Next Article in Journal
Mixed-Ligand Copper(II) Complexes Derived from Pyridinecarbonitrile Precursors: Structural Features and Thermal Behavior
Previous Article in Journal
Structure and Electrochemical Performance of Glasses in the Li2O-B2O3-V2O5-MoO3 System
Previous Article in Special Issue
Enhanced Photocatalytic Performances and Mechanistic Insights for Novel Ag-Bridged Dual Z-Scheme AgI/Ag3PO4/WO3 Composites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Carbon Dots-Based Photocatalysts for Water Treatment Applications

by
Adamantia Zourou
1,
Afrodite Ntziouni
1,
Alexandra Karagianni
1,
Niyaz Alizadeh
2,
Nikolaos Argirusis
2,
Maria Antoniadou
3,
Georgia Sourkouni
4,
Konstantinos V. Kordatos
1 and
Christos Argirusis
1,*
1
School of Chemical Engineering, National Technical University of Athens, Zografou, 15773 Athens, Greece
2
mat4nrg GmbH, Burgstätter Str. 42, 38678 Clausthal-Zellerfeld, Germany
3
School of Chemical Engineering, University of Western Macedonia, Koila, 50150 Kozanis, Greece
4
Clausthal Center of Materials Technology (CZM), Technische Universität Clausthal, Leibnizstr. 9, 38678 Clausthal-Zellerfeld, Germany
*
Author to whom correspondence should be addressed.
Inorganics 2025, 13(9), 286; https://doi.org/10.3390/inorganics13090286
Submission received: 26 June 2025 / Revised: 7 August 2025 / Accepted: 22 August 2025 / Published: 26 August 2025
(This article belongs to the Special Issue Inorganic Photocatalysts for Environmental Applications)

Abstract

Carbon dots (CDs), a rapidly emerging class of zero-dimensional (0-D) nanomaterials with small particle sizes (<10 nm), have garnered significant scientific interest owing to their exceptional physicochemical properties, non-toxicity, low-cost synthesis, and versatile applications. In recent years, the combination of various inorganic photocatalysts (e.g., metal oxides, metal chalcogenides, metal oxyhalides, MXenes, non-metallic semiconductors) with CDs has gained momentum as a promising strategy to enhance their photocatalytic efficiency. By incorporating CDs, researchers have addressed fundamental challenges in photocatalytic systems, including limited light absorption range, rapid electron–hole recombination rate, low quantum efficiency, etc. The present review is focused on the most recent developments in CDs-based heterostructures for advanced photocatalytic applications, particularly in the field of environmental remediation, providing a comprehensive overview of emerging strategies, synthesis approaches, and the resulting enhancements in photocatalytic water treatment applications.

1. Introduction

1.1. The Role of Photocatalysis in Water Treatment

Undoubtedly, water is a fundamental prerequisite for the existence and sustainability of life on Earth. Nevertheless, water quality encounters severe challenges, due to various factors, including natural processes (e.g., geochemical activity, forest fires, etc.), anthropogenic influences (e.g., agricultural and mining activities, industrial discharges, rapid population growth, etc.), and climate change, resulting in increased levels of pollutants and contaminants in water bodies [1]. The presence of emerging contaminants into the aqueous environment, such as heavy metals [2], pesticides [3], pharmaceuticals [4], dyes [5], polycyclic aromatic hydrocarbons [6], microplastics [7], etc., is a major concern for humans, aquatic species, animals, and vegetations. Therefore, their removal from water is vital for the protection of both human health and ecosystems. The scientific community has dedicated efforts to the successful elimination of water pollutants via several physicochemical methods, including photocatalysis [8], membrane separation [9], coagulation/flocculation [10], adsorption [11], biodegradation [12], etc.
Among these techniques, photocatalysis has emerged globally as a sustainable solution because of its high effectiveness, low-cost, and the ability to degrade a wide range of pollutants under mild conditions. Although the term “photocatalysis” has been a subject of substantial confusion and debate regarding its precise definition, it is officially restricted to reactions that occur in the presence of a semiconductor and illumination, according to the International Union of Pure and Applied Chemistry (IUPAC) [13]. Briefly, during a photocatalytic process (Figure 1), the semiconductor absorbs incident photons with energy equal to or greater than its band gap (hv ≥ Eg), leading to the excitation of electrons (e) from the valence band (VB) to the conduction band (CB) and leaving behind holes (h+) in the VB. Electrons and holes can participate in various surface chemical reactions; however, these charge carriers may also undergo a recombination step followed by the production of phonons or heat. When the separated holes interact with H2O, they oxidize donor molecules and form ·OH radicals, while the separated electrons react with dissolved O2 to produce superoxide radicals (O2·), which can further react to generate ·OH. The above reduction/oxidation (redox) reactions are listed below (Equations (1)–(9)):
Semiconductor + hv → h+ + e
e + O2 → O2·−
O2·− + H+·OOH
2 ·OOH → O2 + H2O2
H2O2 + O2·−·OH + OH + O2
H2O2 + hv → 2 ·OH
h+ + H2O → ·OH + H+
h+ + OH·OH
Pollutants + (·OH, h+, e, ·OOH, ·O2) → Degradation Products

