater treatment techniques, such as biological treatments or membrane filtration, creating hybrid systems that integrate these methods could optimize pollutant removal and broaden the application of ZnO in real-world scenarios. Additionally, research into combining ZnO with 2D materials like black phosphorus and carbon nitride could enhance photocatalytic efficiency by improving light absorption and promoting effective charge separation, addressing some of ZnO's inherent limitations.
Future research should concentrate on integrating these strategies by advancing the design of ZnO nanostructures through controlled defects, facet engineering, and lattice strain, which can improve charge separation and broaden activity into the visible light spectrum. Additionally, combining ZnO with plasmonic metals, carbon materials, or narrow band-gap semiconductors can enhance solar absorption and increase degradation efficiency.
The photocorrosion of ZnO in acidic environments necessitates ongoing efforts to enhance its stability, with effective strategies including surface passivation, core–shell structures, and protective coatings. Additionally, green synthesis methods utilizing plant extracts or biopolymers promote environmentally friendly and scalable production processes. The use of immobilized ZnO in forms such as membranes, coatings, and 3D-printed structures is expected to become increasingly prevalent, enabling catalyst recovery and sustained operation in practical applications. Furthermore, integrating ZnO with photothermal materials, adsorption components, solar concentrators, or LED-powered systems presents promising opportunities for the remediation of agricultural runoff in real-world settings.
Conflicts of interest
Authors declare no conflict of interest.
Data availability
No primary research results, software or code have been included and no new data were generated or analysed as part of this review.
Acknowledgements
Dr Haji Muhammad expresses his gratitude to TUBITAK for granting him a visiting scientist fellowship under the “2221-Visiting Scientists Fellowship Programme 2023/6.” Additionally, Dr Asma Siddiqui extends her thanks to TWAS/TUBITAK for awarding her a fellowship through the 2216B Programme 2024/01.
References
- M. Tudi, et al., Agriculture development, pesticide application and its impact on the environment, Int. J. Environ. Res. Publ. Health, 2021, 18(3), 1112 CrossRef CAS PubMed.
- P. Kumar, et al., Impact of pesticides application on aquatic ecosystem and biodiversity: a review, Biol. Bull., 2023, 50(6), 1362–1375 CrossRef.
- S. Bose, et al., Microbial degradation of recalcitrant pesticides: a review, Environ. Chem. Lett., 2021, 19(4), 3209–3228 CrossRef CAS.
- L. Rani, et al., An extensive review on the consequences of chemical pesticides on human health and environment, J. Cleaner Prod., 2021, 283, 124657 CrossRef CAS.
- J. dos Santos Mendonça, et al., Mutagenicity, hepatotoxicity, and neurotoxicity of glyphosate and fipronil commercial formulations in Amazon turtles neonates (Podocnemis expansa), Sci. Total Environ., 2023, 898, 165529 CrossRef PubMed.
- J. H. O'Driscoll, et al., A risk ranking of pesticides in Irish drinking water considering chronic health effects, Sci. Total Environ., 2022, 829, 154532 CrossRef PubMed.
- W. Boedeker, et al., The global distribution of acute unintentional pesticide poisoning: estimations based on a systematic review, BMC Public Health, 2020, 20(1), 1–19 CrossRef PubMed.
- D. Rawtani, et al., Nanotechnology-based recent approaches for sensing and remediation of pesticides, J. Environ. Manage., 2018, 206, 749–762 CrossRef CAS PubMed.
- F. Abdullah, N. A. Bakar and M. A. Bakar, Current advancements on the fabrication, modification, and industrial application of zinc oxide as photocatalyst in the removal of organic and inorganic contaminants in aquatic systems, J. Hazard Mater., 2022, 424, 127416 CrossRef CAS PubMed.
- O. Ounas, B. Lekhlif and J. Jamal-eddine, The facile immobilization of ZnO into a polymer surface for photodegradation of organic contaminants, Mater. Today: Proc., 2020, 30, 816–822 CAS.
- K. Pszczolińska, H. Barchańska and D. Lalek, Comprehensive multiresidue chromatographic methods for monitoring pesticides in agricultural areas and corresponding plant protection zones, Environ. Pollut., 2024, 344, 123422 CrossRef PubMed.
- R. de Oliveira, W. da Silva Martini and A. C. Sant'Ana, Combined effect involving semiconductors and plasmonic nanoparticles in photocatalytic degradation of pesticides, Environ. Nanotechnol. Monit. Manag., 2022, 17, 100657 CAS.
- F. P. Carvalho, Pesticides, environment, and food safety, Food Energy Secur., 2017, 6(2), 48–60 CrossRef.
- C. Chang, et al., Current pesticide profiles in blood serum of adults in Jiangsu Province of China and a comparison with other countries, Environ. Int., 2017, 102, 213–222 CrossRef CAS PubMed.
- M. Tudi, et al., Exposure routes and health risks associated with pesticide application, Toxics, 2022, 10(6), 335 CrossRef CAS PubMed.
- M. Singhal, et al., Microalgae based sustainable bioremediation of water contaminated by pesticides, Biointerface Res. Appl. Chem., 2021, 12, 149–169 Search PubMed.
- G. K. Sidhu, et al., Toxicity, monitoring and biodegradation of organophosphate pesticides: a review, Crit. Rev. Environ. Sci. Technol., 2019, 49(13), 1135–1187 CrossRef CAS.
