Next Article in Journal
Graphitic Carbon Nitride-Based S-Scheme Heterojunctions: Recent Advances in Photocatalytic Dye Degradation
Previous Article in Journal
Efficiency of Microwave-Assisted Surface Grafting of Ni and Zn Clusters on TiO2 as Cocatalysts for Solar Light Degradation of Cyanotoxins
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Applications of Quantum Dots in Photo-Based Advanced Oxidation Processes for the Degradation of Contaminants of Emerging Concern—A Review

Faculty of Chemistry, Warsaw University of Technology, Noakowski Str. 3, 00-661 Warsaw, Poland
*
Author to whom correspondence should be addressed.
Catalysts 2025, 15(6), 591; https://doi.org/10.3390/catal15060591
Submission received: 31 March 2025 / Revised: 16 May 2025 / Accepted: 27 May 2025 / Published: 14 June 2025

Abstract

Nanomaterials are interesting due to their unexpected and unique properties arising from phenomena occurring at the so-called mesoscale (that is, between single atoms and bulk solids). Among nanomaterials, one may distinguish quantum dots, which are highly crystalline nanocrystals with sizes up to c.a. 10 nm. Due to the quantum confinement effect, quantum dots exhibit extraordinary electronic and optical properties and may be utilized in photocatalysis. Semiconducting quantum dots may absorb photons, which results in the excitation of electrons from valence to conducting bands. Excited electrons in the conducting band and positive holes in the valence band may interact with chemical molecules (e.g., with water molecules), forming highly reactive radicals. Consequently, quantum dots may be utilized in advanced oxidation processes based on the action of light (i.e., photo-based advanced oxidation processes). Furthermore, quantum dots have advantages, such as having a tunable energy band gap and relative cost-effectiveness. Advanced oxidation processes are very important in the context of the constantly increasing pollution of the natural environment. Contaminants of emerging concern, such as pesticides, endocrine-disrupting compounds, and flame retardants, are still being detected in naturally present water. Such compounds may be degraded using advanced oxidation processes, utilizing quantum dots as photocatalysts. However, many operational parameters (such as quantum dots’ properties, including the means of their preparation) influence the efficiency of such processes; thus, detailed studies are being conducted.

1. Introduction

The last few decades have seen rapid developments in the field of nanotechnology. Nanomaterials produced by nanoscientists have been found to be useful for many wide-scale applications. This fact has contributed to the rapid development of nanotechnology. The current definition of a nanoparticle, provided by the International Union of Pure and Applied Chemistry, states that a nanoparticle is a particle of any shape but with a size lying in a range from 10−9 to 10−7 m [1]. Nanomaterials lie somewhere between the quantum and macroscopic scales; some of their properties are similar to those of molecules and, simultaneously, some of their other properties are rather similar to those of bulk solids. This fact contributes to the observed extraordinary behavior of nanomaterials. For example, their relatively large surface/volume ratio allows for increased adsorption of chemical molecules on their surface. This feature is especially demanded in catalysis [2]. Nanoparticles also exhibit a relatively high surface energy, which contributes to the ability of nanostructures to participate in various chemical reactions [2]. Quantum dots have especially found applications in the catalytic degradation of various organic compounds. For example, Cd0.5Zn0.5S and carbon dots (CDs) were combined in S-scheme heterojunction photocatalysts (Cd0.5Zn0.5S/Bi2MoO6/CDs) with improved stability and activity regarding the degradation of various antibiotics [3]. Another example of the application of quantum-dot-based photocatalytic degradation is the removal of levofloxacin (a third-generation quinolone antibiotic) by a multicomponent and enhanced fibrous photocatalyst consisting of carbon QDs, CdS, and Ta3N5; the combined heterostructure (CQDs/CdS/Ta3N5) exhibited a kinetic rate constant exceeding that of a CdS/Ta3N5 heterostructure by 2.1 fold and that of standalone Ta3N5 by 39.4 fold [4]. As may be seen from the above examples, the preparation of heterostructured photocatalysts consisting, at least partially, of quantum dots is a relatively new concept that may lead to increased photocatalytic activity [5,6,7].
The main advantage of the application of quantum dots in photocatalysis is their relatively high photocatalytic activity, which is related to their excellent optical and electrical properties [7,8]. On the other hand, the disadvantages of quantum dots include the following: they often require a problematic synthesis process; complicated operation, separation, and dosage; susceptibility to degradation (including chemical and photo-degradation); and they have relatively complicated and expensive characterization techniques. Moreover, one of the main advantages of quantum dots—the ability to generate reactive oxygen species (ROS)—is, at the same time, responsible for their unwanted and hazardous biological activity, which may include, for example, DNA damage [9,10]. Quantum dots may also release toxic elements into their surroundings, which brings additional concerns.
Contaminants of emerging concern (CECs) are a new group of different chemical compounds ranging from metals, polymers, and semiconductors to complex chemical compounds that are being constantly released into aquatic systems due to the immersive and rapid industrial development of various branches. CECs may be exemplified as pesticides, endocrine-disrupting compounds, and flame retardants. Such compounds may be degraded by advanced oxidation processes utilizing quantum dots as photocatalysts. However, many operational parameters (such as quantum dots’ properties, including the means of their preparation) influence the efficiency of such processes; thus, detailed studies are being conducted.
The aim of this review is to collect and discuss the most recent investigations on the synthesis and application of quantum dots in the photo-based removal of contaminants of emerging concern.

2. Quantum Dots—Preparation, Characterization, and Properties

Quantum dots are crystalline nanoparticles with sizes lying typically in a range from 1 to 10 nm and which exhibit semiconducting properties [2]. However, the most important criterion for a nanoparticle to be a quantum dot is for its size to be comparable to the exciton Bohr radius, which, in turn, depends on the kind of material that constitutes the nanoparticle (e.g., for CdSe, the exciton Bohr radius is c.a. 5 nm) [2]. Such a criterion ensures the exhibition of the quantum confinement effect, which leads to the quantization of energy levels in a quantum dot (similarly to the energy levels in a single atom) [2]. This means that even a metallic bulk solid may reveal semiconducting properties at the nanoscale [2].
There are many methods for the preparation of quantum dots, which may be, in general, divided into two subgroups: bottom-up and top-down [2,11]. The latter comprises mainly physical methods and includes mechanical methods (such as high-energy ball milling) and laser ablation, among others [11]. Bottom-up methods, on the other hand, depend on building the nanoparticle from the bottom and are mainly chemical in nature [2,11]. They include various colloidal syntheses (e.g., the precipitation of sulfides and reduction in noble metal cations), the sol–gel method, hydrothermal methods, microwave-assisted synthesis, sonochemical synthesis (i.e., ultrasound assisted), and electrochemical methods, among others [11,12,13].
Characterization techniques of quantum dots are numerous. Microscopic techniques include transmission electron microscopy (TEM), scanning electron microscopy (SEM), scanning transmission electron microscopy (STEM), high-resolution transmission electron microscopy (HRTEM), atomic force microscopy (AFM), and scanning tunneling microscopy (STM). They enable the investigation of the morphology, topology, size, and phase boundaries of samples [2]. Among spectroscopic techniques, one may list energy dispersive X-ray spectroscopy (EDX), Raman spectroscopy, X-ray photoelectron spectroscopy (XPS), extended X-ray absorption fine structure (EXAFS), Mössbauer spectroscopy, inductively coupled plasma atomic emission spectroscopy (ICP-MS), Fourier-transform infrared spectroscopy (FTIR), and ultraviolet-visible spectroscopy (UV-Vis), among others. These spectroscopic techniques provide information on the chemical and elemental composition of quantum dots, their size and phase composition, optical properties (optical band gap), and the local arrangement of atoms within them [2]. Diffraction techniques, such as X-ray powder diffraction (PXRD), give information on the phase composition, crystalline structure, lattice deformations and crystallite sizes [2]. Lastly, absorption techniques such as gas adsorption (e.g., BET measurements) allow for the estimation of the specific surface area of a nanomaterial—a parameter that is especially important in catalytic processes [2].
The semiconducting properties of quantum dots make them suitable for use in photocatalysis. Due to quantum confinement and an increase in the energy band gap, quantum dots absorb photons of specific energy, and that energy is higher than for a bulk material. After photon absorption, electrons from the valence band are promoted to the conduction band. The ‘space’ that remains after the excitation of the electron is called a positive hole. Electrons and holes may recombine and emit energy in the form of heat or light. However, if they do not recombine, they may take part in various chemical reactions that are responsible for their photocatalytic applications. First of all, positive holes may oxidize the surrounding molecules of a solvent (typically: water), forming corresponding radicals. In the case of oxidizing water molecules, hydroxyl radicals are formed which, in chemistry, are known as the second strongest oxidizing agent, after fluorine. These radicals may further react and oxidize organic compounds, causing their degradation. Positive holes may also directly oxidize organic molecules that are adsorbed at the surface of the quantum dot. On the other hand, photogenerated electrons act as chemical reductors and react with oxygen adsorbed on the surface of the quantum dot. Such a reaction produces peroxide radicals, which also have a significant oxidizing capability and may destroy organic molecules.
From the point of view of enhancing photocatalytic activity, the following properties of quantum dots are important:
-
size, which determines the optical band gap of the quantum dot (the higher the band gap, the higher the energy of the photoexcited electrons), surface (a larger surface means more likely absorption of organic molecules), the surface-to-volume ratio (smaller quantum dots have relatively more atoms on the surface that are reactive in contrast to the atoms present inside the quantum dot), and the path of photogenerated carriers to the reactive surface (smaller size means shorter path and thus smaller probability of electron and hole recombination),
-
optical properties, which determine which photons are absorbed and how likely they are to be absorbed (i.e., the absorption coefficient should be high),
-
electric properties, which influences the probability of electron and positive hole recombination.
The properties of quantum dots may be tuned to some extent by doping with various elements, forming heterostructures, and modifying the surface with passive layers or absorbed organic ligands. For example, in a heterostructure, an additional component may increase the specific surface area, improve the separation of photogenerated carriers, modify the optical energy band gap, enhance the absorption of photons, and improve the adsorption of target organic molecules by specific physicochemical interactions.

3. Quantum Dots for Photo-Based Removal of Pesticides

Pesticides are widely used in agriculture, forestry, and horticulture. However, they are hazardous to the environment and considered as contaminants of emerging concern. Photo-based degradation of pesticides is a promising sustainable way for pollutant removal; however, there are various challenges in both the design of materials as well as the selection of optimal operational parameters. Factors such as the pollutant and catalyst concentration, temperature, pH, and the presence of oxidizing agents significantly influence the removal efficiency. Quantum dots and their heterojunctions have great potential in generating reactive oxygen species due to their light absorption, charge carrier dynamics, and adsorption properties.

3.1. Influence of Operational Parameters on the Efficiency and Kinetics of Photo-Based Processes for Removal of Pesticides

3.1.1. Influence of Pollutant Concentration

Photocatalytic pesticide degradation is typically described by the Langmuir-Hinshelwood model; therefore, it follows pseudo-first order kinetics. However, experiments show that removal efficiency decreases as the pollutant concentration rises. This phenomenon is explained by the fact that with an increase in pollutant concentration, more pesticide molecules adsorb on the surface of the catalyst, blocking the active sites, so less reactive oxygen species (ROS) are generated, which causes a decrease in the degradation rate.

3.1.2. Influence of Catalyst Concentration

With the rise of catalyst amount, more reactive oxygen species are generated per unit of time, increasing the degradation rate. However, after some critical point, the addition of the catalyst results in only minor changes [14]; therefore, it is not optimal in terms of cost. Moreover, the excessive concentration of nanoparticles causes a reduction in light penetration and may lead to a decrease in the efficiency of pesticide removal [14]. The phenomenon is depicted in Figure 1.

3.1.3. Influence of Temperature

According to the Arrhenius equation, with an increase in temperature, the reaction rate of the elementary reaction increases. Additionally, temperature rise enhances the charge carrier dynamics, promoting the generation of ROS. On the other hand, an increase in temperature leads to enhanced electron hole recombination, reducing the efficiency of photocatalysis. In addition, the adsorption properties decrease after reaching a certain critical point. Experiments show that the removal rate of certain quantum dots (QDs) enhances with an increase in temperature [15,16,17], while others reach a maximum at a specific temperature and then decrease [18].

3.1.4. Influence of the pH of the Reaction Medium

The influence of pH on the photo-based removal of pesticides is complex. At first, it impacts the charge of the catalyst surface and, therefore, adsorption equilibrium. The second factor is the hydrolysis of pesticides under a certain pH. To optimize adsorption, the pH at the zero point of charge and pKa of pollutants should be considered. Targhan et al. [14] reported the hydrolysis of imidacloprid in alkaline (pH = 9) conditions. The optimal pH is different for each system: in the case of the degradation of carbaryl with CQDs modified with ZnO and TiO2, the optimal pH was found to be 7 [19], for CdS QDs on a molecularly imprinted polymer used in the degradation of buprofezin, it was 10 [17], while for the case of imidacloprid removal with the same catalyst, the optimal pH was found to be 9. In the case of ZnS/Ag2S/thiol groups containing graphene QDs ternary heterojunctions, used in diazinon and fenitrothion degradation, the optimal pH was 3 for both pesticides. It has been reported that the optimal pH for the heterogeneous photo-Fenton process is acidic, because Fe2+ and Fe3+ ions are present in the form of active aqua complexes instead of less reactive hydroxide complexes [20,21].