1.2. Limitations of Inorganic Photocatalysts

Over the years, various semiconductors have been employed as photocatalysts for the elimination of both organic and inorganic contaminants from water. Based on the type and composition, inorganic photocatalysts could be broadly clustered into the following categories: i. metal oxides (e.g., TiO2, ZnO, Fe2O3, BiO2, SrO2, WO3, etc.) [15], ii. metal chalcogenides, including sulfides (e.g., CdS, ZnS, Ag2S, MoS2, In2S3, etc.) [16], selenides (e.g., CdSe, ZnSe, MoSe2, WSe2, Ag2Se, Bi2Se3, etc.) and tellurides (e.g., CdTe, ZnTe, etc.) [17], iii. metal halides/oxyhalides (e.g., AgCl, BiOBr, BiVO4, etc.) [18], iv. 2-D metal carbides/nitrides (MXenes) (e.g., Ti3C2, V2ZnC, Ti3ZnC2, Ti2ZnN, etc.) [19], and v. metal-free materials (e.g., graphitic C3N4 (g–C3N4), etc.) [20]. Each class of materials exhibits distinct physicochemical properties, which make them suitable for various photocatalytic applications such as environmental remediation [21], antibacterial disinfection [22], self-cleaning [23], hydrogen evolution [24] and CO2 reduction [25]. Metal oxides present unique electronic versatility, ranging from conductive to semiconductive and insulating, due to the wide band gap that influences their photocatalytic performance, while they are known for their high chemical and thermal stability [26]. Metal chalcogenides typically show narrower band gaps, compared to metal oxides and thus demonstrate remarkable light-harvesting capacity. In addition, they show effective charge carrier migration capacity [16]. Recently, 2-D MXenes have attracted significant attention due to their interesting properties, such as hydrophilicity, high specific surface area, tunable surface chemistry, and exceptional electrical conductivity, all of which contribute to superior charge carrier dynamics and reduced recombination rates [27]. Last but not least, g–C3N4, with exceptional chemical stability and a favorable electronic band structure, stands as a prominent and sustainable non-metal photocatalyst [20].
Despite their unique advantages, the aforementioned inorganic photocatalysts exhibit certain limitations that may hinder their practical applications. For instance, metal oxides, such as the well-studied TiO2 [28] and ZnO [29], possess large band gaps (~3.2 eV and ~3.3 eV, respectively) and they could be activated under ultraviolet (UV) irradiation. This results in the inefficient exploitation of solar light since UV light only accounts for ~4–5% of the sunlight spectrum. Furthermore, metal oxides suffer from low quantum efficiency, arising from the high charge carrier recombination rate [30]. Metal chalcogenides also undergo rapid carrier recombination of e–h+. In addition, they are susceptible to photo-corrosion and often contain toxic elements (e.g., Cd), limiting their environmental applicability [31]. Metal halides and oxyhalides, including perovskite materials, typically exhibit limited visible (Vis) light absorption and encounter stability issues in the aqueous environment [32]. Moreover, they may release toxic ions (e.g., Pb2+), which raises concerns regarding their environmental safety. Although MXenes have a number of attractive properties due to their unique 2-D structure, they are vulnerable to oxidative degradation reactions with water and/or oxygen because of their inherent defective sites [33]. They also face restacking and aggregation phenomena due to van der Waals interactions between nanosheets, which could hinder their photocatalytic activity [34]. Similarly, a 2-D metal-free g–C3N4 semiconductor faces aggregation tendencies. Furthermore, its modest visible light absorption and rapid charge carrier recombination lead to decreased efficiency [35].
In order to overcome these drawbacks, several strategies have been proposed, including the following: i. doping with metal or non-metal elements so as to decrease the band gap and enhance visible light absorption [36], ii. combination with carbon nanoallotropes [37] or conductive polymers [38] to improve the charge separation and reduce the recombination rates, or iii. surface modification to eliminate toxicity [39] and photo-corrosion [40], as well as agglomeration phenomena [41]. It is evident that the growing demand for enhanced photocatalytic properties has driven ongoing investigations into material combinations, with novel functionalities prompting the scientific community to shift emphasis toward hybrid materials [42].
Following this direction, the integration of inorganic photocatalysts with carbon dots (CDs)—the latest addition in the broad family of carbon-based nanomaterials—has emerged as a promising strategy and has garnered considerable attention over the past few years. CDs are currently at the forefront of research in nanomaterials due to their advantageous properties such as their fascinating optical properties, abundance of surface functional groups, non-toxicity, etc. The aim of this review is to highlight the most recent advances on the scientific area of CDs-based inorganic photocatalysts for water treatment applications, as reported in the literature.