- A. Dehghan, et al., Green synthesis of ZnO/αFe2O3 nano-photocatalyst for efficient removal of carbamate pesticides in wastewater: optimization, mineralization, and financial analysis, Korean J. Chem. Eng., 2024, 41(1), 249–269 CrossRef CAS.
- M. Ahmad, et al., Photocatalytic Degradation of Deltamethrin in Drinking Water Under Visible Light by Using ZnO and TiO2, International Journal of Innovations in Science and Technology, 2025, 7, 57–67 Search PubMed.
- D. Kathuria, et al., A Review on Methods for the Elimination of Imidacloprid: 2018–2022, ChemistrySelect, 2023, 8(34), e202302293 CrossRef CAS.
- S.-M. Lam, et al., Sunlight responsive WO3/ZnO nanorods for photocatalytic degradation and mineralization of chlorinated phenoxyacetic acid herbicides in water, J. Colloid Interface Sci., 2015, 450, 34–44 CrossRef CAS PubMed.
- J. Fenoll, et al., Semiconductor-sensitized photodegradation of s-triazine and chloroacetanilide herbicides in leaching water using TiO2 and ZnO as catalyst under natural sunlight, J. Photochem. Photobiol., A, 2012, 238, 81–87 CrossRef CAS.
- S. Tariq, G. Chotana and A. Rashid, Photocatalytic degradation of paraquat dichloride in the presence of ZnO. WO3 composite, Int. J. Environ. Sci. Technol., 2022, 19(4), 2583–2598 CrossRef CAS.
- Q. Liu, et al., Phototransformation of phosphite induced by zinc oxide nanoparticles (ZnO NPs) in aquatic environments, Water Res., 2023, 245, 120571 CrossRef CAS PubMed.
- W. J. McCormick, et al., Enhanced monitoring of photocatalytic reactive oxygen species: using electrochemistry for rapid sensing of hydroxyl radicals formed during the degradation of coumarin, J. Phys. Chem. A, 2023, 127(23), 5039–5047 CrossRef CAS PubMed.
- S. Ali, et al., Environmental and health effects of pesticide residues, Sustainable Agriculture Reviews 48: Pesticide Occurrence, Analysis and Remediation, 2021, vol. 2 analysis, pp. 311–336 Search PubMed.
- A. Bagheri, N. Emami and C. A. Damalas, Monitoring point source pollution by pesticide use: an analysis of farmers' environmental behavior in waste disposal, Environment, Development and Sustainability: A Multidisciplinary Approach to the Theory and Practice of Sustainable Development, 2023, ch. 7, vol. 25, pp. 6711–6726 Search PubMed.
- X. Zhu, et al., Reductions in water, soil and nutrient losses and pesticide pollution in agroforestry practices: a review of evidence and processes, Plant Soil, 2020, 453(1), 45–86 CrossRef CAS.
- R. Wang, et al., Cleaner agricultural production in drinking-water source areas for the control of non-point source pollution in China, J. Environ. Manage., 2021, 285, 112096 CrossRef CAS PubMed.
- J. Yang, et al., Characteristics of non-point source pollution under different land use types, Sustainability, 2020, 12(5), 2012 CrossRef CAS.
- R. D. A. Acayaba, et al., Occurrence of pesticides in waters from the largest sugar cane plantation region in the world, Environ. Sci. Pollut. Res., 2021, 28, 9824–9835 CrossRef CAS PubMed.
- E. S. Severo, et al., Ecological risk of pesticide contamination in a Brazilian river located near a rural area: a study of biomarkers using zebrafish embryos, Ecotoxicol. Environ. Saf., 2020, 190, 110071 CrossRef CAS PubMed.
- N. L. Scholz, et al., A perspective on modern pesticides, pelagic fish declines, and unknown ecological resilience in highly managed ecosystems, Bioscience, 2012, 62(4), 428–434 CrossRef.
- K. Pohorecka, et al., The exposure of honey bees to pesticide residues in the hive environment with regard to winter colony losses, J. Apicult. Sci., 2017, 61(1), 105–125 CrossRef CAS.
- K. R. Hakeem, M. S. Akhtar and S. N. A. Abdullah, Plant, Soil and Microbes, Springer, 2016 Search PubMed.
- P. Rajak, et al., Agricultural pesticides–Friends or foes to biosphere?, J. Hazard. Mater. Adv., 2023, 10, 100264 CAS.
- M. S. Sankhla, et al., Water contamination through pesticide & their toxic effect on human health, Int. J. Res. Appl. Sci. Eng. Technol., 2018, 6(1), 967–970 CrossRef.
- M. Ataei and M. Abdollahi, A systematic review of mechanistic studies on the relationship between pesticide exposure and cancer induction, Toxicol. Appl. Pharmacol., 2022, 116280 CrossRef CAS PubMed.
- K. G. Bernal-González, et al., Organophosphate-Pesticide-Mediated Immune Response Modulation in Invertebrates and Vertebrates, Int. J. Mol. Sci., 2023, 24(6), 5360 CrossRef PubMed.
- A. Fucic, et al., Reproductive health risks associated with occupational and environmental exposure to pesticides, Int. J. Environ. Res. Publ. Health, 2021, 18(12), 6576 CrossRef CAS PubMed.
- W. Lai, Pesticide use and health outcomes: evidence from agricultural water pollution in China, J. Environ. Econ. Manag., 2017, 86, 93–120 CrossRef.