3.1.5. Presence of Oxidizing Agents

The presence of H2O2 causes enhanced generation of ROS. Targhan et al. [15] reported that in the case of imidacloprid removal by a binary composite of CQDs containing thiol groups and CdS QDs, the addition of H2O2 at a concentration of 20 ppm increased the removal efficiency from 75.60% to 88.06%. However, a further increase in H2O2 concentration led to only a minor rise in removal efficiency. This result is consistent with research by Pervez et al. [21], in which GQDs/α-FeOOH used QDs for the photo-Fenton degradation of ciprofloxacin (CIP), and where the influence of the H2O2 concentration was investigated in the range of 0.125–0.75 mM. The difference between 0.50 mM and 0.75 mM was not significant compared to the increase from 0.25 mM to 0.50 mM. Therefore, 0.50 mM concentration was chosen as optimal.

3.2. Examples of the Application of Quantum Dots in Photo-Based Degradation of Pesticides

Figure 2 presents the molecular structures of pesticides that were degraded using various quantum dots.

3.2.1. Photocatalytic Degradation

Targhan et al. [15] studied the photocatalytic degradation of imidacloprid by CdS QDs passivated with CQDs containing thiol groups for stabilization. Under simulated visible light, at pH 9 and after 90 min of irradiation, the removal efficiency was 90.13%. The influence of components was investigated—photocatalytic performance decreased in the seria CQDs-SH/CdS QDs > CQDs-SH = CdS QDs, i.e., the formation of heterojunctions increased the observed efficiency. Photocatalytic experiments using a UV-C lamp as the light source demonstrated that it is more effective than simulated visible light. The optimal pH was basic—experiments were conducted in a pH range of 5–9; the highest degradation rate was observed at pH 9. This phenomenon was caused by higher adsorption, as a photocatalyst surface has a negative charge (pH at zero point of charge 5.2), while imidacloprid at pH 9 is positively charged (pKa = 11.12). The influence of the photocatalyst concentration led to the conclusion that an increased catalyst dose provides the generation of more ROS per unit of time; however, an excessive concentration of catalyst causes reduced light penetration of the system, and the removal rate decreases. The effect of H2O2 addition was studied, and it was found that an increase in its concentration from 0 to 20 ppm led to a strong increase in photodegradation efficiency, i.e., from 75.60% to 88.06%. A further increase in the H2O2 concentration led to a minor enhancement in imidacloprid degradation. Those authors connected this observation with the fact that at relatively high concentrations, H2O2 can potentially act as an ·OH radical scavenger. With an increase in temperature from 25 to 40 °C, the degradation rate increased due to an enhancement in charge carrier mobility. Reusability tests showed a slight decrease in degradation over three cycles. The products of degradation were studied with the LC–MS technique (Figure 3); 6-chloronicotinaldehyde (Figure 3-(3)) was considered as the main product of degradation. Imidacloprid may be hydrolyzed to imidacloprid-urea (Figure 3-(2)), then degraded to 6-chloronicotinaldehyde, and oxidized to 6-chloronicotinic acid (Figure 3-(4)).
Targhan et al. [14] used terpyridine (Tpy)-coated CdS QD loaded on a metal–organic framework, i.e., ZIF-67 (Zeolitic imidazolate framework-67), nanocomposite for the photocatalytic degradation of imidacloprid. To investigate the partial influence of these components on the composite degradation efficiency of ZIF-67, CdS/Tpy QDs and ZIF-CdS/Tpy were compared. The removal efficiency was found to be 56.3%, 72.2%, and 91.0%, respectively. The effect of pH was investigated in the range of 5–9; it turned out that better degradation was obtained under alkaline conditions, with a removal efficiency at pH 9 of 80.9%, while at pH 5, it was only 59.2%. Under alkaline conditions, the hydrolysis of imidacloprid was also observed. Without a catalyst, the removal efficiency was 12.50%. Experiments with scavenger addition revealed that hydroxyl radicals are responsible for degradation. Removal rates under dark conditions, which characterize the adsorption equilibrium, increased as follows: CdS/Tpy QDs 15.4%, ZIF-67 43.6%, and ZIF-CdS/Tpy 47.4%. The BET surface area and the total pore volume of ZIF-CdS/Tpy were 1611.1 m2/g and 0.6805 cm3/g, respectively. These results suggest that the large specific surface area of ZIF-67 enhanced the adsorption of pollutants, while CdS/Tpy QDs ensured the generation of hydroxyl radicals and the degradation of imidacloprid.
Muangmora et al. [19] obtained carbon quantum dots using the hydrothermal method from spent coffee and modified them with ZnO and TiO2. CQD modification increased the BET surface area: the values were 10.64 m2/g, 11.44 m2/g, and 24.148 m2/g for ZnO, TiO2, and TiO2/ZnO/CQD respectively. The combination of oxides and CQDs was studied in terms of carbaryl photodegradation. The removal efficiency growth was as follows: TiO2 < ZnO < TiO2/ZnO < TiO2/CQD < ZnO/CQD < TiO2/ZnO/CQD. The latter showed an almost six times higher degradation rate than TiO2 (0.0570 vs. 0.0098 min−1). With the increase in the carbaryl concentration, the rate of degradation decreased. Experiments with scavenger addition revealed that hydroxyl radicals make a major contribution to the process, while ·O2− radicals play a minor role. The effect of pH was investigated in the range of 3–11, and the highest degradation rate was observed at pH 7. Reusability tests showed a decrease from 99.01% to 84.09% of removal efficiency after 5 cycles. Those authors connected this decrease with the loss of catalyst during the separating and residual adsorption of carbaryl molecules.
Malik et al. [17] modified CdS QDs with molecularly imprinted polymer (MIP) using gingerol as a template [19] to obtain CdS/MIP and for the comparison of synthesized CdS QDs modified with non-imprinted polymer CdS/NIP. They then used them for the photocatalytic removal of Imidacloprid and Buprofezin. The best removal was observed at pH 10 for Buprofezin and pH 9 for Imidacloprid. Under a temperature of 20 °C and an optimized pH, after 2 h removal, the efficiencies for Imidacloprid were 81% in the case of CdS/MIP and 76% for CdS/NIP. Under the same conditions, the degradation of Buprofezin reached 72% and 68%, respectively. With the increase in catalyst concentration, the degradation rate increased. The rise of temperature from 20 to 60 °C enhanced removal efficiency. The increase in pesticide concentration led to a lower degradation rate, as the pollutant adsorbed onto the catalyst surface and blocked reactive sites, preventing the formation of ROS.
Idrees et al. [18] synthesized carbon and phosphorus co-doped boron nitride quantum dots (CP-BNQDs) with a hydrothermal method and investigated the photocatalytic degradation of a fungicide, i.e., chlorothalonil, and nine other molecules. In the case of chlorothalonil, the observed kinetics were of the pseudo-first order type, with a rate parameter of 0.0053 min−1. An additional investigation conducted on the degradation of the antibiotic tetracycline showed that the optimal temperature was 35 °C. Photocatalytic experiments with the addition of scavengers revealed that the active oxygen species causing degradation were both ·OH and ·O2–.
Nekooei et al. [16] obtained ZnS QD heterojunctions with graphene quantum dots containing thiol groups (S-G GQDs) and modified with Ag2S. The obtained ternary ZnS/SG QDs/Ag2S was utilized in the photocatalytic degradation of diazinon and fenitrothion. According to UV-VIS and COD (chemical oxygen demand) analyses, photocatalytic degradation consisted of two main steps: 1. the degradation of the initial molecule, and 2. the degradation of intermediates. UV-VIS absorbance decreased significantly in the first 35–40 min of the process, while the COD remained almost constant; thus, a notably longer time is needed for the decomposition of intermediates. The influence of temperature was well described by the Arrhenius equation. The activation energies of photocatalytic degradation of diazinon and fenitrothion were 4.8 and 5.1 kJ/mol, respectively. The pH effect was investigated in the wide range of 1–11, and the best efficiency was at pH 3 for both pesticides. The influence of individual components of ternary ZnS/SG QDs/Ag2S and their combination was also investigated. The degradation rate presented the same pattern for both investigated pesticides and increased as follows: ZnS QDs < ZnS/SG QDs < ZnS QDs/Ag2S < ZnS/SG QDs/Ag2S < Ag2S. Although Ag2S by itself showed better performance, the price of silver is high, so ZnS/SG QDs/Ag2S is more sustainable for commercial reasons. ICP analysis showed that the proportion of silver in the final catalyst was only 0.9%. The reusability test showed no significant changes over four cycles compared to the first cycle.
John et al. [22] modified Bi2O3 nanoparticles with 5 wt. % of nitrogen and sulfur co-doped carbon quantum dots (N,S-CQDs) and the studied photocatalytic degradation of 2,4-dichloropenoxyacetic acid (2,4-D) and diuron. After 2 h of sunlight radiation, the TOC (total organic carbon) removal rate of 2,4-D was 92% and 97% for diuron. High-resolution mass spectrometry showed that the main product of 2,4-D degradation was the less toxic 2,4-dichloroanisole, while the degradation of diuron led to the creation of 3,4-dichloroaniline.
Abbasi et al. [23] synthesized InP QDs and InP QDs covered with a ZnS QD core–shell structure by applying a solvothermal method. The degradation of dyes, hydrocarbons, and the pesticide deltamethrin was studied under a 300 W Xe lamp. Interestingly, InP/ZnS QDs showed better performance in the degradation of dyes and hydrocarbons while, in the case of deltamethrin, the result was 90.29% after 90 min versus 93.36% in the case of InP QDs. Reusability tests conducted on dyes proved good stability of the catalyst over three cycles.

3.2.2. Photo-Fenton Degradation

Swedha et al. [20] used a Ni2CuS4 Qds @ Fe3O4 nanocomposite with different amounts of Fe3O4, from 10% to 30%, in the photo-Fenton process for the degradation of the pesticide bromoxynil (BRX) and the antibiotic cefixime (CEF). Tests were conducted under a 1000 W halogen lamp with 20 mg/L concentration of catalyst and pollutant and 100 µL of H2O2. Photocatalytic tests were conducted on the following systems: only H2O2 (without catalyst), Ni2CuS4 QDs, Fe3O4, Ni2CuS4 QDs @ 10%Fe3O4, Ni2CuS4 QDs @ 20%Fe3O4, and Ni2CuS4 QDs @ 30%Fe3O4. The highest degradation rate was in the case of Ni2CuS4 QDs @ 30% Fe3O4 for both BRX (0.0070 min−1) and CEF (0.0038 min−1). The lowest rate of degradation was in the case of the reaction with only H2O2, which proved that the photo-Fenton process was dominant. The effect of the pollutant and catalyst dosage showed that a higher catalyst dosage at the beginning caused higher degradation, but further increase led to a reaction between Fe3+ and ·OH radicals. An increase in pollutant concentration led to a higher percentage of degradation, but further increase caused decline due to the limitation of the generated reactive oxygen states. The influence of pH was investigated in a range of 4–9, and the highest degradation rate was obtained with acidic pH. The presence of anions that are common in real wastewater samples, e.g., Cl, CO32−, NO3, PO43−, and SO42−, caused a minor decrease in degradation. Reusability tests proved high stability and reusability over six cycles. Experiments with scavengers revealed that ·OH and ·O2− radicals were responsible for the degradation, while the addition of hole and electron scavengers did not significantly influence the process.
Pervez et al. [21] obtained nanocomposites from graphene quantum dots with α-FeOOH and utilized them in the photo-Fenton degradation of ciprofloxacin (CIP). Figure 4 illustrates the influence of operational parameters on removal efficiency. To investigate the impact of the GQD amount in the composite, different portions of GQDs, from 9 to 15 mg, were added during synthesis to 14.456 g of FeSO4 · 7H2O. The degradation efficiency was more than 90% in each case, and the highest value was in the case of 13.5 mg GQDs. For comparison, pure α-FeOOH resulted in only 52.6% photo-Fenton degradation. After 60 min, 93.73% of the pesticide was degraded; for comparison, during the Fenton process in dark conditions, the removal efficiency was only 52.50%, proving the significant effect of light in this process. With an increase in H2O2 concentration, i.e., in the range of 0.125–0.75 mM, the degradation rate increased; however, the difference between 0.50 mM and 0.75 mM was not significant compared to the increase from 0.25 mM to 0.50 mM. Therefore, 0.50 mM concentration was chosen as optimal. The influence of pH was investigated in the range of 3.01–10.40. CIP has two dissociation constants (pKa1 = 6.05 and pKa2 = 8.37), and the point of zero charge of GQDs/α-FeOOH is between 8.5 and 9.0. Generally, higher rates were obtained in acidic conditions: at pH 3.01 and 4.95, removal was the same (93.7%), while increases of pH to 7.00, 9.04, and 10.40 resulted in a decrease in the degradation rate to 79.78%, 58.7%, and 22.0% respectively. A series of experiments with the addition of radical inhibitors revealed that ·OH, ·O2−, and 1O2 play important roles in the process.
The applications of quantum dots in photo-based removal processes of pesticides are summarized in Table 1.