2. The Combination of Inorganic Photocatalysts with Carbon Dots (CDs)

2.1. CDs (Structure, Properties, Synthesis)

In 2004, Xu et al. [43] serendipitously observed fluorescent carbon nanoparticles during the electrophoretic purification of single-walled carbon nanotubes. This observation, while not directly related to CDs, contributed to the early conceptualization of CDs as a unique class of carbon-based nanomaterials. They are typically small size (<10 nm) quasi-spherical nanoparticles, consisting of a graphitic, amorphous or semi-crystalline carbon core, surrounded by a large number of functional groups, as shown in Figure 2a. Due to the development of several top-down and bottom-up methods, as well as the abundance of precursor materials (e.g., graphite, polymers, small organic molecules, agricultural by-products, etc.), different structures of CDs could be obtained. Therefore, based on their structure, CDs are classified into the following groups: i. carbon quantum dots (CQDs), ii. graphene quantum dots (GQDs), iii. carbon nanodots (CNDs), and iv. carbon polymer dots (CPDs) (Figure 2b). Briefly, CQDs consist of a crystalline graphitic core surrounded by a multitude of functional groups. Similarly, GQDs are composed of segments of one or a few (<5) graphene (or graphene oxide) sheets encompassed by various surface functional groups. On the other hand, CNDs possess an amorphous carbon core. In contrast to the aforementioned categories, CPDs exhibit a low degree of carbonization and contain numerous functional groups and polymeric chains both in the core and the outer shell [44].
Due to their outstanding physicochemical properties, CDs have been utilized in numerous research domains, including photocatalysis. First and foremost, CDs are well known for their strong light absorption in the UV–Vis region (Figure 3a). This is attributed to the π–π* electron transitions associated with sp2-hybridized C=C bonds in the carbon core, as well as the n–π* electron transitions taking place in the shell, facilitated by the presence of abundant functional groups (e.g., -OH, -COOH, -NH2, etc.) [46]. Furthermore, the photoluminescence (PL) is probably one of the most fascinating characteristics of CDs; however, its origin still remains a highly debated aspect. Studies have shown that both carbon core (Figure 3b) and surface state (Figure 3c) play a critical role in the PL mechanism. More specifically, sp2 carbon cores could produce size-dependent fluorescence because of the quantum confinement effect, whereas surface functional groups could introduce localized energy levels, which trap excitons, leading to radiative emission [47]. Another important aspect in the optical performance of CDs is their resistance to photochemical degradation, which makes them ideal candidates for applications where materials are exposed to light for prolonged time and high photostability is required [48]. In addition to their remarkable optoelectronic properties, CDs are environmentally friendly and non-toxic nanomaterials [49], making them highly attractive to a wide range of technological fields. In addition, the ability to synthesize CDs using low-cost precursors, such as biomass waste and industrial or agricultural by-products, highlights their cost-effectiveness and sustainability [50]. Lastly, the rich surface functionality of CDs results not only in excellent water solubility, but also in the interactions with other molecules (e.g., water pollutants) or materials, forming novel heterostructures with enhanced properties.
CDs are synthesized via numerous techniques, which are classified into top-down and bottom-up methods (Figure 3d); the former approach involves the break-down of larger carbon-based structures (e.g., graphene), while the latter approach is based on the polymerization and carbonization of small organic molecules. The synthesis of CDs via top-down approaches involves techniques such as laser ablation [51], arc discharge, and chemical/electrochemical oxidation [52], which typically demand high cost and advanced equipment. On the other hand, bottom-up methods including hydrothermal/solvothermal synthesis [53,54,55] are facile and cost-effective. Notably, the synthesis method of CDs strongly influences their structural and physicochemical properties. CDs produced via a top-down approach typically present a wide size distribution and low quantum yield due to the introduction of surface defects during the harsh processing conditions. Nevertheless, the graphitic core is preserved, since the precursors are graphitic materials (e.g., graphite, graphene, etc.), resulting in enhanced conductivity. In addition, it has been suggested that the top-down methods enable superior control over structure and purity of the final products. In contrast, CDs prepared via bottom-up methods, such as hydrothermal synthesis, present improved control over size and surface passivation, leading to higher fluorescence quantum yields, as well as tunable optical properties. However, by-products and agglomerates could be formed as a concurrent process during the bottom-up synthesis of CDs, which may compromise their physicochemical characteristics. Therefore, further purification of the final product, via filtration, dialysis, electrophoresis, or chromatography is needed; this step is crucial to ensure the quality and reproducibility of the as-synthesized CDs, especially when aimed at optoelectronic or biomedical applications.
Figure 3. (a) Light absorption of CDs in the UV–Vis range [56]. The fluorescent properties of CDs are associated with (b) the quantum confinement effect and (c) the surface state. (Adopted from [47] with permission). (d) Top-down and bottom-up synthesis methods of CDs [57].
Figure 3. (a) Light absorption of CDs in the UV–Vis range [56]. The fluorescent properties of CDs are associated with (b) the quantum confinement effect and (c) the surface state. (Adopted from [47] with permission). (d) Top-down and bottom-up synthesis methods of CDs [57].
Inorganics 13 00286 g003