- A. G. Akerdi and S. H. Bahrami, Application of heterogeneous nano-semiconductors for photocatalytic advanced oxidation of organic compounds: a review, J. Environ. Chem. Eng., 2019, 7(5), 103283 CrossRef CAS.
- A. Boudebbouz, et al., Pesticide residues levels in raw cow's milk and health risk assessment across the globe: a systematic review, Environ. Adv., 2022, 9, 100266 CrossRef CAS.
- S. Lorn, et al., Pyrethroid susceptibility in Stomoxys calcitrans and Stomoxys indicus (Diptera: Muscidae) collected from cattle farms in southern Thailand, Insects, 2022, 13(8), 711 CrossRef PubMed.
- R. Giugliano, et al., Monitoring of Non-Maximum-Residue-Level Pesticides in Animal Feed: A Study from 2019 to 2023, Toxics, 2024, 12(9), 680 CrossRef CAS PubMed.
- S. Bogialli, et al., Development of a multiresidue method for analyzing herbicide and fungicide residues in bovine milk based on solid-phase extraction and liquid chromatography–tandem mass spectrometry, J. Chromatogr. A, 2006, 1102(1–2), 1–10 CrossRef CAS PubMed.
- M. Miclean, E. A. Levei and O. Cadar, Organochlorine Pesticides in Dairy Cows' Diet and the Carryover into Milk in NW Romania, Sustainability, 2024, 16(1), 434 CrossRef CAS.
- L. Zhao, et al., The Associations between Organophosphate Pesticides (OPs) and Respiratory Disease, Diabetes Mellitus, and Cardiovascular Disease: A Review and Meta-Analysis of Observational Studies, Toxics, 2023, 11(9), 741 CrossRef CAS PubMed.
- J. Kaushal, M. Khatri and S. K. Arya, A treatise on Organophosphate pesticide pollution: current strategies and advancements in their environmental degradation and elimination, Ecotoxicol. Environ. Saf., 2021, 207, 111483 CrossRef CAS PubMed.
- J. Perry, et al., Organophosphate exposure and the chronic effects on farmers: a narrative review, Rural Remote Health, 2020, 20(1), 206–222 Search PubMed.
- R. Jayaraj, P. Megha and P. Sreedev, Organochlorine pesticides, their toxic effects on living organisms and their fate in the environment, Interdiscip. Toxicol., 2016, 9(3–4), 90 CrossRef CAS PubMed.
- M. A. Mohammadkhani, et al., Insights into Organochlorine Pesticides Exposure in the Development of Cardiovascular Diseases: A Systematic Review, Arch. Iran. Med., 2023, 26(10), 592 CrossRef PubMed.
- G.-H. Lee and K.-C. Choi, Adverse effects of pesticides on the functions of immune system, Comp. Biochem. Physiol., Part C:Toxicol. Pharmacol., 2020, 235, 108789 CAS.
- S. Moreira, et al., Carbamate pesticides: shedding light on their impact on the male reproductive system, Int. J. Mol. Sci., 2022, 23(15), 8206 CrossRef CAS PubMed.
- A. Chrustek, et al., Current research on the safety of pyrethroids used as insecticides, Medicina, 2018, 54(4), 61 CrossRef PubMed.
- I. Hołyńska-Iwan and K. Szewczyk-Golec, Pyrethroids: how they affect human and animal health?, Medicina, 2020, 56(11), 582 CrossRef PubMed.
- A.-M. Saillenfait, D. Ndiaye and J.-P. Sabaté, Pyrethroids: exposure and health effects–an update, Int. J. Hyg Environ. Health, 2015, 218(3), 281–292 CrossRef CAS PubMed.
- X. Chen, et al., Simultaneous removal of triadimefon and dinotefuran by a new biochar-based magnesium oxide composite in water: performances and mechanism, Sep. Purif. Technol., 2024, 336, 126213 CrossRef CAS.
- Z.-H. Diao, et al., Removal of herbicide atrazine by a novel biochar based iron composite coupling with peroxymonosulfate process from soil: synergistic effect and mechanism, Chem. Eng. J., 2021, 409, 127684 CrossRef CAS.
- N. M. Islam, et al., Development of a novel catalytic membrane reactor for heterogeneous catalysis in supercritical CO2, Int. J. Mol. Sci., 2010, 11(1), 164–172 CrossRef CAS PubMed.
- D. Sud and P. Kaur, Heterogeneous photocatalytic degradation of selected organophosphate pesticides: a review, Crit. Rev. Environ. Sci. Technol., 2012, 42(22), 2365–2407 CrossRef CAS.
- S. Ahmed, et al., Recent developments in physical, biological, chemical, and hybrid treatment techniques for removing emerging contaminants from wastewater, J. Hazard Mater., 2021, 416, 125912 CrossRef CAS PubMed.
- S. Gangola, et al., Biotechnological tools to elucidate the mechanism of pesticide degradation in the environment, Chemosphere, 2022, 296, 133916 CrossRef CAS PubMed.
- J. Colina-Márquez, F. Machuca-Martínez and W. Salas, Enhancement of the potential biodegradability and the mineralization of a pesticides mixture after being treated by a coupled process of TiO2-based solar photocatalysis with constructed wetlands, Ing. Compet., 2013, 15(2), 181–190 Search PubMed.
- G. R. M. Echavia, F. Matzusawa and N. Negishi, Photocatalytic degradation of organophosphate and phosphonoglycine pesticides using TiO2 immobilized on silica gel, Chemosphere, 2009, 76(5), 595–600 CrossRef CAS PubMed.