4. Quantum Dots for Photo-Based Removal of Flame Retardants

Flame retardants are substances that reduce or lower the overall fire risk of materials. These substances are widely used in composites, textiles, clothes, plastics, building supplies, bedding, furniture, etc. [24]. Flame retardants may be categorized as, for example, inorganic, halogenides (based on bromine and chlorine), and organo-phosphorus.
Brominated flame retardants contaminate the environment, mainly water, dust, air, and soil, from which exposure to humans occurs, potentially affecting human health. Brominated flame retardants may act as endocrine disruptors—they can interfere with hormonal function [25].
Brominated aromatic compounds (including brominated flame retardants) demonstrate the ability to absorb the electromagnetic waves from the UV region. A sufficient amount of captured electromagnetic energy causes the facile homolytic breaking of the C–Br bond and provokes further structural rearrangements, including the breaking of the ether bond, which leads to the photodegradation process [26].
Quantum dots can boost the process of photodecomposition—they generate an electric field that enhances the catalytic activity on the conduction band on the conductor’s surface [27].

4.1. Examples of the Application of Quantum Dots in the Photo-Based Degradation of Flame Retardants

4.1.1. Photodegradation of 4-Bromophenol by Graphitic Carbon Nitride with Graphene Quantum Dots

Halophenols, such as 4-bromophenol are obstinate organic substances, which are used in the production of many substances, such as brominated flame retardants (BFRs), polymer intermediates, drilling fluid, pharmaceuticals, and wood preservatives.
Brominated phenols (BPs) exhibit toxicity. The degradation processes of these substances are slow, and they contaminate water and soil in the long term. The environmental pollution connected with these flame retardants is hard to overcome via natural processes. Notably, 4-BP is considered a priority toxic halogenated phenol by the USEPA (the United States Environmental Protection Agency) based on its mutagenic, genotoxic, and carcinogenic nature [25,27]. Diverse decomposition procedures (chemical, physical, and biological treatment procedures) have been utilized for the treatment of halophenols in wastewater.
Graphitic carbon nitride functionalized with graphene quantum dots (CN@GQN; Figure 5) exhibits an effective photodegradation of 4-bromophenol. CN@GQN is obtained by the in-plain induction of GQDs into exfoliated graphitic carbon nitride via visible light-driven photodegradation.
CN-Ex (carbon nitride) can convert electrons from the valence band, because it has exceptional conductivity. GQDs are responsible for generating an electric field to improve the catalytic activity on the conduction band. A metal-free CN@GQD composite showed a relatively high photodegradation rate of 4-BP under visible light, which could be debrominated and oxidized separately on the conduction and valence bands, respectively [27].
Research has shown that the inclusion of GQD and CN layers might be responsible for the modification of the band structure and the enhancement of hydrophilicity and photocatalytic effectiveness [25].
Graphene-based and various transition-metals catalysts have been considered as photocatalysts for the degradation of bromophenols due to their low cost, strong resistance to chemical reactions, and enhanced charge carrier separation.
CN@GQN, as a non-metallic composite, is more advantageous than metallic catalysts with regard to metal leaching, incomplete pollutant degradation, and the generation of toxic sludge, which are present in the case of metallic photocatalysts [27].
Visible light represents about 43% of the total solar spectrum, while UV light comprises only 5% (Figure 6). For this reason, decomposition under visible light could be more beneficial in practical terms. During a chemical analysis, a composite of GQD was made via the hydrothermal process, built by the induction into the layers and edges of the CN. The application of GQD@CN for the mineralization of 4-BP under LED visible light (λ = 420 nm) irradiation at room temperature led the researchers to the conclusions that the synergistic effect of GQD and CN had caused the complete mineralization of 4-BP, and CN@GQD markedly facilitated the red shift of the band gap and extensive solar light absorption [28].
The recycling ability of the catalyst demonstrated great potential as a metal-free catalyst for the treatment of micropollutants in the aquatic environment. It also indicated that the above-mentioned method would be effective and beneficial for the degradation of 4-bromophenol [25].

4.1.2. Double Heterojunction Carbon Quantum Dots (CQDs)/CeO2/BaFe12O19 for the Degradation of Tetrabromobisphenol A (TBBPA)

Tetrabromobisphenol A (TBBPA) is often applied as a flame retardant in insulated wires, printed circuit boards, and polycarbonate plastics. TBBPA has been detected in soil, sediment samples, and living organisms [29,30]. As a brominated flame retardant, TBBPA may seriously affect metabolism, cause hormone secretion disorders, and even damage the endocrine system, bones, and brain [30]. Efficient techniques for the elimination of TBBPA from the environment are crucial due to the reasons outlined above [29,30].
Double heterojunction carbon quantum dot/CeO2/BaFe12O19 magnetic separation photocatalysts (CCBFOMSPs) were prepared via a polyacrylamide gel method and low-temperature sintering combined with a hydrothermal method. The photocatalytic activity of this catalyst for the degradation of TBBPA was improved by regulation of the electromagnetic interaction between the BaFe12O19 contents and the photo-generated electrons.
The construction of a special heterojunction turned out to be an effective method to enhance the photocatalytic activity of the BaFe12O19 photocatalyst. Cerium dioxide (CeO2) is a wide-gap semiconductor of n-type conductivity, which may be used to construct, together with BaFe12O19, a composite photocatalyst with n-n type heterojunctions. Carbon quantum dots (CQDs) may be additionally incorporated to serve as charge transport carriers, improving the transfer and separation of charge carriers in n-n type BaFe12O19/CeO2 photocatalysts. In this method, CQDs could be made using glucose as a raw material by the hydrothermal method.
Figure 7a shows the structure of Tetrabromobisphenol A. Figure 7b shows the time-dependent photocatalytic degradation of TBBPA in the presence of Samples S1, S2, S3, and S4 under simulated sunlight radiation (S1—ultrafine CeO2 nanoparticles, S2—15% CQDs and 5 wt. % BFO/CeO2, S3—15% CQDs and 10 wt. % BFO/CeO2, 15% CQDs and 10 wt. % BFO/CeO2). Figure 7c,d shows plots of ln(Ct/C0) vs. irradiation time and the first-order kinetic constant (k) of TBBPA in the presence of samples S1, S2, S3, and S4 under simulated sunlight radiation, respectively. The degradation rate of sample S3 was 3.29 times that of sample S1. The results showed that the constructed heterojunction could effectively degrade the Tetrabromobisphenol A [30].
The photocatalytic activity of the used photocatalysts could be linked to the oxidation and reduction capacity of positive holes (hVB+), hydroxyl radicals (•OH), and superoxide radicals (•O2). It was observed that the saturation magnetization of CCBFOMSPs increased with an increasing BaFe12O19 content. However, the photocatalytic activity of CCBFOMSPs first increased and then decreased with an increase of BaFe12O19 content. This trend was consistent with the trend of the visible light absorption coefficient among CCBFOMSP samples. The results showed an internal correlation between magnetic and optical properties and photocatalytic activities [30]. Experiments demonstrated that CCBFOMSP is an effective photocatalyst in the photodecomposition of Tetrabromobisphenol A.

4.1.3. TiO2 Nanomaterials with Various Types of QD to Remove Tetrabromobisphenol-A

A F-TiO2/g-C3N4 heterojunction was successfully applied to degrade widely used flame retardants. Nanostructured TiO2, due to its unique photochemical and electrochemical properties, has been widely applied in photocatalysis. TiO2 can soak up only a few solar photons and scarcely responds to visible light because of its large bandgap of 3.27 eV. There are special mechanisms that are used to synthesize high-performance TiO2-based materials to improve the absorption competence for solar light and the potential for charge separation and transportation. Those special mechanisms involve doping, controlling heterojunctions, and compositing with quantum dots (C, Ge, ZnS, CdS) [31].
A synthetic procedure for the preparation of a F-TiO2/g-C3N4 heterojunction photocatalyst with a three-dimensional structure included one-step calcination, which improved the photocatalytic activity of the F-TiO2/g-C3N4 photocatalysts. Remarkably, compared with pure g-C3N4, F-TiO2, and TiO2/g-C3N4, the photodegradation performance (under visible light) of F-TiO2/g-C3N4 toward TBBPA was greatly enhanced [31]. This was due to the enhancement of the separation efficiency of photogenerated electron-hole pairs [31].

4.1.4. Detection of Melamine by Digital Image Colorimetry Using Carbon Nitride Quantum Dots in a Cellulose Matrix with a Smartphone-Based Portable Device

Melamine derivatives are salts of melamine (an organic base) with organic or inorganic acids such as cyanuric, boric, and phosphoric or polyphosphoric acids. For example, melamine cyanurate (MC) is used in the production of halogen-free flame-retarded unfilled and mineral-filled Pas (Polyamides) for electronic and electrical applications [32].
To integrate carbon nitride quantum dots into a cellulose matrix, carbon nitride quantum dots (CNQDs) were embedded in a sodium carboxymethyl cellulose (CMC) matrix to form a CNQD-CMC film, which was then used to explore the room temperature phosphorescence (RTP) of CNQDs. Due to the internal hydrogen bonding interactions between CMC and CNQDs, the non-radiative relaxation process was suppressed. During the experiments, a simple, background-free miniature device integrated with a CNQD-CMC film was made. In the presence of MEL, the yellow RTP color of the CNQD-CMC film was photographed using a smartphone. The Color Recognizer APP was used to recognize the red (R) value for the quantitative detection of MEL. As a result, a digital image colorimetry (DIC) determination of MEL was achieved due to the visible RTP color change of the CNQD-CMC film [33].

5. Quantum Dots for Photo-Based Removal of Dyes

Industrial effluents containing dyes make water unfit for drinking. Water pollution is nowadays a concern to the global community, since wastewater from the leather, textile, food, and chemical industries may contain hazardous dyes which can cause serious environmental issues. Various organic dyes are commonly used as coloring agents in the cosmetics, leather, textile, printing, paper, plastic, pharmaceutical, and rubber industries. Heterogeneous photocatalysis using QDs is one of many technologies for the advanced degradation of water contaminated with organic dyes.

5.1. Examples of the Application of Quantum Dots for the Photo-Based Degradation of Dyes

5.1.1. Carbon Quantum Dots with TiO2 Nanocomposite for the Photodegradation of Rhodamine B

Rhodamine B (RhB) (Figure 8) is a toxic, non-biodegradable dye. It exhibits carcinogenic and neurotoxic effects, as well as the ability to cause several diseases in humans [34,35]. Rhodamine B is a synthetic dye utilized in the manufacturing of dye lasers, fireworks, stamp pad inks, ballpoint pens, and carbon sheets [35].
Pure TiO2 and a carbon quantum dot (CQD)-doped TiO2 nanocomposite (CQDs/TiO2 nanocomposite) were prepared by a sol–gel method for the photocatalytic removal of toxic Rhodamine B [36].
TiO2 is a photocatalyst that needs another compound for the modification of its nanoparticle structure to enhance its photocatalytic activity. This might be achieved by coupling with materials such as semiconductors, metals, dyes, and nonmetal elements. These additives may form the energies of the intermediate states in the band gap. Carbon nanomaterials have been demonstrated to efficiently enhance photocatalytic reactions due to their superior conductivity, capacitance, absorption of visible light, and chemical stability. For example, carbon quantum dots (CQDs), as a kind of novel and environmental carbon nanomaterial, have drawn much attention due to their multiple synthesis methods, favorable biocompatibility, and unique photo-induced electron transfer. It was demonstrated that the absorption wavelength of the composite material underwent a red shift to the visible region in comparison to pure TiO2. When the content of CQDs in the composite was 5% (based on the mass of TiO2), the obtained CQDs/TiO2 exhibited excellent catalytic activity for the degradation of RhB. The degradation degree of CQDs/TiO2 was 85.47%, which was 5.5 times higher than that of standalone TiO2. The degree of catalytic degradation of RhB remained above 80% even after four cycles. Investigations of the photocatalytic mechanism of RhB by CQDs/TiO2 revealed that positive holes (h+), superoxide anion radicals (•O2), hydrogen peroxide (H2O2), and hydroxyl radicals (•OH) were the main active species in this process [37].
To obtain a composite photocatalyst, carbon quantum dots (CQDs) were embedded onto the external surface of the TiO2 nanotubes (TNTs). CQDs were produced via the hydrothermal method. The results indicated that the CQDs played an important role in the visible-light photocatalytic process [37].
The TiO2-doped CQD heterostructure possessed higher photocatalytic activity under UV-light irradiation compared to TiO2. The photocatalytic degradation of Rhodamine B was improved when TiO2-doped CQDs were used instead of TiO2 [37].