2.2. In Situ and Ex Situ Synthesis of CDs-Based Heterostructures

The integration of CDs with various inorganic materials, such as metal oxides, metal chalcogenides, MXenes, metal oxyhalides or non-metallic materials can be achieved through two main approaches: in situ and ex situ synthesis (Figure 4). The in situ synthesis involves three plausible routes: i. utilization of pre-formed CDs and inorganic precursors, ii. utilization of carbon-based precursors and pre-formed inorganic materials, or iii. employing the starting materials of both CDs and inorganic materials, simultaneously. By enabling the in situ synthesis of CDs alongside inorganic materials, either sequentially or simultaneously, strong interactions between the components and enhanced homogeneous dispersion are achieved, attracting the interest of materials scientists and engineers. Among various in situ strategies, hydrothermal or solvothermal synthesis is the most commonly employed and widely discussed in the literature; this is due to its low-cost, environmental friendliness, simplicity, control over reaction conditions and the ability to obtain products with high crystallinity and purity. During the hydrothermal/solvothermal method, the precursor mixture is placed into a Teflon-lined autoclave and heated under pressure, higher than atmospheric. The physicochemical properties of the as-prepared materials depend on the selected precursors and solvent, as well as the reaction conditions (e.g., temperature, time, pH, etc.). This review highlights a wide range of studies that rely on the in situ synthesis of CDs-based heterostructures using a hydrothermal/solvothermal process. On the other hand, the ex situ approach refers to the combination of CDs and inorganic materials that have been prepared separately in a prior step. Although the ex situ approach is more favorable for the large-scale production of hybrid materials, it presents some drawbacks, including low uniform dispersion and relatively weak interactions among the components. However, the nature of interactions depends on the synthesis method. For instance, when CDs and inorganic particles are mixed under ambient conditions via magnetic stirring, electrostatic interactions are more likely to dominate, whereas hydrothermal conditions may promote the formation of both covalent and non-covalent bonds.