- Z. Zihan, et al., Photocatalytic degradation of an organophosphorus pesticide using a ZnO/rGO composite, RSC Adv., 2020, 10, 11929–11938 RSC.
- J. M. Poyatos, et al., Advanced oxidation processes for wastewater treatment: state of the art, Water, Air, Soil Pollut., 2010, 205, 187–204 CrossRef CAS.
- A. Oliveira, et al., Solar photo-chemistry for environmental remediation-advanced oxidation processes for industrial wastewater treatment, Molecular Photochemistry-Various Aspects, InTech, Rijeka, 2012, pp. 195–223 Search PubMed.
- M. Saad, et al., Ultrasonically enhanced photocatalytic degradation of methylene blue by nano-CoFe2O4-immobilized Saccharomyces cerevisiae yeast composite as a photo-Fenton catalyst: a central composite design study, Int. J. Biol. Macromol., 2025, 147068 CrossRef CAS PubMed.
- A. Shamim, et al., Removal of pesticide pollutants from aqueous waste utilizing nanomaterials via photocatalytic process: a review, Int. J. Environ. Sci. Technol., 2024, 21(4), 4653–4684 CrossRef CAS.
- J. Wang and R. Zhuan, Degradation of antibiotics by advanced oxidation processes: an overview, Sci. Total Environ., 2020, 701, 135023 CrossRef CAS PubMed.
- A. Hossain, et al., Kinetics of degradation of eosin Y by one of the advanced oxidation processes (AOPs)—Fenton’s process, Am. J. Anal. Chem., 2016, 7(12), 863–879 CrossRef CAS.
- U. Hübner, et al., Advanced oxidation processes for water and wastewater treatment–guidance for systematic future research, Heliyon, 2024, 10(9), e30402 CrossRef PubMed.
- G. Lu, et al., Recent advances in metal-organic frameworks-based materials for photocatalytic selective oxidation, Coord. Chem. Rev., 2022, 450, 214240 CrossRef CAS.
- T. Katagi, Direct photolysis mechanism of pesticides in water, J. Pestic. Sci., 2018, 43(2), 57–72 CrossRef CAS PubMed.
- S. S. Patil, et al., Photocatalytic degradation of organic dyes and binary mixture of dyes by Fe and Cu doped ZnO nanoparticles under artificial light and sunlight, Inorg. Chem. Commun., 2025, 172, 113751 CrossRef CAS.
- J. Schneider, et al., Photocatalysis: Fundamentals and Perspectives, Royal Society of Chemistry, 2016 Search PubMed.
- M. R. D. Khaki, et al., Application of doped photocatalysts for organic pollutant degradation-a review, J. Environ. Manag., 2017, 198, 78–94 CrossRef CAS PubMed.
- G. Ren, et al., Recent advances of photocatalytic application in water treatment: a review, Nanomaterials, 2021, 11(7), 1804 CrossRef CAS PubMed.
- M. Devi, S. Praharaj and D. Rout, Industrial problems and solution towards visible light photocatalysis, in Nanostructured Materials for Visible Light Photocatalysis, Elsevier, 2022, pp. 535–567 Search PubMed.
- M. Mrowetz, et al., Oxidative power of nitrogen-doped TiO2 photocatalysts under visible illumination, J. Phys. Chem. B, 2004, 108(45), 17269–17273 CrossRef CAS.
- Z. Moradi, S. Z. Jahromi and M. Ghaedi, Design of active photocatalysts and visible light photocatalysis, in Interface Science and Technology, Elsevier, 2021, pp. 557–623 Search PubMed.
- A. Samal, et al., Photocatalytic degradation and kinetic investigations of ZnO-SnO2 heterostructures for treatment of methyl violet using non-conventional approach, Inorg. Chem. Commun., 2024, 159, 111809 CrossRef CAS.
- V. Kitsiou, et al., Heterogeneous and homogeneous photocatalytic degradation of the insecticide imidacloprid in aqueous solutions, Appl. Catal., B, 2009, 86(1–2), 27–35 CrossRef CAS.
- S. I. Wani and A. S. Ganie, Ag2O incorporated ZnO− TiO2 nanocomposite: ionic conductivity and photocatalytic degradation of an organic dye, Inorg. Chem. Commun., 2021, 128, 108567 CrossRef CAS.
- M. Hadei, et al., A comprehensive systematic review of photocatalytic degradation of pesticides using nano TiO 2, Environ. Sci. Pollut. Res., 2021, 28, 13055–13071 CrossRef CAS PubMed.
- K. Balkus, Metal oxide nanotube, nanorod, and quantum dot photocatalysis, New and Future Developments in Catalysis, 2013, pp. 213–244 Search PubMed.
- D. K. Sharma, et al., A review on ZnO: fundamental properties and applications, Mater. Today: Proc., 2022, 49, 3028–3035 CAS.
- S. Munir, S. M. Shah and H. Hussain, Effect of carrier concentration on the optical band gap of TiO2 nanoparticles, Mater. Des., 2016, 100(92), 64–72 CrossRef.
- T. Zhang, et al., Tunning the bandgap of MnO2 homojunction by building active high-index facet to achieve rapid electron transfer for enhanced photocatalytic sterilization, J. Mater. Sci. Technol., 2024, 168, 265–275 CrossRef CAS.
- K. A. Mohammed, et al., Capping agent effect on optical properties of Fe2O3 nanoparticles, Mater. Today: Proc., 2022, 56, 2010–2015 CAS.