5.1.2. Carbon Quantum Dots for the Photodegradation of Methylene Blue

Methylene blue (methylthioninium chloride) (MB) is a cationic organic dye from the phenothiazine family. It is a tricyclic phenothiazine compound that is soluble in water and some organic solvents. Over the years, it has been used for the treatment of malaria, methemoglobinemia, and other pathologies [38].
Carbon Quantum Dots (CQDs), as a typical functional carbonaceous material, have some excellent characteristics, such as luminescence, low toxicity, good up-converting, and relatively low cost of synthesis. The as-prepared samples of CQDs and TNTs exhibited significant photocatalytic activity in the case of the high initial concentration of the methylene blue solution (30 mg/L) within 50 min under visible-light irradiation. This excellent photocatalytic activity was mainly due to the up-conversion photoluminescence property of the CQDs, which can convert longer wavelength light (>600 nm) absorbed from visible light into light of shorter wavelengths (<600 nm) and activate the titania nanotubes to produce electron-hole pairs. Additionally, the stacking interaction of the π–π bonds between the CQDs and the benzene rings in molecules of dyes would improve the absorption ability of the CQD/TNT photocatalysts.
Carbon quantum dots (CQDs) were embedded onto the external surface of TiO2 nanotubes (TNTs) through an improved hydrothermal method. The results indicated that the CQDs played an important role in the visible-light photocatalytic process—the electrons photogenerated from the TNTs could be trapped by the CQDs and retard the recombination of photoexcited [39].

5.1.3. TiO2 and GQD Compounds for the Photodegradation of Methylene Blue

TiO2/GQD compounds absorb electromagnetic waves from the visible spectrum of sunlight and thus may be utilized for photocatalytic organic matter degradation. GQDs show relatively good physicochemical properties, such as stable photoluminescence, thermal conductivity, and improved surface grafting [40]. GQDs could be additionally modified with polyethylenimine (PEI) and polyethylene glycol (PEG).
Polymer-modified GQDs showed enhanced photocatalytic organic matter decomposition [40]. The degradation ability of these modified photocatalysts was demonstrated on the basis of the decomposition of methylene blue. The photocatalytic rates of degradation decreased in the following order: GQDs, GQDs-PEIs, GQDs-PEGs. Thus, polymer-modified GQDs could qualitatively control the degradation rate of MB [40].
GQDs were synthesized via a hydrothermal method using citric acid as a precursor. The highly fluorescent GQDs were highly soluble and stable in water [40].

5.1.4. Phosphorus-Doped Graphene Quantum Dots Loaded on TiO2 for Enhanced Photodegradation of the Methyl Orange

Methyl orange (MO) dye is carcinogenic, mutagenic, and toxic to aquatic organisms. Methyl orange (MO) is an anionic azo dye, having high chemical stability; it belongs to the sulfonated azo group (Figure 9) [41].
Photocatalytic efficiency may be enhanced by reducing the recombination of photogenerated electron-hole pairs. Transferring the photogenerated carriers to adsorbates is a typical way to diminish the direct recombination probability. It is possible to reduce this probability in TiO2 by forming a p-n heterojunction with p-type phosphorus-doped graphene quantum dots (P-GQDs). In such a heterojunction, the photogenerated holes are transferred from TiO2 to P-GQDs, which reduces the direct recombination of carriers. This results in the improvement of the Methyl Orange (MO) photodegradation efficiency to 95.5% after 14 min [42].
Graphene quantum dots (GQDs) are 0D carbon nanomaterials with a size of less than 10 nm. They have the following advantages over traditional semiconductor quantum dots: simple synthesis, unique boundary effects, and biocompatibility [42].
A P-GQD/TiO2 photocatalyst was synthesized via facile hydrothermal method. It exhibited efficient photocatalytic activity in the MO degradation process under UV light irradiation. In the hybrid photocatalyst, photogenerated carriers were transferred by the electric fields generated by light excitation and formed by the p-n junction, which, in general, led to the high effectivity of the photodegradation of MO [42].

5.1.5. Methyl Orange Piezodegradation via GQD/ZnO S-Scheme—A Different Method of Degradation with QD Usage

Alternatively, there are other methods for the degradation using QDs which do not depend on the usage of light. Piezoelectric materials are a class of materials with non-central symmetries. Under the action of mechanical forces, they exhibit the captivating phenomenon of piezoelectric polarization, which causes the reorganization of surface charges and starts subsequent redox reactions. Consequently, part of the charge screened by the polarized charge is released and may participate in the catalytic reaction [43].
Piezocatalysis is a newly emerging catalysis technology that relies on the piezopotential and piezoelectric properties of the catalysts [44]. The piezocatalytic technique offers a unique advantage, as it can convert mechanical energy into chemical energy without the need for light or electricity [45].
The GQD/ZnO heterojunction exhibited relatively good piezocatalytic activity, such as a relatively high reaction kinetic rate constant (0.051 min−1) and degradation efficiency (96.1% after 60 min) [45].
A GQD/ZnO S-scheme heterojunction has been applied as a piezocatalyst [45]. The S-scheme charge transfer pathway and piezocatalytic mechanism of GQD/ZnO were revealed, based on both experimental and theoretical analyses [45]. The process involves facilitating charge transfer and improving the redox ability [45].

6. Quantum Dots for Photo-Based Removal of Endocrine Disrupting Compounds

Endocrine disrupting compounds (EDCs) are chemical compounds that can interfere with the hormonal system and may produce harmful effects in both humans and wildlife.
Quantum dots have been applied in the photo-based removal of the following EDCs (see Table 2): Ethylparaben, Bisphenol A, Sulfamethoxazole, Norfloxacin, Diclofenac, Dimethyl Phthalate, Para-nitrophenol, and Sulfadiazine.

6.1. Examples of the Application of Quantum Dots for the Photo-Based Degradation of Endocrine Disrupting Compounds

6.1.1. Degradation of Sulfamethoxazole (SMX) Using a Hybrid CuOx–BiVO4/SPS/Solar System

BiVO4 catalysts (0.75, 3.0, and 10 wt. %) were synthesized using polyol-reduction with ethylene glycol as the reductant [46,47]. As may be seen in Figure 10, all obtained catalysts are able to remove all SMX with only the help of solar power and sodium persulfate activation (SPS). If only either SPS or solar was used with a catalyst, the efficiency could not exceed 60%, but with both, the contaminant was removed fairly quickly [48]. Among all of the tested CuxBVO4 catalysts, the best performing one was Cu3.0BVO4, which was further proven by another test where all catalysts were used to remove SMX. Cu3.0BVO4 was able to remove the entirety of the contaminant in 25 min, while the others did it in 30 min. Next, the effect of reactive species scavengers on SMX degradation was examined within the same environment as in Figure 10. The results showed that ultrapure water and 10 mM of t-BuOH did not make affect SMX degradation, but 10 mM of MeOH caused degradation of around 90% after 60 min, and a sample containing 10 mM of EDTA could remove more than 30% of SMX after 60 min. Lastly, SPS consumption was studied. For normal SPS/solar, there was a small change in SPS concentration after 60 min, while for Cu3.0BVO4/SPS, 10% was gone after 60 min, and for Cu3.0BVO4/SPS/Solar around 40% was gone by 60 min. This occurred under solar light irradiation, whereby pairs of electron-holes were formed in copper-promoted BiVO4 and SPS reacted with electrons in the semiconductor’s band forming SO4o− radicals that could participate in SMX degradation.

6.1.2. Hybrid Quantum Dots of Cadmium-Doped in Chitosan Forming Cd/CdIn2S4 for the Photodegradation of Ofloxacin (OFL) and Para-Nitrophenol (PNP)

Mishra et al. synthesized inorganic-organic hybrid quantum dots of cadmium-doped in chitosan, forming Cd/CdIn2S4 for the photodegradation of ofloxacin (OFL) and para-nitrophenol (PNP) [49].
The main reason why this catalyst was made was because CdIn2S4 has shown an ideal gap for a catalyst under visible light radiation, i.e., 2.18 eV. The catalyst was produced via a hydrothermal technique for electrocatalytic application, and Cd/CdIn2S4@Ch quantum dots were synthesized via the solvothermal method [50], but at a slightly higher temperature of 160 °C. Subsequently, the following research was done.
The first experiment was to determine the optimal amount of H2O2 for the best catalyst performance. This was done because H2O2 plays a critical part in the photo-Fenton -like oxidation process which was used here [49]. After testing in a range of 0–0.5 mL with a space of 0.1 mL between, the optimal amount of H2O2 was shown to be 0.2/0.3 mL for OFL/PNP (conditions: OFL/PNP—20/15 ppm, catalyst dosage 8 mg, pH = 7). The pH of the solution was greatly influenced by the surface charge of the synthesized Cd/CdIn2S4@Ch quantum dots; thus, to maximize degradation, the optimum pH must be calculated, including with the types of pollutants present in wastewater. The calculated values of optimal pH for OFL and PNP were 7 and 3, respectively, and these were used in later tests. The next experiment was to determine the influence of contaminant concentration on the degradation performance of our CDs; the best results were yielded for 20 ppm of OFL, but 15 and 10 ppm were very similar. Meanwhile, the best results for PNP were with 15 ppm. The decrease in catalyst performance with higher contaminant concentration may have been caused by the narrowing of photon path lengths. Also, higher amounts of contaminant would require a heavier catalyst load, which could enhance solution’s opacity and, in the end, inhibit photodegradation.

6.1.3. Carbon Dots Modified with g-C3N4 for Photo Oxidation of Bisphenol-A (BPA)

Iqbal et al. obtained carbon dots modified with g-C3N4 for the photo-oxidation of Bisphenol-A (BPA) under visible light irradiation [51]. The first obtained was g-C3N4 by using simple thermal condensation of urea. In this process, 10 g of urea was placed in a crucible and dried at 80 °C overnight, followed by heat treatment at 550 °C with a heating rate of 5 °C/min for 3 h. After heating, the obtained light-yellow powder was ground into a fine powder. Then, carbon quantum dots (CDs) were synthesized with a one-step microwave-assisted approach. The precursor for this synthesis was glucose, of which 70 mg was dissolved in 20 mL of ultrapure water in a beaker. After that, the beaker was placed into a microwave oven and heated at 500 W for 30 min. The obtained light to dark brown solid with some clusters was further dissolved with 10 mL of ultrapure water to produce a CD solution, which was then kept in the fridge. With both C3N4 and CDs having been obtained, doped CDs were synthesized by mixing 10 g of urea with 40 mL of ultrapure water, followed by the addition of 0.5, 1.0, and 1.5 mL CD solution for different doped catalysts. Next, the solution was mixed under a magnetic stirrer and heated at 80 °C until it was completely dry. Next, the mixture was cooled to form a solid, ground to a powder, and calcined at 550 °C for 2 h at a heating rate of 5 °C/min. With all catalysts obtained, the degradation of Bisphenol A was researched and the following results were obtained.
High-Resolution TEM microstructure was analyzed, and the size of CDs ranged from 2.2 nm to 3.9 nm, with an average of 3.75 nm. Besides that, the temperature used in the synthesis was not high enough, so the CDs may still have been multi-crystalline or amorphous and still containing significant amount of sp2 clusters.
The band gap was measured; higher doping showed a decreased band gap: 1.5 CDs/g-C3N4 (2.11 eV), 1.0 CDs/g-C3N4 (2.14 eV), 0.5 CDs/g-C3N4 (2.26 eV), g-C3N4 (2.46 eV).
As shown in Figure 11, a lower pH strongly favored the degradation of BPA on this catalyst [51]. Higher pH caused BPA to exist as an anion or even a dianion, which caused BPA to be repelled from the catalyst, which was also negatively charged, thus lowering its oxidation. This was why the adsorption of BPA on the surface of 1.5 CDs/gC3N4 was also high, at pH 3–7. Thus, pH 10 was chosen as the best for tests, because of a good balance between the BPA adsorbed on the surface of the catalyst and the amount of catalyst that generated e/h+ pairs.
The next test was the investigation of the influence of the initial concentration of BPA on BPA removal. The test was done with 5, 10, 20, and 30 mg/L (catalyst dosage 30 mg/L, solution volume 100 mL, and pH = 10), with 5 and 10 mg/L being optimal. In the case of higher dosages, the solution was probably less transparent, causing less light to be transported to the surface of the catalyst and making it work with less efficiency. Lastly, research were done on the effect of catalyst dosage and the effect of CD loading on BPA removal. All tests showed similar results, but the best performance was observed with 30 mg/L of 1.5 CD/g-C3N4 catalysts.