2.3. CDs–Metal Oxide Heterostructures

Metal oxides constitute a versatile class of semiconductor materials with energy band gaps typically ranging from 1.0 to 4.0 eV (Figure 5a), which are widely employed in photocatalytic applications, including water treatment, due to their favorable electronic properties, chemical, and thermal stability, as well as their abundance [26]. Nevertheless, they face certain limitations, such as low exploitation of the full solar spectrum and low quantum efficiency due to the high charge carrier recombination rate, hindering their overall performance. Thus, various strategies have been proposed in order to enhance their photocatalytic activity. Among them, the modification of metal oxides with non-toxic and low-cost CDs has garnered considerable attention in the scientific community, reflected in the growing number of research articles on the above topic (Figure 5b).
Within the broad family of metal oxides, titanium dioxide (TiO2) stands out as the most well-known photocatalyst. Due to its cost-effectiveness, chemical stability, and intrinsic non-toxicity, TiO2 is still one of the most attractive photocatalytic materials. Yet, its wide band gap (≈3.2 eV for anatase, ≈3.0 eV for rutile) restricts absorption almost exclusively to UV light, which represents <5% of the solar spectrum. In addition, the fast recombination of photogenerated electrons and holes keeps the quantum yield low. By introducing dopants, tailoring nanostructures (nanotubes, mesoporous spheres, hierarchical films) to increase the active surface or creating heterojunctions with CDs, researchers have succeeded in extending its absorption into the visible range and enhancing charge separation, thereby unlocking its potential for solar-driven degradation of dyes, pesticides, and other contaminants in agro-industrial wastewaters.
To begin with, various CDs-TiO2 hybrid photocatalysts have been used for the degradation of various compounds such as organic dyes and antibiotics, including [58] tetracycline (TC) and amoxicillin (AMX), as well as the reduction of hexavalent Cr (VI). For instance, Camilli et al. [59] investigated the role of different surface-functionalities of CDs in the enhancement of the visible-light photocatalytic performance of TiO2 nanoparticles (NPs). More specifically, OH-functionalized, N-functionalized, and P-functionalized CDs were synthesized and used for the production of CDs-coated TiO2 hybrid materials, followed by photocatalytic experiments in the degradation of glucose. In all cases, the addition of CDs to TiO2 led to decrease of the band gap from 3.08 eV for the pristine TiO2 down to 2.92 eV, 2.83 eV, and 2.89 eV for OH-CDs-TiO2, N-CDs-TiO2, and P-CDs-TiO2, respectively, whereas the N-CDs heterostructure exhibited the best performance in glucose degradation. In addition, this work enhances the sustainability and green nature of the photocatalytic process by incorporating CDs obtained from waste materials. Recently, a CDs–TiO2 hybrid material was developed by Sendao et al. [60], utilizing corn stover as a precursor of CDs, which is a major agricultural waste. The heterostructure was tested under visible light on multiple dyes, including Methylene Blue (MB), Rhodamine B (RhB), Orange II, and Methyl Orange (MO), achieving a degradation efficiency of 90.8%, 81.7%, 67.9%, and 37.0%, respectively. Furthermore, the performance of the photocatalyst in a mixed-dye system consisting of MB and RhB, was also examined, where CDs–TiO2 showed a selective preference toward MB over RhB with a degradation efficiency of 89.1% and 43.7%, respectively. Ma et al. [61] developed, via hydrothermal treatment, a CDs–TiO2 hybrid photocatalyst, using CDs derived from peanut shells. The hybrid material was tested in order to evaluate its photocatalytic performance in the reduction of Cr (VI), as it is presented in Figure 6. Various experimental factors, such as CDs-content (1–3%), photocatalyst dosage, and pH value were investigated. Regarding the CDs loading, the hybrid material containing 1% of CDs presented the best performance, whereas the increase of the photocatalyst concentration from 0.1 g/100 mL to 0.3 g/100 mL led to complete reduction of Cr (VI) within 60 min. Furthermore, the CDs–TiO2 photocatalyst was tested in domestic sewage water achieving a reduction efficiency of 90.09%, whilst also exhibiting photocatalytic disinfection of E. coli. Herein, it is important to note that five regeneration cycles were conducted for the hybrid material presenting the best performance (CDs (1%)–TiO2) in order to evaluate its practical applicability. Interestingly, even after the last cycle, the reduction efficiency was 96.83%. The characterization of the material before and after recycling with several techniques (e.g., FT-IR, XRD, XPS), showed that the chemical environment of the photocatalyst did not undergo significant alterations. Consequently, the combination of enhanced photocatalytic performance and excellent stability makes it well-suited for real-world photocatalytic applications.