- S. Pazouki and N. Memarian, Effects of Hydrothermal temperature on the physical properties and anomalous band gap behavior of ultrafine SnO2 nanoparticles, Optik, 2021, 246, 167843 CrossRef CAS.
- A. Dilshad, et al., Araucaria cunninghamii Linn leaves silver nanoparticles for simultaneous naked-eye detection of iron II and III, Moroc. J. Chem., 2025, 13(3), 1290 Search PubMed.
- H. Kisch, Semiconductor photocatalysis—mechanistic and synthetic aspects, Angew. Chem., Int. Ed., 2013, 52(3), 812–847 CrossRef CAS PubMed.
- N. Premalatha and P. Rex, A comprehensive review on photocatalytic degradation of organophosphorus pesticide using ZnO coupled photocatalysts, Desalination Water Treat., 2024, 320, 100753 CrossRef CAS.
- W. Li, Photocatalysis of oxide semiconductors, J. Aust. Ceram. Soc., 2013, 41–46 CAS.
- S. H. Khan and B. Pathak, Zinc oxide based photocatalytic degradation of persistent pesticides: a comprehensive review, Environ. Nanotechnol., Monit. Manage., 2020, 13, 100290 Search PubMed.
- V. Gurylev and T. P. Perng, Defect engineering of ZnO: review on oxygen and zinc vacancies, J. Eur. Ceram. Soc., 2021, 41(10), 4977–4996 CrossRef CAS.
- P. Veerakumar, et al., Palladium and silver nanoparticles embedded on zinc oxide nanostars for photocatalytic degradation of pesticides and herbicides, Chem. Eng. J., 2021, 410, 128434 CrossRef CAS.
- A. Saifaldeen and E. E. Al-Abodi, Exciting new nan composite by using the solar energy to remove BG dye pollutants from waste water, in AIP Conference Proceedings, AIP Publishing, 2021 Search PubMed.
- A. A. Shaikh, et al., Synthesis and characterization of Ag doped ZnO nanomaterial as an effective photocatalyst for photocatalytic degradation of Eriochrome Black T dye and antimicrobial agent, Inorg. Chem. Commun., 2023, 151, 110570 CrossRef.
- N. Miguel, et al., Photocatalytic degradation of pesticides in natural water: effect of hydrogen peroxide, Int. J. Photoenergy, 2012, 2012(1), 371714 Search PubMed.
- S. Liang, et al., A novel ZnO nanoparticle blended polyvinylidene fluoride membrane for anti-irreversible fouling, J. Membr. Sci., 2012, 394, 184–192 CrossRef.
- M. Kouhail, et al., A Comparative study between TiO2 and ZnO photocatalysis: photocatalytic degradation of textile dye, in IOP Conference Series: Materials Science and Engineering, IOP Publishing, 2020 Search PubMed.
- D. Štrbac, et al., Photocatalytic degradation of Naproxen and methylene blue: comparison between ZnO, TiO2 and their mixture, Process Saf. Environ. Prot., 2018, 113, 174–183 CrossRef.
- S. Shahzad, S. Javed and M. Usman, A review on synthesis and optoelectronic applications of nanostructured ZnO, Front. Mater., 2021, 8, 613825 CrossRef.
- H. AlMohamadi, et al., Photocatalytic activity of metal-and non-metal-anchored ZnO and TiO2 nanocatalysts for advanced photocatalysis: comparative study, Catalysts, 2024, 14(7), 420 CrossRef CAS.
- M. Shekofteh-Gohari, et al., Magnetically separable nanocomposites based on ZnO and their applications in photocatalytic processes: a review, Crit. Rev. Environ. Sci. Technol., 2018, 48(10–12), 806–857 CrossRef CAS.
- M. Bhattu, R. Acevedo and A. Shnain, A comprehensive review on the synthesis routes, properties and potential applications of ZnFe2O4 ferrites, in E3S Web of Conferences, EDP Sciences, 2024 Search PubMed.
- N. T. T. Truc, et al., The advanced photocatalytic degradation of atrazine by direct Z-scheme Cu doped ZnO/g-C3N4, Appl. Surf. Sci., 2019, 489, 875–882 CrossRef CAS.
- Y. N. Kanafin, et al., A review on WO3 photocatalysis used for wastewater treatment and pesticide degradation, Heliyon, 2025, 11, e40788 CrossRef CAS PubMed.
- H. J. Jung, et al., Enhanced photocatalytic degradation of lindane using metal–semiconductor Zn@ ZnO and ZnO/Ag nanostructures, J. Environ. Sci., 2018, 74, 107–115 CrossRef CAS PubMed.
- T. Sansenya, et al., Hydrothermal synthesis of ZnO photocatalyst for detoxification of anionic azo dyes and antibiotic, J. Phys. Chem. Solids, 2022, 160, 110353 CrossRef CAS.
- E. Prabakaran and K. Pillay, Synthesis of N-doped ZnO nanoparticles with cabbage morphology as a catalyst for the efficient photocatalytic degradation of methylene blue under UV and visible light, RSC Adv., 2019, 9(13), 7509–7535 RSC.
- R. Saravanan, et al., Comparative study on photocatalytic activity of ZnO prepared by different methods, J. Mol. Liq., 2013, 181, 133–141 CrossRef CAS.
- N. T. Nguyen and V. A. Nguyen, Synthesis, characterization, and photocatalytic activity of ZnO nanomaterials prepared by a green, nonchemical route, J. Nanomater., 2020, 2020(1), 1768371 Search PubMed.