6.1.4. Photocatalytic Degradation of Fluoroquinolone and Norfloxacin (NOFX) by Mn:ZnS Quantum Dots

The synthesis of Mn:ZnS QDs was as follows [52]. First, solutions of sodium sulfide, zinc acetate, and manganese carbonate in water were prepared. Precursors for the obtaining process were 0.5 M solutions of Zn(CH3COO)2*H2O and Na2S. Then, 29.75 mL of 0.01 M MnCO3 solution was added to 49.50 mL of Zn(CH3COO)2*H2O. Subsequently, cetylpyridinium chloride (1 At. wt. %) was added as the capping agent. Next, the Na2S solution was added dropwise. Soon, a white precipitate appeared, and stirring continued for 15 min. The reaction was then refluxed at 60 °C and centrifuged at 5000 rpm for 5 min. Lastly, the precipitate was filtered through Whattman filter paper and washed to eliminate impurities [52].
After the reaction, an XRD diffractogram showed that Mn:ZnS QDs comprised 83.92 wt. % Zn, 13.32 wt. % S and 2.77 wt. % Mn. The band gap was also measured Via Tauc’s relation; for pure ZnS, it was 3.88 eV, and for Mn-doped ZnS QDs, it was found to be 4.12, 4.39, 4.6, and 4.5 eV for 0.5%, 1%, 3%, and 5% Mn doping, respectively [52].
The first test done was the investigation of the efficiency of ZnS QDs for NOFX removal (15 mg/L). The best results were obtained with Mn:ZnS+UV, achieving degradation above 80%, followed by pure ZnS+UV, with removal of around 80% of NOFX. The next test conducted was the investigation of the effect of the initial pH of the NOFX solution on its removal (Figure 12). pH plays a significant role in photocatalytic aqueous oxidation because it influences the charge of the catalyst, the size of catalyst particles, and the positions of valence and conduction bands.
As may be seen from Figure 12, the optimal pH for the degradation of those drugs was found to be 10. The decrease in drug degradation at lower and higher pH was due to surface changes in the catalyst. The ZnS point of zero charge is between 7 and 7.5, so at lower pH values, too much of the contaminant was absorbed, whereas at higher pH, there was not much NOFX adsorbed on the surface of the catalyst.
Next, the amount of catalyst for this pH was researched. As shown in Figure 13, the optimal amount of catalyst was 60 mg for 25 mL of the drug with 60 min of irradiation. Higher amounts of the catalyst also worked well, achieving results of around 75% degraded NOFX. This was because higher amounts of the catalyst caused reduced light penetration, while the initial increase in degradation for lower values was possibly due to the augmentation in the active site accessibility with increased loading of the catalyst.
Lastly, the effect of the initial concentration of the drug was studied; the optimal amount was found to be 15 ppm, with a degradation efficiency of around 86%, while for 10 and 20 ppm, degradation lowered to 60 and around 45%, respectively. The decline in efficiency with rising drug concentration was because a higher proportion of drug molecules were adsorbed on the surface of QDs, leading to a decreased numbers of active sites, thus creating a lesser amount of OH radicals. Also, because of the presence of more molecules, fewer photons could reach the catalyst surface.

6.1.5. Photocatalytic Degradation of Diclofenac (DCF) by Nitrogen-Doped Carbon Quantum Dot-Graphitic Carbon Nitride (CNQD)

The prepared compounds were bulk g-C3N4 and g-C3N4 nanosheets [53]. The bulk was obtained via the thermal polymerization of urea. Firstly, 10 g of urea was placed in a 50 mL porcelain crucible with a lid and heated to 550 °C, applying a heating rate of 300 K/h. This temperature was sustained for 3 h. After cooling to room temperature, a bright yellow solid was obtained in the amount of 0.65 g. It was then crushed into powder (denoted as bCN). Afterward, bCN was subjected to a second polymerization via heating in a partially covered crucible. The yield was 0.53 g of a light-yellow substance which was crushed to powder, making g-C3N4 nanosheets (CNSs). An elemental analysis of bCN and CNS yielded the following: C 29.4, H 2.0, and N 55.9 (wt. %), and C 31.1, H 3.8, and N 59.3 (wt. %).
After that Nitrogen-Doped Carbon Quantum Dots (NCQDs) were synthesized via hydrothermal treatment [53]. In 15 mL of deionized water, 3 g of citric acid and 1 g of urea were combined and dissolved. After stirring at 500 rpm for 30 min, the mixture was heated at 150 °C for 4 h in a PTFE-lined autoclave. The obtained dark-brown solution was centrifuged at 12,000 rpm for 20 min. After that, the solution was filtered with an ultrafiltration membrane (MwCO = 2000) to remove impurities. The final solution was freeze-dried to obtain solid NCQD samples. An analysis of the powder showed that it contained C 36.0, H 6.3, and N 14.7 in wt. %.
Lastly, CNQDs were prepared with different concentrations of NCQD (2, 4, 6, and 8 wt. %) via the hydrothermal method. Firstly, 1 g of CNS was ultrasonically treated in 90 mL of deionized water for 30 min to exfoliate and delaminate the sheet’s structure. Then, ultrasonication was applied for 15 min to disperse NCQDs in 10 mL of distilled water. After that, two solutions were mixed and stirred at room temperature for 30 min. Then, the resulting solution was put into a PTFE-lined autoclave and heated to 120 °C for 4 h. The precipitate was collected, washed three times with distilled water, centrifuged at 6000 rpm for 6 min, and then dried at 70 °C overnight. Four types of CNQDs were created based on NCQD concentration (2, 4, 6, and 8 wt. %).
Finally, a three-step process was carried out in which bCN was made via thermal polymerization with urea as the precursor. bCN underwent a second round of thermal polymerization to modify its morphology. NCQDs then were synthesized using citric acid and urea in the second step. The collected NCQD powder was added to the surface of CNS and the co-catalyst via a hydrothermal method at 120 °C for 4 h. After this process, the degradation of DCF via this catalyst was examined.
The band gap values of the obtained bCN and CNQD were between 2.65 eV to 2.76 eV, while the values of the band gaps of CNS and CNQDs were 2.80 eV and less than 1.00 eV, respectively.
Photolysis and the kinetic rate constant were both measured and graphed (Figure 14).
The initial pH and DCF concentrations were 3.17 and 10 mg/L, respectively. As may be seen in the graph (Figure 14), the best performing catalyst was CNQD-6, with DCF removal of 62%. More or less doped catalysts performed worse in the photolysis tests. This was because while the doping of catalysts helps with visible light absorption and improves charge separation of positive holes (h+) and electrons (e), after a certain point, it can also cause a shield effect and inhibit the process.
Next, photolysis of DCF with different scavengers such as AO (ammonium oxalate), IPA (isopropanol), and p-BQ was done, because they may scavenge both •O2 radicals and h+. The experiment showed that IPA did not influence degradation, but while AO caused a small decrease in degradation, the sample with p-BQ stopped the degradation of DCF after removing 40% of it, probably due to some equilibrium established in the solution.
Lastly, TOC (total organic carbon content) removal during the photocatalysis of DCF over CNQD-6 and the removal of DCF with CNQD-6 were studied. TOC was removed by 34% in 3 h and up to 75% after 10 h. This indicated that an increase in reaction time enhanced DCS mineralization. The removal of DCF by CNQD-6 was measured over three cycles using HPLC. With each cycle, the was slightly lower, i.e., from 62% to 51%, which showed the great stability of the catalyst. During each cycle, C-N=C sp2 hybridized aromatic N atoms were reported as critical components for this oxidation, and during the last cycle, only a slight loss in this band was observed, which indicated excellent stability of g-C3N4.

6.1.6. Porous Fe2O3 Nanoparticles Coupled with CdS Quantum Dots for the Degradation of Bisphenol A

In a study conducted by Liang and colleagues, porous Fe2O3 nanoparticles were coupled with CdS quantum dots for the degradation of Bisphenol A (BPA) [54]. MIL-100(Fe) was synthesized by a direct hydrothermal synthesis. A-Fe2O3 was prepared by heating MIL-100(Fe) at 300 °C for 2 h with a heating rate of 5 °C/min in air; afterwards, the temperature was increased to 450 °C at a rate of 1 °C/min. After reaching 450 °C, the ceramic crucible was removed and a reddish brown powder was observed (F450). Next was the synthesis of X-CdS/450. The first step was to immerse 100 mg of F450 in 20 mL of 0.1 M Cd(NO3)2 aqueous solution for 30 s and then centrifuge the mixture with distilled water at 4000 rpm for 5 min. Then, the collected sample was immersed in 20 mL of 0.1 M Na2S aqueous solution for 30 s and again centrifuged with distilled water. The cycle was repeated 1, 3, 5, and 7 times for different catalyst samples. Next, the cycle samples were dried in a N2 stream. The obtained powder was X-CdS/F450 (X = 1, 3, 5, 7, where x is the number of times of immersion). Lastly, pure CdS was prepared by dispersing 0.1 M Cd(NO3)2 in 200 mL distilled water containing 0.1 M Na2S, which was then stirred overnight. After that, the resulting precipitate was centrifuged with distilled water and dried to obtain a bright-yellow solid. Next, research was conducted, and the results were as follows.
Porous F450 had a surface area of 201 m2/g and a pore volume of 0.26 cm3/g, while modified CdS QDs had a surface area of 79 m3/g. The next experiments were to assess the degradation of Bisphenol A via the obtained catalysts, the effect of H2O2 dosage on BPA degradation, the effect of pH, and how each catalyst worked under different conditions. The best performing catalyst in different environments was 5-CdS/Fe2O3+H2O2, which means that BPA degradation probably occurred through a Fenton-like pathway. The next test further confirmed that among all X-CdS/F450 catalysts, 5-CdS/F450 performed the best, with a degradation time of BPA 30 min. This was because a lower content of CdS led to insufficient visible light absorption, while too much CdS caused the agglomeration of CdS QD nanoclusters which overlapped on the surface and decreased its photoactivity. A pH of 4 was found to be optimal. Finally, the effect of the H2O2 dosage on the degradation of BPA was studied; the best value was 50 µL. This amount of H2O2 produced the most •OH radicals. Lower values of H2O2 produced fewer radicals, while higher concentrations caused the radicals to scavenge to form HOO• radicals, with lower oxidation capacity. The last test was to evaluate the reusability of 5-CdS/F450. After each test, the catalyst was recovered by centrifuge, washed with ethanol and water, centrifuged at 4000 rpm for 5 min, and dried in a vacuum at 100 °C for 4 h. After four cycles, the photocatalytic efficiency decreased from 100% to around 90%, which showed the great stability of the obtained catalyst.

6.1.7. Marimo-like Bi2WO6 and Mammillaria-like ZnO for the Photodegradation of Dimethyl Phthalate (DMP)

Chin et al. formed a composite catalyst using marimo-like Bi2WO6 and mammillaria-like ZnO for application in the photodegradation of dimethyl phthalate (DMP) [55]. Marimo-like Bi2WO6 was synthesized with the hydrothermal method. Firstly, 6 mmol of Bi(NO3)3*5H2O was dissolved in 0.4 M HNO3 and sonicated for 10 min. Subsequently, 3 mmol Na2WO4 solution was added dropwise into the Bi(NO3)3 solution under stirring. The mixture was stirred for 60 min and placed in a Teflon-lined autoclave for 16 h at 175 °C. The product was then washed with ethanol and deionized water and then dried at 65 °C for 24 h. Next, the BWZ composite was made by ultrasonicating 4 mmol of Zn(NO3)2*6H2O and 0.0462 g of marimo-like Bi2WO6 in 80 mL deionized water for 30 min. Next, 0.024 mol of NaOH was added slowly, and the mixture was stirred for 17 h. The collected products were washed several times with ethanol and deionized water and dried at 65 °C in an oven. The dried products were then calcinated at 450 °C for 2 h to obtain 15-BWZ. With this method, other catalysts were also made with different contents of Bi2WO6 (15, 20, and 25), denoted as 15-BWZ, 20-BWZ, and 25-BWZ (Figure 15).
After obtaining the catalysts, research was conducted. The band gaps of each catalyst were measured; the band gap of pure ZnO was estimated to be 3.20 eV and that of Bi2WO6 at 2.94 eV.
First, the photodegradation of DMP under different conditions was tested. The best catalyst was proven to be 20-BWZ (20 wt. % Bi2WO6/ZnO), achieving degradation of 86.6%, whereas pure ZnO and Bi2WO6 only degraded 59.5% and 27.1%, respectively. This effect could be attributed to the fact that a low amount of Bi2WO6 can lead to fewer heterojunctions, while an excessive amount can become the recombination center of charge carriers. The next test was catalyst recycling. During this test, the 20-BWZ composite lost around 14% of its photoactivity (decrease from 86.6% to 72%) after four cycles, which could be connected to the unavoidable loss of catalyst during the photodegradation of DMP. Also, phototoxicity removal by the catalyst was assessed because of possible environmental risks associated with photocatalytically treated DMP. During the test, the 20-BWZ composite showed great phytotoxicity removal, being able to remove a high level of phytotoxicity (82.8%), which shows that the 20-BWZ composite is not only stable in terms of degrading DMP, but also can efficiently detoxify a treated DMP solution.