Several research groups have engineered CDs–TiO2-based photocatalysts with more complex morphologies to enhance visible-light harvesting and promote charge-carrier separation at heterojunctions. On this basis, Rawat et al. [62] reported a two-steps synthesis procedure of CDs–TiO2 photocatalytic heterostructure (Figure 7a). Firstly, CDs were prepared via microwave irradiation of allylamine and ascorbic acid precursors. Following this, the CDs–TiO2 mixture was converted into CDs-decorated TiO2 nanotubes (NTs) through hydrothermal treatment. The CDs–TiO2 NTs outperform both unmodified TiO2 and CDs–TiO2 NPs in the visible-light degradation of TC. In addition, the band gap shrinks from 3.02 eV for TiO2 to 2.96 eV for CDs–TiO2 NPs and 2.92 eV for CDs–TiO2 NTs (Figure 7b). The results indicate that the morphological characteristics of TiO2 have a significant effect in the photocatalytic performance.
Moreover, a CDs-decorated Ni-doped TiO2 hybrid photocatalyst in which single Ni atoms are incorporated into the TiO2 lattice and CDs onto the surface of TiO2 was developed by Bui et al. [63]. The photocatalytic performance was evaluated by testing the degradation of ciprofloxacin (CPX). Integrating CDs onto Ni-doped TiO2, results in a visible-light-responsive photocatalyst for antibiotic degradation, combining single-atom-induced oxygen vacancies (OVs) alongside CDs-driven up conversion and improved charge separation. Under visible-light irradiation for 150 min, the degradation efficiency of commercial TiO2 was 27%, while sol-gel synthesized TiO2 achieved 60% removal. Introducing Ni into the TiO2 lattice increased degradation to 80%, achieving the same percentage of degradation as CDs-decorated TiO2 (without Ni). However, the CDs–Ni-doped TiO2 hybrid photocatalyst achieved 97% CPX removal from water. Furthermore, Taghiloo et al. [64] reported the design of a multifunctional CQD–TiO2-based hydrogel, which combines dye adsorption and visible-light photocatalysis for wastewater treatment. Photocatalytic degradation studies of MB were performed evaluating the photocatalyst dose, initial pollutant concentration and pH value. Moderate pollutant concentration, an optimal photocatalyst dose (8 mg/20 mL), and mildly basic conditions (pH = 8–10) yielded the best combined adsorption–photodegradation performance. The CQDs–TiO2–based hydrogel achieved 97% degradation of MB under visible-light irradiation, outperforming bare TiO2 NPs. This degradation efficiency is attributed to the synergistic combination of dye adsorption by the hydrogel’s network and photocatalytic degradation at the CQDs–TiO2 interface. Moreover, the photocatalytic-based hydrogel maintained its high performance over five consecutive degradation cycles, demonstrating good reusability.
Zinc oxide (ZnO) is also a well-studied semiconductor, chemically and physically stable, exhibiting high photosensitivity and biocompatibility, making it a promising photocatalytic material. However, its relatively wide bandgap energy of 3.37 eV limits its photocatalytic efficiency under visible light. Due to their high surface-to-volume ratio and nanoscale size, ZnO NPs serve as effective photocatalysts. To enhance their performance, N-doping, or/and the incorporation of nanocarbon-based supporting materials, such as CDs, have been shown to reduce the bandgap energy and suppress the recombination rate of electron–hole (e–h+) pairs, thereby improving overall photocatalytic activity.