- Y. Kedruk, et al., Facile low-cost synthesis of highly photocatalytically active zinc oxide powders, Front. Mater., 2022, 9, 869493 CrossRef.
- Y. Wang, et al., Solvothermal synthesis of ZnO nanoparticles for photocatalytic degradation of methyl orange and p-nitrophenol, Water, 2021, 13(22), 3224 CrossRef CAS.
- L. Ndlwana, et al., Sustainable hydrothermal and solvothermal synthesis of advanced carbon materials in multidimensional applications: a review, Materials, 2021, 14(17), 5094 CrossRef CAS PubMed.
- S. A. Mousa, et al., Enhanced photocatalytic activity of green synthesized zinc oxide nanoparticles using low-cost plant extracts, Sci. Rep., 2024, 14(1), 16713 CrossRef CAS PubMed.
- S. Abou Zeid and Y. Leprince-Wang, Advancements in ZnO-Based Photocatalysts for Water Treatment: A Comprehensive Review, Crystals, 2024, 14(7), 611 CrossRef CAS.
- A. Hezam, et al., Smart plasmonic Ag/Ag2O/ZnO nanocomposite with promising photothermal and photodynamic antibacterial activity under 600 nm visible light illumination, J. Photochem. Photobiol., A, 2023, 435, 114322 CrossRef CAS.
- M. R. Shakil, et al., Single-doped and multidoped transition-metal (Mn, Fe, Co, and Ni) ZnO and their electrocatalytic activities for oxygen reduction reaction, Inorg. Chem., 2018, 57(16), 9977–9987 CrossRef CAS PubMed.
- K. Rekha, et al., Structural, optical, photocatalytic and antibacterial activity of zinc oxide and manganese doped zinc oxide nanoparticles, Phys. B, 2010, 405(15), 3180–3185 CrossRef CAS.
- A. Taha, E. Da'na and H. A. Hassanin, Modified activated carbon loaded with bio-synthesized Ag/ZnO nanocomposite and its application for the removal of Cr (VI) ions from aqueous solution, Surf. Interfaces, 2021, 23, 100928 CrossRef CAS.
- M. L. Levy, et al., Key recommendations for primary care from the 2022 Global Initiative for Asthma (GINA) update, npj Prim. Care Respir. Med., 2023, 33(1), 7 CrossRef PubMed.
- S. Iqbal, et al., Design Ag-doped ZnO heterostructure photocatalyst with sulfurized graphitic C3N4 showing enhanced photocatalytic activity, Mater. Sci. Eng., B, 2021, 272, 115320 CrossRef CAS.
- F. Ajala, et al., The influence of Al doping on the photocatalytic activity of nanostructured ZnO: the role of adsorbed water, Appl. Surf. Sci., 2018, 445, 376–382 CrossRef CAS.
- R. E. Adam, et al., Synthesis of Mg-doped ZnO NPs via a chemical low-temperature method and investigation of the efficient photocatalytic activity for the degradation of dyes under solar light, Solid State Sci., 2020, 99, 106053 CrossRef CAS.
- R. Sonkar, et al., Cu doped ZnO nanoparticles: correlations between tuneable optoelectronic, antioxidant and photocatalytic activities, J. Phys. Chem. Solids, 2024, 185, 111715 CrossRef CAS.
- S. Rani, et al., Highly Efficient Photocatalytic Properties of La-Doped ZnO over Pristine ZnO for Degradation of 2-Chlorophenol from Aquatic Agriculture Waste, Chem. Afr., 2023, 6(4), 1981–1990 CrossRef CAS.
- A. Satheesh, et al., UV Assisted Enhanced Photodegradation of Carbaryl Pesticide with Ag Doped ZnO NPs, J. Sci. Res., 2024, 16(2), 561–573 CrossRef CAS.
- M. Shirzad-Siboni, et al., Enhancement of photocatalytic activity of Cu-doped ZnO nanorods for the degradation of an insecticide: kinetics and reaction pathways, J. Environ. Manag., 2017, 186, 1–11 CrossRef CAS PubMed.
- C. Sayury Miyashiro and S. Hamoudi, Palladium and graphene oxide doped ZnO for aqueous acetamiprid degradation under visible light, Catalysts, 2022, 12(7), 709 CrossRef CAS.
- Y. Zandsalimi, et al., Photocatalytic removal of 2, 4-dichlorophenoxyacetic acid from aqueous solution using tungsten oxide doped zinc oxide nanoparticles immobilised on glass beads, Environ. Technol., 2022, 43(5), 631–645 CrossRef CAS PubMed.
- C. Di Valentin and G. Pacchioni, Trends in non-metal doping of anatase TiO2: B, C, N and F, Catal. Today, 2013, 206, 12–18 CrossRef CAS.
- C. Tang, et al., Enhancement of degradation for nitrogen doped zinc oxide to degrade methylene blue, Phys. B, 2020, 583, 412029 CrossRef CAS.
- S. Ramachandran and A. Sivasamy, Synthesis and characterization of nanocrystalline N-doped semiconductor metal oxide and its visible photocatalytic activity in the degradation of an organic dye, J. Environ. Chem. Eng., 2018, 6(3), 3770–3779 CrossRef CAS.
- P. Perillo and M. Atia, C-doped ZnO nanorods for photocatalytic degradation of p-aminobenzoic acid under sunlight, Nano-Struct. Nano-Objects, 2017, 10, 125–130 CrossRef CAS.