6.1.8. Hydrothermally Synthesized rGO-TiO2 Composite for Ethylparaben Degradation

Firstly, graphene oxide (GO) was synthesized. This was done by adding 120 mL of concentrated sulfuric acid, 2.5 g of graphite, and 2.5 g of NaNO3 into a beaker under agitation in a cold bath [56]. Subsequently, 15 g of KMnO4 was added slowly in small doses at T < 20 °C. The suspension was continuously agitated for 2 h at 35 °C. Next, 325 mL of water was added to the cold mixture, raising the temperature to 90 °C. After that, 8.83 mL of 33% H2O2 was added to reduce KMnO4, and then agitation was maintained for 30 min. The oxidized material was washed with 10% HCl, and the suspension was centrifuged and washed several times with water until a neutral pH was achieved. The product was then dried in an oven at 60 °C for 24 h to obtain graphene oxide, which was dispersed in 250 mL of water and sonicated for 1 h. After sonication, the dispersion was centrifuged for 30 min at 8000 rpm.
The next prepared compound was a rGO-TiO2 composite. This was done with the hydrothermal method, by adding 3.7 mL of titanium isopropoxide to 3.3 mL of triethanolamine to obtain a 0.5 M Ti(IV) solution. rGO-TiO2 composites were obtained by adding different amounts of GO dispersion (1 mg/mL) to 42.9 mL of a mixture of water with ethanol (proportions 1:14) under continuous agitation. Subsequently, 8.6 mL of 0.5 M Ti(IV) solution was added, followed by agitating for 24 h to obtain a homogeneous solution which was then heated in a stainless steel reactor for 24 h at 180 °C. The resulting solid was washed three times with ethanol, centrifuged at 13,000 rpm for 10 min, and oven-dried at 60 °C. The obtained composites were designated as xrGO-TiO2, with x being the GO content (4, 7, 10, or 30%) [56].
Lastly, rGO-P25 composites were prepared by the same method. Different amounts of GO were added to 90 mL of a water:ethanol (2:1) mixture, which was then sonicated for 30 min. Next, 300 mg of P25 was added and then agitated for 2 h. Next, the mixture was placed into a stainless steel reactor and heated at 120 °C for 3 h. The resulting composite was recovered by filtration, washed several times with deionized water, and dried at 70 °C for 12 h.
Thermogravimetric analysis curves were made for TiO2 and the obtained catalysts. The experiment showed curves in which we could distinguish four mass loss regions [56]. The first one was the slight initial mass loss, corresponding to dehydration and the elimination of water molecules adsorbed on the surface (20–200 °C). The second one (200–325 °C) corresponded to the decomposition of labile oxygenated groups to GO sheets. The third one (325–600 °C) corresponded to the combustion of carbon to more stable groups. Finally, the fourth one (600–800 °C) corresponded to the dihydroxylation process; thus, a higher % of GO in the catalyst caused it to be worse in operation at higher temperatures. The photocatalytic degradation of EtP was measured; the results are shown in Figure 16.
As shown in Figure 16, the worst-performing method of photodegradation of EtP was direct photolysis, while the best was photodegradation with a catalyst containing 7% rGO-TiO2, which was able to photodegrade 98.6% of EtP in 40 min. The presence of rGO sheets in TiO2 favored photocatalytic activity, because they improved the adsorption capacity by increasing the surface area. Graphene is an electron acceptor which suppresses the recombination of photo-generated charge carriers and enhances EtP degradation. rGO acted as a sensitizer by donating electrons to TiO2. Lastly, the presence of C-O-Ti bonds reduced the band gap value to 2.55 eV, increasing the concentration of the generated radicals for the reaction. It was also found that rGO of 7% was the optimal because, for lower values, there were not many particles, so the surface still grew, while for higher values of rGO, performance decreased, because excess particles could cover active sites on the TiO2 surface or act as recombination centers [56].
A comparison of various studies on the photo-based degradation of endocrine disrupting compounds using quantum dots is presented in Table 3.

7. Conclusions

The utilization of quantum dots in photo-based wastewater treatment is a promising and developing area of research. In this study, the analyzed QDs and their heterojunctions showed potential for the photocatalytic and photo-Fenton removal of organic pesticides. For flame retardants and dyes, the main similarity was that, among these substances, the most commonly used are GQD, CQD, and CNQD. Carbon/graphitic quantum dots improved the ability for photodegradation under the visible light. The application of quantum dots on the surface of the different compounds to achieve photodecomposition is a novel, environmentally friendly, and low-cost kind of photodegradation. The reactive oxygen species produced in the process degrade pollutants into organic intermediates and mineralization products; however, total oxidation requires a long time, and only few studies to date have considered TOC, while most research has been focused only on the initial pollutant. Moreover, operational parameters significantly influence the process and therefore should be optimized for each system individually.

Funding

This research received no external funding.