ZnO decorated with CQDs have been presented by several research groups for the photocatalytic degradation of dyes, especially MB [65,66,67,68], but also MO [69]. In most cases, the CDs–ZnO hybrid materials are synthesized via hydrothermal or solvothermal methods. In their study, Bhavsar et al. [70] achieved the increase of photodegradation of MB, MO, and RhB from 82.26%, 46.46%, and 62.87% assigned to pristine ZnO NPs, to 90.87%, 57.95%, and 86.70% for CDs-decorated ZnO, respectively. In general, various studies indicate that CDs introduce more active sites on the ZnO surface, while they increase the absorbance in the UV–Vis light range, thus improving the overall photocatalytic performance. In another work conducted by Hidayat et al. [66], a 91% degradation of the MB dye within 30 min is reported, where Ayu et al. [65] used CDs–N-doped ZnO photocatalyst and achieved decolorization of MB under basic conditions (90.1%), neutral pH (83.4%) and acidic pH (67.2%). The latter study also investigates the recyclability performance of the hybrid material after three reaction cycles in neutral pH (pH = 7.04). The photocatalytic efficiency was 83.4%, 70.6%, and 58.2% for cycle 1, cycle 2, and cycle 3, respectively. In addition, the structure of the photocatalytic material before and after its utilization for 3 cycles was characterized via XRD analysis and the results revealed that there was no difference in the crystal structure of the CDs–N-doped ZnO sample. Similarly, the morphology of the photocatalytic material remained unchanged, based on FE-SEM characterization. As a result, the hybrid material presented high photocatalytic performance and stability.
Experimental parameters, such as the pH of the solution, catalyst dosage, initial concentration of pollutants (Figure 8), and temperature affect the photocatalytic performance of the hybrid materials, as was investigated by Karaca et al. [71]. In their study, ZnO nanorods coated with Rheum ribes-derived N-doped CDs were synthesized via hydrothermal method for the photocatalytic degradation of TC. It was reported that the pH of the solution plays a critical role in the photocatalytic oxidation processes, as it could influence both the oxidative agents and the surface properties of the photocatalyst. Recently, in a similar work by Nugroho et al. [72], the performance of ZnO functionalized with coconut water-derived CDs in the photocatalytic degradation of ofloxacin (OFX), as well as reactive red azo dye (RR141), was investigated under the influence of several parameters (e.g., pH, photocatalyst dosage, pollutant concentration). Based on the optimum experimental conditions, the hybrid material achieved a degrading rate of 98% for RR141 and 96% for ofloxacin. Notably, Nugroho et al. [73] proceeded with a later study, investigating the degradation of other pharmaceutical compounds. The ZnO photocatalyst functionalized with Rosa indica petals-derived CDs achieved 99% degradation efficiency of both AMX and CPX in an acidic environment. The role of catalyst dosage and pollutant concentration was also examined, indicating that total degradation of amoxicillin and ciprofloxacin occurs at lower levels of concentration (5–10 ppm) and a higher dosage of catalyst (75 mg), within 40 min. Furthermore, as it was indicated, the photocatalytic activity of the hybrid material remained effective after five cycles. Based on XRD and SEM analysis, there were no significant alterations in the structural and morphological characteristics of the photocatalyst, before and after the photodegradation experiments, indicating its enhanced stability.
Table 1 presents a selection of noteworthy studies published from 2023 up to now, reflecting the most recent developments in the combination of CDs with TiO2 and ZnO, as well as other metal oxides, for effective water treatment via photocatalysis.
Lastly, the broad family of metal oxides includes magnetic materials, such as ferrites (MFe2O4, where M is a divalent metal cation). Ferrites (e.g., Fe3O4, CoFe2O4, CuFe2O4, etc.) have recently drawn the interest of the scientific community for various photocatalytic applications due to their narrow bang gap, strong absorption of visible light, and high chemical and thermal stability, as well as their magnetic properties which make them easily separable from the reaction mixture. Several studies have been devoted to developing ferrites with improved photocatalytic properties through different strategies, such as modification with noble metals, metal doping, and fabrication of heterostructures [88]. Among them, functionalization of ferrites with CDs is one of the most cost-effective and environmentally friendly strategies. Table 2 summarizes the most recent publications (2023–present) on the field of magnetic CD–ferrites heterostructures for photocatalytic water treatment applications.