- Y. Jin, et al., Synthesis of caged iodine-modified ZnO nanomaterials and study on their visible light photocatalytic antibacterial properties, Appl. Catal., B, 2019, 256, 117873 CrossRef CAS.
- R. J. Ramalingam, et al., Synthesis, characterization and optical properties of sulfur and fluorine doped ZnO nanostructures for visible light utilized catalysis, Optik, 2017, 148, 325–331 CrossRef.
- C.-C. Lin and Y.-J. Chiang, Preparation of coupled ZnO/SnO2 photocatalysts using a rotating packed bed, Chem. Eng. J., 2012, 181, 196–205 CrossRef.
- I. Zgura, et al., Wet chemical synthesis of ZnO-CdS composites and their photocatalytic activity, Mater. Res. Bull., 2018, 99, 174–181 CrossRef CAS.
- F. Attaria, et al., Photocatalytic Removal of the Endocrine Disruptor Chlorpyrifos using a ZnO. MoO3 Composite Supported on Al2O3, ChemistrySelect, 2024, 9(23), e202400627 CrossRef CAS.
- S. H. Khan, B. Pathak and M. Fulekar, Synthesis, characterization and photocatalytic degradation of chlorpyrifos by novel Fe: ZnO nanocomposite material, Nanotechnol. Environ. Eng., 2018, 3, 1–14 CrossRef CAS.
- R. Yadav, et al., Photocatalytic degradation of textile dyes using β-CD-CuO/ZnO nanocomposite, J. Phys. Chem. Solids, 2022, 165, 110691 CrossRef CAS.
- P. A. Roozbahani, H. Behnejad and M. Hamzehloo, Synthesis, analysis, and application of zinc oxide and bismuth-based nanocomposites (ZnO/BaBi2O6) as a photocatalyst for organic dyes degradation, Inorg. Chem. Commun., 2024, 168, 112821 CrossRef CAS.
- N. Mishra, et al., Enhanced photocatalytic degradation of hazardous dyes under visible light with biogenically synthesized ZnO-decked multi-walled carbon nanotubes (ZnO/MWCNT) nanocomposite: catalyst fabrication, performance and mechanistic insight, Nano-Struct. Nano-Objects, 2024, 39, 101200 CrossRef CAS.
- M. I. Rahmah, R. S. Sabry and W. J. Aziz, Preparation of superhydrophobic Ag/Fe2O3/ZnO surfaces with photocatalytic activity, Surf. Eng., 2021, 37(10), 1320–1327 CrossRef CAS.
- S. Roy, J. Darabdhara and M. Ahmaruzzaman, ZnO-based Cu metal–organic framework (MOF) nanocomposite for boosting and tuning the photocatalytic degradation performance, Environ. Sci. Pollut. Res., 2023, 30(42), 95673–95691 CrossRef CAS PubMed.
- I. Daou, O. Zegaoui and A. Elghazouani, Physicochemical and photocatalytic properties of the ZnO particles synthesized by two different methods using three different precursors, C. R. Chim., 2017, 20(1), 47–54 CrossRef CAS.
- F. M. Collares, et al., Exploring needle-like zinc oxide nanostructures for improving dental resin sealers: design and evaluation of antibacterial, physical and chemical properties, Polymers, 2020, 12(4), 789 CrossRef CAS PubMed.
- M. Kwoka, et al., Novel insight on the local surface properties of ZnO nanowires, Nanotechnology, 2020, 31(46), 465705 CrossRef CAS PubMed.
- I. Ayoub, et al., Advances in ZnO: manipulation of defects for enhancing their technological potentials, Nanotechnol. Rev., 2022, 11(1), 575–619 CrossRef CAS.
- P. Nandi and D. Das, Photocatalytic degradation of rhodamine-B dye by stable ZnO nanostructures with different calcination temperature induced defects, Appl. Surf. Sci., 2019, 465, 546–556 CrossRef CAS.
- Q. Nie, et al., High piezo–catalytic activity of ZnO/Al2O3 nanosheets utilizing ultrasonic energy for wastewater treatment, J. Cleaner Prod., 2020, 242, 118532 CrossRef CAS.
- J. Zhang and J. Li, The oxygen vacancy defect of ZnO/NiO nanomaterials improves photocatalytic performance and ammonia sensing performance, Nanomaterials, 2022, 12(3), 433 CrossRef CAS PubMed.
- T. Bi, et al., Preparation of flower-like ZnO photocatalyst with oxygen vacancy to enhance the photocatalytic degradation of methyl orange, Appl. Surf. Sci., 2023, 614, 156240 CrossRef CAS.
- A. T. T. Pham, et al., Oxygen vacancy-activated thermoelectric properties of ZnO ceramics, Ceram. Int., 2024, 50(2), 3511–3518 CrossRef CAS.
- B. Rajeev, S. Yesodharan and E. Yesodharan, Sunlight activated ZnO mediated photocatalytic degradation of acetophenone in water, IOSR J. Appl. Chem., 2016, 55–70 CAS.
- N. P. de Moraes, et al., Effect of synthesis medium on structural and photocatalytic properties of ZnO/carbon xerogel composites for solar and visible light degradation of 4-chlorophenol and bisphenol A, Colloids Surf., A, 2020, 584, 124034 CrossRef CAS.