Data Availability Statement

No new data were created or analyzed in this study.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Compendium of Chemical Terminology. Available online: https://goldbook.iupac.org/ (accessed on 28 February 2025).
  2. Cademartiri, L.; Ozin, G.A. Concepts of Nanochemistry; Wiley-WCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2009. [Google Scholar]
  3. Li, S.; Yan, R.; Cai, M.; Jiang, W.; Zhang, M.; Li, X. Enhanced antibiotic degradation performance of Cd0.5Zn0.5S/Bi2MoO6 S-scheme photocatalyst by carbon dot modification. J. Mater. Sci. Technol. 2023, 164, 59–67. [Google Scholar] [CrossRef]
  4. Li, S.; Rong, K.; Wang, X.; Shen, C.; Yang, F.; Zhang, Q. Design of Carbon Quantum Dots/CdS/Ta3N5 S-Scheme Heterojunction Nanofibers for Efficient Photocatalytic Antibiotic Removal. Acta Phys. Chim. Sin. 2024, 40, 2403005. [Google Scholar] [CrossRef]
  5. Wang, H.; Zhang, L.; Chen, Z.; Hu, J.; Li, S.; Wang, Z.; Liu, J.; Wang, X. Semiconductor heterojunction photocatalysts: Design, construction, and photocatalytic performances. Chem. Soc. Rev. 2014, 43, 5234. [Google Scholar] [CrossRef]
  6. Chiu, Y.-H.; Chang, T.-F.M.; Chen, C.-Y.; Sone, M.; Hsu, Y.-J. Mechanistic Insights into Photodegradation of Organic Dyes Using Heterostructure Photocatalysts. Catalysts 2019, 9, 430. [Google Scholar] [CrossRef]
  7. Li, H.; Zhou, Y.; Tu, W.; Ye, J.; Zou, Z. State-of-the-Art Progress in Diverse Heterostructured Photocatalysts toward Promoting Photocatalytic Performance. Adv. Funct. Mater. 2015, 25, 998–1013. [Google Scholar] [CrossRef]
  8. Wen, J.; Zhou, L.; Tang, Q.; Xiao, X.; Sun, S. Photocatalytic degradation of organic pollutants by carbon quantum dots functionalized g-C3N4: A review. Ecotoxicol. Environ. Saf. 2023, 262, 115133. [Google Scholar] [CrossRef]
  9. Wang, D.; Zhu, L.; Chen, J.-F.; Dai, L. Can graphene quantum dots cause DNA damage in cells? Nanoscale 2015, 7, 9894–9901. [Google Scholar] [CrossRef] [PubMed]
  10. Yao, Y.; Zhang, T.; Tang, M. The DNA damage potential of quantum dots: Toxicity, mechanism and challenge. Environ. Pollut. 2023, 317, 120676. [Google Scholar] [CrossRef]
  11. Kamble, S.; Bhosale, K.; Mohite, M.; Navale, S. Methods of Preparation of Nanoparticles. Int. J. Adv. Res. Sci. Commun. Technol. 2023, 3, 121–127. [Google Scholar] [CrossRef]
  12. Mohamed, W.A.A.; El-Gawad, H.A.; Mekkey, S.; Galal, H.; Handal, H.; Mousa, H.; Labib, A. Quantum dots synthetization and future prospect applications. Nanotechnol. Rev. 2021, 10, 1926–1940. [Google Scholar] [CrossRef]
  13. Drbohlavova, J.; Adam, V.; Kizek, R.; Hubalek, J. Quantum Dots—Characterization, Preparation and Usage in Biological Systems. Int. J. Mol. Sci. 2009, 10, 656–673. [Google Scholar] [CrossRef] [PubMed]
  14. Targhan, H.; Rezaei, A.; Aliabadi, A.; Zheng, H.; Cheng, H.; Aminabhavi, T.M. Adsorptive and Photocatalytic Degradation of Imidacloprid Pesticide from Wastewater via the Fabrication of ZIF-CdS/Tpy Quantum Dots. Chem. Eng. J. 2024, 482, 148983. [Google Scholar] [CrossRef]
  15. Targhan, H.; Rezaei, A.; Aliabadi, A.; Ramazani, A.; Zhao, Z.; Shen, X.; Zheng, H. Photocatalytic Removal of Imidacloprid Pesticide from Wastewater Using CdS QDs Passivated by CQDs Containing Thiol Groups. Sci. Rep. 2024, 14, 530. [Google Scholar] [CrossRef]
  16. Nekooei, A.; Miroliaei, M.R.; Shahabi Nejad, M.; Sheibani, H. Enhanced Visible-Light Photocatalytic Activity of ZnS/S-Graphene Quantum Dots Reinforced with Ag2S Nanoparticles. Mater. Sci. Eng. B 2022, 284, 115884. [Google Scholar] [CrossRef]
  17. Malik, A.Q.; Tabasum, S.; Rani, S.; Lokhande, P.; Singh, P.P.; Mooney, J.; Singh, J.; Alberto, H.-A.C.; Sharma, A.; Aepuru, R.; et al. Fluorescent CdS QDs Modified with Molecular Imprinted Polymer for the Photodegradation of Imidacloprid and Buprofezin Pesticides Under Visible Light. J. Inorg. Organomet. Polym. 2023, 33, 3468–3484. [Google Scholar] [CrossRef]
  18. Idrees, S.A.; Jamil, L.A.; Omer, K.M. Silver-Loaded Carbon and Phosphorous Co-Doped Boron Nitride Quantum Dots (Ag@CP-BNQDs) for Efficient Organic Waste Removal: Theoretical and Experimental Investigations. ACS Omega 2022, 7, 37620–37628. [Google Scholar] [CrossRef]
  19. Muangmora, R.; Rojviroon, O.; Kemacheevakul, P.; Chuangchote, S.; Rajendran, R.; Phouheuanghong, P.; Arumugam, P.; Paramasivam, S.; Rojviroon, T. Spent Coffee Ground-Derived Carbon Quantum Dot Composite with Metal Oxides for Photocatalytic Degradation of Carbaryl in Water and Antibacterial Application. J. Water Process Eng. 2025, 70, 107145. [Google Scholar] [CrossRef]
  20. Swedha, M.; Okla, M.K.; Abdel-Maksoud, M.A.; Kokilavani, S.; Kamwilaisak, K.; Sillanpää, M.; Khan, S.S. Photo-Fenton System Fe3O4/NiCu2S4 QDs towards Bromoxynil and Cefixime Degradation: A Realistic Approach. Surf. Interfaces 2023, 38, 102764. [Google Scholar] [CrossRef]
  21. Pervez Md, N.; Ma, S.; Huang, S.; Naddeo, V.; Zhao, Y. Photo-Fenton Degradation of Ciprofloxacin by Novel Graphene Quantum Dots/α-FeOOH Nanocomposites for the Production of Safe Drinking Water from Surface Water. Water 2022, 14, 2260. [Google Scholar] [CrossRef]
  22. John, B.K.; Mathew, B. Nitrogen and Sulphur Co-Doped Carbon Quantum Dot Integrated Bismuth Oxide Nanocomposite for Photocatalytic Degradation and Electrochemical Sensing Applications. Opt. Mater. 2023, 139, 113819. [Google Scholar] [CrossRef]
  23. Abbasi, M.; Aziz, R.; Rafiq, M.T.; Bacha, A.U.R.; Ullah, Z.; Ghaffar, A.; Mustafa, G.; Nabi, I.; Hayat, M.T. Efficient Performance of InP and InP/ZnS Quantum Dots for Photocatalytic Degradation of Toxic Aquatic Pollutants. Environ. Sci. Pollut. Res. 2024, 31, 19986–20000. [Google Scholar] [CrossRef] [PubMed]
  24. Grover, T.; Khandual, A.; Chatterjee, K.N.; Jamdagni, R. Flame retardants: An overview. Colourage 2014, 61, 29–36. [Google Scholar]
  25. Feiteiro, J.; Mariana, M.; Cairrao, E. Health toxicity effects of brominated flame retardants: From environmental to human exposure. Environ. Pollut. 2021, 285, 117475. [Google Scholar] [CrossRef] [PubMed]
  26. Saeed, A.; Altarawneh, M.; Siddigue, K.; Conesa, J.A.; Ortuno, N.; Dlugogorski, B.Z. Photodecomposition properties of brominated flame retardants (BFRs). Ecotoxicol. Environ. Saf. 2020, 192, 110272. [Google Scholar] [CrossRef]
  27. Sahu, R.S.; Dubey, A.; Shih, Y.-h. Novel metal-free in-plane functionalized graphitic carbon nitride with graphene quantum dots for effective photodegradation of 4-bromophenol. Carbon 2021, 182, 89–99. [Google Scholar] [CrossRef]
  28. Bernerd, F.; Passeron, T.; Castiel, I.; Marionnet, C. The Damaging Effects of Long UVA (UVA1) Rays: A Major Challenge to Preserve Skin Health and Integrity. Int. J. Mol. Sci. 2022, 23, 8243. [Google Scholar] [CrossRef]
  29. Zhang, Z.; He, D.; Zhou, Y.; Bai, E.; Qu, J.; Zhang, Y.-n. Fabrication of black phosphorus/CdS heterostructure with enhancement photocatalytic degradation activity for tetrabromobisphenol A and toxicity prediction of intermediates. Environ. Res. 2024, 256, 119060. [Google Scholar] [CrossRef]
  30. Wang, S.; Chen, X.; Fang, L.; Gao, H.; Han, M.; Chen, X.; Xia, Y.; Xie, L.; Yang, H. Double heterojunction CQDs/CeO2/BaFe12O19 magnetic separation photocatalysts: Construction, structural characterization, dye and POPs removal, and the interrelationships between magnetism and photocatalysis. Nucl. Anal. 2022, 3, 100026. [Google Scholar] [CrossRef]
  31. Chen, P.; Di, S.; Qiu, X.; Zhu, S. One-step synthesis of F-TiO2/g-C3N4 heterojunction as highly efficient visible-light-active catalysts for tetrabromobisphenol A and sulfamethazine degradation. Appl. Surf. Sci. 2022, 587, 152889. [Google Scholar] [CrossRef]
  32. Vahabi, H.; Lopez-Cuesta, J.-M.; Chivas-Joly, C. 6-High-performance fire-retardant polyamide materials. In Novel Fire Retardant Polymers and Composite Materials; Woodhead Publishing: Sawston, UK, 2017; pp. 147–170. [Google Scholar] [CrossRef]
  33. Kong, Y.; Lei, T.; He, Y.; Song, G. Background-free room temperature phosphorescence and digital image colorimetry detection of melamine by carbon nitride quantum dots in cellulose matrix with smartphone-based portable device. Food Chem. 2022, 390, 133135. [Google Scholar] [CrossRef]
  34. Al-Gheethi, A.A.; Azhar, Q.M.; Kumar, P.S.; Yusuf, A.A.; Al-Buriahi, A.K.; Mohamed, R.M.S.R.; Al-shaibani, M.M. Sustainable approaches for removing Rhodamine B dye using agricultural waste adsorbents: A review. Chemosphere 2022, 287, 132080. [Google Scholar] [CrossRef] [PubMed]
  35. Oladoye, P.O.; Kadhom, M.; Khan, I.; Aziz, K.H.H.; Alli, Y.A. Advancements in adsorption and photodegradation technologies for Rhodamine B dye wastewater treatment: Fundamentals, applications, and future directions. Green Chem. Eng. 2024, 5, 440–460. [Google Scholar] [CrossRef]
  36. Chen, J.; Shu, J.; Anqi, Z.; Juyuan, H.; Yan, Z.; Chen, J. Synthesis of carbon quantum dots/TiO2 nanocomposite for photo-degradation of Rhodamine B and cefradine. Diam. Relat. Mater. 2016, 70, 137–144. [Google Scholar] [CrossRef]
  37. Tong, S.; Zhou, J.; Ding, L.; Zhou, C.; Liu, Y.; Li, S.; Meng, J.; Zhu, S.; Chatterjee, S.; Liang, F. Preparation of carbon quantum dots/TiO2 composite and application for enhanced photodegradation of rhodamine B. Colloids Surf. A Physicochem. Eng. Asp. 2022, 648, 129342. [Google Scholar] [CrossRef]
  38. Silva, M.; Caro, V.; Guzman, C.; Perry, G.; Areche, C.; Cornejo, A. Cahpter 1—α-Synuclein and tau, two targets for dementia. Stud. Nat. Prod. Chem. 2020, 67, 1–25. [Google Scholar] [CrossRef]
  39. Zhao, F.; Rong, Y.; Wan, J.; Hu, Z.; Peng, Z.; Wang, B. High photocatalytic performance of carbon quantum dots/TNTs composites for enhanced photogenerated charges separation under visible light. Catal. Today 2018, 315, 162–170. [Google Scholar] [CrossRef]
  40. Fan, J.; Li, D.; Wang, X. Effect of modified graphene quantum dots on photocatalytic degradation property. Diam. Relat. Mater. 2016, 69, 81–85. [Google Scholar] [CrossRef]
  41. Aljuaid, A.; Almehmadi, M.; Alsaiari, A.A.; Allahyani, M.; Abdulaziz, O.; Alsharif, A.; Alsaiari, J.A.; Saih, M.; Alotaibi, R.T.; Khan, I. g-C3N4 Based Photocatalyst for the Efficient Photodegradation of Toxic Methyl Orange Dye: Recent Modifications and Future Perspectives. Molecules 2023, 28, 3199. [Google Scholar] [CrossRef] [PubMed]
  42. Guo, Z.; Wu, H.; Li, M.; Tang, T.; Wen, J.; Li, X. Phosphorus-doped graphene quantum dots loaded on TiO2 for enhanced photodegradation. Appl. Surf. Sci. 2020, 526, 146724. [Google Scholar] [CrossRef]
  43. Dong, H.; Zhou, Y.; Wang, L.; Chen, L.; Zhu, M. Oxygen vacancies in piezocatalysis: A critical review. Chem. Eng. J. 2024, 487, 150480. [Google Scholar] [CrossRef]
  44. Wang, K.; Han, C.; Li, J.; Qiu, J.; Sunarso, J.; Liu, S. The Mechanism of Piezocatalysis: Energy Band Theory or Screening Charge Effect? Angew. Chem. Int. Ed. 2022, 61, e202110429. [Google Scholar] [CrossRef] [PubMed]
  45. Ning, X.; Hao, A.; Chen, R.; Khan, M.F.; Jia, D. Constructing of GQDs/ZnO S-scheme heterojunction as efficient piezocatalyst for environmental remediation and understanding the charge transfer mechanism. Carbon 2024, 218, 118772. [Google Scholar] [CrossRef]
  46. Kanigaridou, Y.; Petala, A.; Frontistis, Z.; Antonopoulou, M.; Solakidou, M.; Konstantinou, I.; Deligiannakis, Y.; Mantzavinos, D.; Kondarides, D.I. Solar photocatalytic degradation of bisphenol A with CuOx/BiVO4: Insights into the unexpectedly favorable effect of bicarbonates. Chem. Eng. J. 2017, 318, 39–49. [Google Scholar] [CrossRef]
  47. Petala, A.; Bontemps, R.; Spartatouille, A.; Frontistis, Z.; Antonopoulou, M.; Konstantinou, I.; Kondarides, D.I.; Mantzavinos, D. Solar light-induced degradation of ethyl paraben with CuOx/BiVO4: Statistical evaluation of operating factors and transformation by-products. Catal. Today 2017, 280, 122–131. [Google Scholar] [CrossRef]
  48. Kouvelis, K.; Kampioti, A.A.; Petala, A.; Frontistis, Z. Degradation of Sulfamethoxazole Using a Hybrid CuOx–BiVO4/SPS/Solar System. Catalysts 2022, 12, 882. [Google Scholar] [CrossRef]
  49. Mishra, S.R.; Gadore, V.; Ahmaruzzaman, M. Inorganic–organic hybrid quantum dots for AOP-mediated photodegradation of ofloxacin and para-nitrophenol in diverse water matrices. Npj Clean Water 2023, 6, 78. [Google Scholar] [CrossRef]
  50. Ling, C.; Ye, X.; Zhang, J.; Zhang, J.; Zhang, S.; Meng, S.; Fu, X.; Chen, S. Solvothermal synthesis of CdIn2S4 photocatalyst for selective photosynthesis of organic aromatic compounds under visible light. Sci. Rep. 2017, 7, 27. [Google Scholar] [CrossRef] [PubMed]
  51. Iqbal, A.; Shittu, F.B.; Ibrahim, M.N.M.; Bakar, N.H.H.A.; Yahaya, N.; Rajappan, K.; Hussin, M.H.; Danial, W.H.; Wilson, L.D. Photoreactive Carbon Dots Modified g-C3N4 for Effective Photooxidation of Bisphenol-A under Visible Light Irradiation. Catalysts 2022, 12, 1311. [Google Scholar] [CrossRef]
  52. Patel, J.; Singh, A.K.; Carabineiro, S.A.C. Assessing the Photocatalytic Degradation of Fluoroquinolone Norfloxacin by Mn:ZnS Quantum Dots: Kinetic Study, Degradation Pathway and Influencing Factors. Nanomaterials 2020, 10, 964. [Google Scholar] [CrossRef]
  53. Awang, H.; Peppel, T.; Strunk, J. Photocatalytic Degradation of Diclofenac by Nitrogen-Doped Carbon Quantum Dot-Graphitic Carbon Nitride (CNQD). Catalysts 2023, 13, 735. [Google Scholar] [CrossRef]
  54. Liang, R.; He, Z.; Zhou, C.; Yan, G.; Wu, L. MOF-Derived Porous Fe2O3 Nanoparticles Coupled with CdS Quantum Dots for Degradation of Bisphenol A under Visible Light Irradiation. Nanomaterials 2020, 10, 1701. [Google Scholar] [CrossRef] [PubMed]
  55. Chin, Y.-H.; Sin, J.-C.; Lam, S.-M.; Zeng, H.; Lin, H.; Li, H.; Huang, L.; Mohamed, A.R. 3-D/3-D Z-Scheme Heterojunction Composite Formed by Marimo-like Bi2WO6 and Mammillaria-like ZnO for Expeditious Sunlight Photodegradation of Dimethyl Phthalate. Catalysts 2022, 12, 1427. [Google Scholar] [CrossRef]
  56. Ruidíaz-Martínez, M.; Álvarez, M.A.; López-Ramón, M.V.; Cruz-Quesada, G.; Rivera-Utrilla, J.; Sánchez-Polo, M. Hydrothermal Synthesis of rGO-TiO2 Composites as High-Performance UV Photocatalysts for Ethylparaben Degradation. Catalysts 2020, 10, 520. [Google Scholar] [CrossRef]