- R. Garg, et al., Eliminating pesticide quinalphos from surface waters using synthesized GO-ZnO nanoflowers: characterization, degradation pathways and kinetic study, Chemosphere, 2022, 286, 131837 CrossRef CAS PubMed.
- W. M. Daqa, et al., Potential Applications of Chitosan-Coated Zinc Oxide Nanoparticles for Degrading Pesticide Residues in Environmental Soils, Crystals, 2023, 13(3), 391 CrossRef CAS.
- H. Esfandian, M. R. Cherati and M. Khatirian, Electrochemical behavior and photocatalytic performance of chlorpyrifos pesticide decontamination using Ni-doped ZnO-TiO2 nanocomposite, Inorg. Chem. Commun., 2023, 111750 Search PubMed.
- S. L. Ezung, et al., Photocatalytic degradation of the organophosphorus insecticide chlorpyrifos in aqueous suspensions using a novel activated carbon ZrO2-ZnO nanocomposite under UV light, Korean J. Chem. Eng., 2023, 40(6), 1360–1372 CrossRef CAS.
- S. Yasmeen, et al., Photocatalytic Degradation of Organic Pollutants—Nile Blue, Methylene Blue, and Bentazon Herbicide—Using NiO-ZnO Nanocomposite, Nanomaterials, 2024, 14(5), 470 CrossRef CAS PubMed.
- M. D. Sani, V. K. Abbaraju and N. Venugopal, Synthesis of Cerium-Doped Zinc Oxide Nanocomposites and Their Application for Photocatalytic Degradation of Lambda-Cyhalothrin in Agricultural Runoff Under Natural Solar Irradiation, Int. J. Nanosci., 2024, 2350089 CrossRef CAS.
- S. Mahadi Danjuma, V. D. N. Kumar Abbaraju and N. V. S. Venugopal, Photocatalytic Degradation of Lambda-Cyhalothrin (Pyrethroid) in Wastewater using Zinc Oxide Nanoparticle, Rasayan J. Chem., 2023, 16(3), 1396–1402 CrossRef.
- N. S. Shah, et al., Enhanced solar light photocatalytic performance of Fe-ZnO in the presence of H2O2, S2O82−, and HSO5− for degradation of chlorpyrifos from agricultural wastes: toxicities investigation, Chemosphere, 2022, 287, 132331 CrossRef CAS PubMed.
- D. Pathania, et al., Bio-synthesized Cu–ZnO hetero-nanostructure for catalytic degradation of organophosphate chlorpyrifos under solar illumination, Chemosphere, 2021, 277, 130315 CrossRef CAS PubMed.
- X. An, et al., Core-shell P-laden biochar/ZnO/g-C3N4 composite for enhanced photocatalytic degradation of atrazine and improved P slow-release performance, J. Colloid Interface Sci., 2022, 608, 2539–2548 CrossRef CAS PubMed.
- S. Dehghan, et al., Preparation and photocatalytic performance of reduced graphene oxide/ZnO nanocatalyst for degradation of metalaxyl from aqueous solution: effect of operational parameters, mineralisation and toxicity bioassay, Int. J. Environ. Anal. Chem., 2022, 102(18), 7112–7134 CrossRef CAS.
- S. Dehghan, et al., Visible-light-driven photocatalytic degradation of metalaxyl by reduced graphene oxide/Fe3O4/ZnO ternary nanohybrid: influential factors, mechanism and toxicity bioassay, J. Photochem. Photobiol., A, 2019, 375, 280–292 CrossRef CAS.
- S. Tariq, G. Chotana and A. Rashid, Photocatalytic degradation of paraquat dichloride in the presence of ZnO. WO 3 composite, Int. J. Environ. Sci. Technol., 2021, 1–16 Search PubMed.
- H. R. Yucra-Condori, et al., Photocatalytic degradation of methamidophos in water using zinc oxide as a photocatalyst, Energy Nexus, 2024, 15, 100317 CrossRef CAS.
- S. Ullah, et al., Efficient removal of the organophosphate pesticide, profenofos using polymer-stabilized microporous Fe2O3-ZnO nanocomposite: kinetic and thermodynamic analysis, Appl. Surf. Sci., 2024, 662, 160027 CrossRef CAS.
- S. Ullah, et al., Efficient photocatalytic degradation of profenofos by CuO-ZnO nanocomposite, J. Photochem. Photobiol., A, 2024, 115787 CrossRef CAS.
- S. S. Stalin and E. K. V. Jino, Fabrication of Cu doped ZnO nanocrystals hybridised with graphene oxide nanosheets as an efficient solar light driven photocatalyst for the degradation of Quinalphos pesticide in aqueous medium, J. Water Environ. Nanotechnol., 2023, 8(2), 94–107 CAS.
- H. Adabavazeh, et al., Synthesis of polyaniline decorated with ZnO and CoMoO4 nanoparticles for enhanced photocatalytic degradation of imidacloprid pesticide under visible light, Polyhedron, 2021, 198, 115058 CrossRef CAS.
- Z. Zhu, et al., Photocatalytic degradation of an organophosphorus pesticide using a ZnO/rGO composite, RSC Adv., 2020, 10(20), 11929–11938 RSC.
- M. Naghizadeh, M. A. Taher and A.-M. Tamaddon, Facile synthesis and characterization of magnetic nanocomposite ZnO/CoFe2O4 hetero-structure for rapid photocatalytic degradation of imidacloprid, Heliyon, 2019, 5(11), e02870 CrossRef CAS PubMed.
|
| This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.