  57. Silva, V.; Lima, D.L.D.; de Matos Gomes, E.; Almeida, B.; Calisto, V.; Baptista, R.M.F.; Pereira, G. Electrospun Nanofiber Dopped with TiO2 and Carbon Quantum Dots for the Photocatalytic Degradation of Antibiotics. Polymers 2024, 16, 2960. [Google Scholar] [CrossRef]
  58. Tian, M.; Hu, C.; Yu, J.; Chen, L. Carbon quantum dots (CQDs) mediated Z-scheme g–C3N4–CQDs/BiVO4 heterojunction with enhanced visible light photocatalytic degradation of Paraben. Chemosphere 2023, 323, 138248. [Google Scholar] [CrossRef]
  59. Guo, R.; Zeng, D.; Xie, Y.; Ling, Y.; Zhou, D.; Jiang, L.; Jiao, W.; Zhao, J.; Li, S. Carbon nitride quantum dots (CNQDs)/TiO2 nanoparticle heterojunction photocatalysts for enhanced ultraviolet-visible-light-driven bisphenol a degradation and H2 production. Int. J. Hydrogen Energy 2020, 45, 22534–22544. [Google Scholar] [CrossRef]
  60. Jorge, K.L.-V.; Guzmán-Mar, L.; Montalvo-Herrera, T.J.; Mendiola-Alvarez, S.Y.; Villanueva-Rodríguez, M. Efficient photocatalytic removal of four endocrine-disrupting compounds using N-doped BiOBr catalyst under UV-Vis radiation. J. Environ. Chem. Eng. 2021, 9, 106185. [Google Scholar] [CrossRef]
  61. Hou, J.; Li, H.; Tang, Y.; Sun, J.; Fu, H.; Qu, X.; Xu, Z.; Yin, D.; Zheng, S. Supported N-doped carbon quantum dots as the highly effective peroxydisulfate catalysts for bisphenol F degradation. Appl. Catal. B Environ. 2018, 238, 225–235. [Google Scholar] [CrossRef]
  62. Jiang, H.; Zhong, Y.; Tian, K.; Pang, H.; Hao, Y. Enhanced photocatalytic degradation of bisphenol A over N,S-doped carbon quantum dot-modified MIL-101(Fe) heterostructure composites under visible light irradiation by persulfate. Appl. Surf. Sci. 2022, 577, 151902. [Google Scholar] [CrossRef]
Figure 1. Comparison of imidacloprid degradation in the presence of different amounts of catalysts (A) Ref. [14] catalyst ZIF-CdS/Tpy; (B) Ref. [15] catalyst CQDs-SH/CdS QDs. (Reproduced with permission from Ref. [14] and under the Creative Commons license from Ref. [15].) Reaction conditions: 35 W simulated visible light, the concentration of imidacloprid = 10 ppm, pH = 7, t = 25 °C.
Figure 1. Comparison of imidacloprid degradation in the presence of different amounts of catalysts (A) Ref. [14] catalyst ZIF-CdS/Tpy; (B) Ref. [15] catalyst CQDs-SH/CdS QDs. (Reproduced with permission from Ref. [14] and under the Creative Commons license from Ref. [15].) Reaction conditions: 35 W simulated visible light, the concentration of imidacloprid = 10 ppm, pH = 7, t = 25 °C.
Catalysts 15 00591 g001
Figure 2. Chemical structures of the discussed pesticides.
Figure 2. Chemical structures of the discussed pesticides.
Catalysts 15 00591 g002
Figure 3. Products of imidacloprid degradation determined by LC-MS technique. Reproduced from Ref. [15] under the Creative Commons license.
Figure 3. Products of imidacloprid degradation determined by LC-MS technique. Reproduced from Ref. [15] under the Creative Commons license.
Catalysts 15 00591 g003
Figure 4. Influence of GQD doping dosage (a); H2O2 concentration (b); initial pH values (c); changes of H2O2 concentration under various initial pH values (d); removal rate of pesticide ciprofloxacin (CIP), antibiotic tetracycline (TC), dyes: methylene blue (MB), Orange II (OII), and Bisphenol A (BPA) (e). Reproduced from Ref. [21] under the Creative Commons license.
Figure 4. Influence of GQD doping dosage (a); H2O2 concentration (b); initial pH values (c); changes of H2O2 concentration under various initial pH values (d); removal rate of pesticide ciprofloxacin (CIP), antibiotic tetracycline (TC), dyes: methylene blue (MB), Orange II (OII), and Bisphenol A (BPA) (e). Reproduced from Ref. [21] under the Creative Commons license.
Catalysts 15 00591 g004
Figure 5. The process of synthesis of CN@GQD via the hydrothermal approach. Reproduced with permission from Ref. [27].
Figure 5. The process of synthesis of CN@GQD via the hydrothermal approach. Reproduced with permission from Ref. [27].
Catalysts 15 00591 g005
Figure 6. A diagram showing how much of the total solar spectrum is represented by visible light (VL), how much is represented by UV light, and how much is represented by IRA, as well as the energy/penetration of the types of light emissions. Reproduced under the Creative Commons license from Ref. [28].
Figure 6. A diagram showing how much of the total solar spectrum is represented by visible light (VL), how much is represented by UV light, and how much is represented by IRA, as well as the energy/penetration of the types of light emissions. Reproduced under the Creative Commons license from Ref. [28].
Catalysts 15 00591 g006
Figure 7. (a) Molecular structure of Tetrabromobisphenol A (TBBPA). (b) Change in TBBPA concentration over time during photocatalytic degradation using heterojunctions with varied content of BFO/CeO2; (c) plots of ln(Ct/C0) vs. degradation time and (d) exhibited first-order kinetic constants (k) of TBBPA in the presence of samples S1, S2, S3, and S4 under simulated sunlight radiation. Reproduced under the Creative Commons license from Ref. [30].
Figure 7. (a) Molecular structure of Tetrabromobisphenol A (TBBPA). (b) Change in TBBPA concentration over time during photocatalytic degradation using heterojunctions with varied content of BFO/CeO2; (c) plots of ln(Ct/C0) vs. degradation time and (d) exhibited first-order kinetic constants (k) of TBBPA in the presence of samples S1, S2, S3, and S4 under simulated sunlight radiation. Reproduced under the Creative Commons license from Ref. [30].
Catalysts 15 00591 g007
Figure 8. The molecular structure of Rhodamine B dye.
Figure 8. The molecular structure of Rhodamine B dye.
Catalysts 15 00591 g008
Figure 9. The chemical structure of Methyl Orange.
Figure 9. The chemical structure of Methyl Orange.
Catalysts 15 00591 g009
Figure 10. Change of SMX concentration during removal processes using solar/SPS, solar/catalyst, SPS/catalyst, and solar/SPS/catalyst for (A) 0.75 Cu.BVO, (B) 3.0 Cu.BVO, and (C) 10.0 Cu.BVO in UPW. Experimental conditions: 0.5 mg/L SMX, 100 mg/L SPS, and 500 mg/L catalyst. Reproduced from Ref. [48] under the Creative Commons license.
Figure 10. Change of SMX concentration during removal processes using solar/SPS, solar/catalyst, SPS/catalyst, and solar/SPS/catalyst for (A) 0.75 Cu.BVO, (B) 3.0 Cu.BVO, and (C) 10.0 Cu.BVO in UPW. Experimental conditions: 0.5 mg/L SMX, 100 mg/L SPS, and 500 mg/L catalyst. Reproduced from Ref. [48] under the Creative Commons license.
Catalysts 15 00591 g010
Figure 11. The (a) effect of different pH and (b) pseudo first-order kinetic fitting curves under the visible light irradiation of BPA on its degradation. Reaction conditions: 1.5 CDs/gC3N4 dosage = 30 mg L−1, BPA concentration = 10 mg L−1 and solution volume = 100 mL. Reproduced from Ref. [51] under the Creative Commons license.
Figure 11. The (a) effect of different pH and (b) pseudo first-order kinetic fitting curves under the visible light irradiation of BPA on its degradation. Reaction conditions: 1.5 CDs/gC3N4 dosage = 30 mg L−1, BPA concentration = 10 mg L−1 and solution volume = 100 mL. Reproduced from Ref. [51] under the Creative Commons license.
Catalysts 15 00591 g011
Figure 12. The effect of pH on the photocatalytic degradation of NOFX in the presence of Mn:ZnS QDs under optimal conditions (25 mL of NOFX, 60 min irradiation, 60 mg QDs). Reproduced from Ref. [52] under the Creative Commons license.
Figure 12. The effect of pH on the photocatalytic degradation of NOFX in the presence of Mn:ZnS QDs under optimal conditions (25 mL of NOFX, 60 min irradiation, 60 mg QDs). Reproduced from Ref. [52] under the Creative Commons license.
Catalysts 15 00591 g012
Figure 13. The effect of variation of the amount of catalyst, Mn:ZnS QDs, on the photocatalytic degradation of NOFX under optimal conditions (25 mL of drug, pH 10, 60 min irradiation). Reproduced from Ref. [52] under the Creative Commons license.
Figure 13. The effect of variation of the amount of catalyst, Mn:ZnS QDs, on the photocatalytic degradation of NOFX under optimal conditions (25 mL of drug, pH 10, 60 min irradiation). Reproduced from Ref. [52] under the Creative Commons license.
Catalysts 15 00591 g013
Figure 14. Changes in the concentration of DCF during photocatalytic degradation (a) using bCN, CNS, and CNQD photocatalysts; DCF degradation efficiency and kinetic rate constant of each photocatalyst under the irradiation of visible light (b). Reproduced from Ref. [53] under the Creative Commons license.
Figure 14. Changes in the concentration of DCF during photocatalytic degradation (a) using bCN, CNS, and CNQD photocatalysts; DCF degradation efficiency and kinetic rate constant of each photocatalyst under the irradiation of visible light (b). Reproduced from Ref. [53] under the Creative Commons license.
Catalysts 15 00591 g014
Figure 15. Schematic diagram of synthetic route for fabricating BWZ composites. Reproduced from Ref. [55] under the Creative Commons license.
Figure 15. Schematic diagram of synthetic route for fabricating BWZ composites. Reproduced from Ref. [55] under the Creative Commons license.
Catalysts 15 00591 g015
Figure 16. The kinetics of photocatalytic degradation of EtP under UV radiation in the presence of rGO-TiO2 composites as a function of treatment time. [EtP]0 = 0.30 × 10−3 mol/L, [catalyst]0 = 0.7 g/L. Reproduced from Ref. [56] under the Creative Commons license.
Figure 16. The kinetics of photocatalytic degradation of EtP under UV radiation in the presence of rGO-TiO2 composites as a function of treatment time. [EtP]0 = 0.30 × 10−3 mol/L, [catalyst]0 = 0.7 g/L. Reproduced from Ref. [56] under the Creative Commons license.
Catalysts 15 00591 g016
Table 1. Summary of discussed examples of the usage of quantum dots in photo-based pesticide degradation.
Table 1. Summary of discussed examples of the usage of quantum dots in photo-based pesticide degradation.
MaterialPesticideRemoval Efficiency/Constant of Degradation RateConditionsRef.
CQDs-SH/CdS QDsImidacloprid75.60% under simulated visible light
92.20% under UV-C
catalyst dose 1 g/L
pesticide dose 10 ppm
pH 7
t = 25 °C
Degradation time 90 min
[15]
ZnS/SG QDs/Ag2S 5 mg/25 mLDiazinon
Fenitrothion
0.053 min−1
0.056 min−1
catalyst dose (0.2 g/L)
pesticide dose (10 g/L)
t = 25 °C
60 W LED light
[16]
CdS QDs/MIPImidacloprid (pH = 9)
Buprofezin (pH = 10)
81%
72%
catalyst dose 10 ppm
pesticide dose 10 ppm
Degradation time 120 min
UV light
[17]
CdS QDs/NIPImidacloprid (pH = 9)
Buprofezin (pH = 10)
76%
68%
catalyst dose 10 ppm
pesticide dose 10 ppm
Degradation time 120 min
UV light
[17]
Ag@CP-BNQDs Chlorothalonil0.0053 min−1 catalyst dose (1 g/L)
pesticide dose (25 ppm)
t = 25 °C
5 W of blue LED
[18]
ZIF-CdS/Tpy QDs (0.7 g/L)Imidacloprid90.95%
93.12%
catalyst dose 0.7 g/L
pesticide dose 10 ppm
35 W LED lamp
Degradation time 90 min
pH 7
[14]
InP QDs
InP/ZnS QDs
Deltamethrin93.36%
90.29%
catalyst dose (10 mg/kg)
300 W Degradation time 90 min
Xe lamp
[23]
TiO2/ZnO/CQD Carbaryl99.01%, 0.0570 min−1catalyst dose 1 g/L
pesticide dose 5 mg/L
Degradation time 60 min
[19]
N,S-CQD-Bi2O3 2,4-D
Diuron
92%
97%
catalyst dose 25 mg/30 mL
pesticide dose 20 mg/L
degradation time 2 h
[22]
Ni2CuS4 QDs@30% Fe3O4Bromoxynil0.0070 min−1catalyst dose 20 mg/L
pesticide dose 20 mg/L
1000 W halogen lamp
degradation time 200 min
amount of H2O2 100 µL
[20]
GQD/α-FeOOHCiprofloxacin93.73%catalyst dose 0.25 mg/L
pesticide dose 10 mg/L
350 W Xe lamp
degradation time 60 min
H2O2 dose 0.50 mM
[21]
Table 2. A summary of endocrine disrupting compounds degraded in photo-based processes using quantum dots.
Table 2. A summary of endocrine disrupting compounds degraded in photo-based processes using quantum dots.
No.Chemical StructureName
1Catalysts 15 00591 i001Ethylparaben
2Catalysts 15 00591 i002Bisphenol A
3Catalysts 15 00591 i003Sulfamethoxazole
4Catalysts 15 00591 i004Norfloxacin
5Catalysts 15 00591 i005Diclofenac
6Catalysts 15 00591 i006Dimethyl Phthalate
7Catalysts 15 00591 i007Para-nitrophenol
8Catalysts 15 00591 i008Sulfadiazine
Table 3. Comparison of the described catalysts, including information on degraded contaminant, efficiency, time, and conditions.
Table 3. Comparison of the described catalysts, including information on degraded contaminant, efficiency, time, and conditions.
CatalystCECEfficiency [%]Time [min]ConditionsRef.
3.0 Cu.BVOsulfamethoxazole10022Solar/SPS[48]
Cd/CdIn2S4@Chp-Nitrophenol96.730H2O2, pH = 3[49]
ofloxacin85.590H2O2, pH = 7
1.5CDs/g-C3N4BPAOver 90180pH = 10[51]
Mn:ZnSNorfloxacin8660pH = 10 [52]
CNQD-6Diclofenac62180Light irradiation[53]
5-CdS/F450BPA10030pH = 4, H2O2[54]
20-BWZDimethyl Phthalate86.690Light irradiation[55]
7% rGO-TiO2Ethylparaben98.640Light irradiation[56]
PA66/TiO2/CQDsAMXOver 90120pH = 8[57]
SDZ88240pH = 8
GCN-CQDs/BVOBenzyl Paraben85.7150pH = 4, light[58]
CNQDs-2.0/TiO2BPA10080UV light[59]
N20BiOBrE2100240pH = 6.5[60]
EE2100240
4TOP100240
BPA100240
NCQD(2.10)/Al2O3-700BPF100120pH = 6.4, 25 °C[61]
5-N,S:CQD/MIL-101(Fe)BPA10060PS/VIS[62]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Matyszczak, G.; Yedzikhanau, A.; Jasiak, C.; Bojko, N.; Krawczyk, K. Applications of Quantum Dots in Photo-Based Advanced Oxidation Processes for the Degradation of Contaminants of Emerging Concern—A Review. Catalysts 2025, 15, 591. https://doi.org/10.3390/catal15060591

AMA Style

Matyszczak G, Yedzikhanau A, Jasiak C, Bojko N, Krawczyk K. Applications of Quantum Dots in Photo-Based Advanced Oxidation Processes for the Degradation of Contaminants of Emerging Concern—A Review. Catalysts. 2025; 15(6):591. https://doi.org/10.3390/catal15060591

Chicago/Turabian Style

Matyszczak, Grzegorz, Albert Yedzikhanau, Christopher Jasiak, Natalia Bojko, and Krzysztof Krawczyk. 2025. "Applications of Quantum Dots in Photo-Based Advanced Oxidation Processes for the Degradation of Contaminants of Emerging Concern—A Review" Catalysts 15, no. 6: 591. https://doi.org/10.3390/catal15060591

APA Style

Matyszczak, G., Yedzikhanau, A., Jasiak, C., Bojko, N., & Krawczyk, K. (2025). Applications of Quantum Dots in Photo-Based Advanced Oxidation Processes for the Degradation of Contaminants of Emerging Concern—A Review. Catalysts, 15(6), 591. https://doi.org/10.3390/catal15060591

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

Article Metrics

Back to TopTop