2.4. CDs–Metal Chalcogenide Heterostructures

Herein, the most recent literature on the combination of CDs with metal chalcogenides for photocatalytic degradation of water pollutants is discussed. Notably, the research in this technological field is focused on metal sulfides and there appears to be lack of publications on the integration of CDs with metal selenides or metal tellurides (in the past three years). Thus, the present section is dedicated to CDs–metal sulfide heterostructures and their enhanced photocatalytic activity. It is interesting to note that a basic search on Scopus using the phrase “carbon dots AND metal sulfides” generates a record of more than 400 papers published during the past decade (Figure 9), covering a wide range of applications, including environmental remediation, biomedicine, sensing, and energy-related processes. To begin with, metal sulfides are compounds in which a sulfur anion is combined with a cationic metal or semimetal to form MxSy with stoichiometric compositions such as MS, M2S, M3S4, and MS2. Based on their elemental composition, metal sulfides are categorized as binary (e.g., CdS, MoS2, In2S3, etc.), ternary (e.g., ZnIn2S4), or multinary [100].
Among numerous binary metal sulfides, cadmium sulfide (CdS) stands out as a highly promising direct bandgap semiconductor. CdS has attracted considerable attention in the field of photocatalysis due to its narrow band gap (2.42 eV) and excellent light harvesting capability. However, its photocatalytic efficiency could be limited by several factors, such as photogenic charge–hole recombination, photo corrosion, and agglomeration phenomena. In order to overcome these issues, various strategies have been proposed, including structural and morphological alterations, incorporation of co-catalysts, surface sensitization, etc. Among them, the most promising is the functionalization of the CdS surface with CDs, which is attributed to the electron-accepting and transport characteristics of CDs [101,102]. In their study, Choudhary et al. [101] investigated the decoration of CdS with CDs prepared by neem leaves for the photocatalytic degradation of the pharmaceutical compound CPX (Figure 10a). More specifically, CDs–CdS hybrid materials containing two different amounts of CDs were synthesized via a hydrothermal method for 150 °C for 6 h. The CPX degradation performance of CDs, CdS, CDs (1 mL)–CdS, and CDs (2 mL)–CdS was observed to be 51%, 68%, 73%, and 74%, respectively (Figure 10b).
Copper sulfide (CuxS) is a p-type semiconducting material with a band gap varying from 1.2 eV to 2.2 eV, based on the percentage of copper [81]. Cheng et al. presented the modification of CuS with Cu-doped CDs and the utilization of the heterostructure as a Fenton-like photocatalyst for the degradation of the antibiotic TC with the assistance of H2O2 [103]. Based on the morphological characterization results, observed in Figure 10c,d, Cu-doped CDs have been successfully incorporated onto CuS nanospheres. The photocatalytic activity of Cu-doped CDs–CuS was determined by the photodegradation of TC. In each experiment, 100 mg of photocatalyst were added into 100 mL of 10 ppm TC aqueous solution, followed by the addition of 1 mL of H2O2 (30%) to initiate the Fenton-like reaction. The TC degradation efficiency was significantly enhanced (by 85%) in a Fenton-like system using the hybrid material (Figure 10e). The proposed mechanism of photocatalytic degradation of the TC is presented in Figure 10f, as well as Equations (10)–(16). More specifically, under illumination, CuS absorbed energy and produced photogenerated e and h+ on its CB. Following this, the photogenerated e transferred to the interface of the Cu-doped CDs–CuS (Equation (10)), which prompted the reduction of Cu2+ to Cu+ (Equation (11)). As a result, Cu+ reacted with O2 to produce active radicals O2· (Equation (12)). In addition, reduction of Cu2+ by H2O2 leads to the generation of Cu+ and O2· (Equation (13)), while Cu+ could be oxidized by H2O2 to form Cu2+ and ·OH (Equation (14)). Electrons play a crucial role in enhancing the efficiency of Fenton-like photocatalytic reactions by accelerating the reduction of Cu2+ to Cu+, thereby facilitating the regeneration of Cu2+ active sites. The photogenerated holes (h+) are driven toward the Cu-doped CDs, where they react with surface hydroxyl groups to form ·OH radicals (Equation (15)). The reactive species O2·, h+, and ·OH generated on the Cu-doped CDs–CdS hybrid material diffuse into the solution, breaking down TC into smaller molecules (Equation (16)) and ultimately mineralizing it into CO2 and H2O. Furthermore, recyclability experiments were conducted for both pristine CuS and Cu-doped CDs–CuS. The results showed that after four cycles, the photocatalytic efficiency was decreased by 35% and 15%, respectively. Notably, XPS analysis of the hybrid material after its utilization in the photocatalytic degradation of TC demonstrated that Cu2+ ions were almost undetectable, indicating minimal leaching during the reaction cycles. Lastly, it is interesting to note that the modification of CdS with Cu-doped CDs resulted in excellent charge separation and photoresponse stability, as illustrated in Figure 10g.
CuS + hv → e + h+
Cu2+ + e → Cu+
Cu+ + O2 → O2·− + Cu2+
Cu2+ + H2O2 → Cu+ + O2·− + 2H+
Cu+ + H2O2 → Cu2+ + ·OH
Cu-doped CDs-OH + h+ → Cu-doped CDs + ·OH
O2/h+/·OH + TC → degradation products
Subsequently, the exceptional physicochemical properties of 2-D nanomaterials, defined by atomically thin layers, have enabled groundbreaking advancements. In this context, 2-D molybdenum disulfide (MoS2) has gained considerable attention and widespread application in various fields. Bulk MoS2 is an indirect-bandgap semiconductor (1.29 eV), which alters to a direct-bandgap semiconductor (1.90 eV) when reduced to monolayers. Although the narrow bandgaps of MoS2 enhance its light absorption capability, the misaligned energy levels hinder the efficient photocatalytic oxidation and reduction processes. Thus, the functionalization of MoS2 with low-cost CDs with up-conversion fluorescent and photogenerated electron-transfer properties has been proposed by several groups.
Qu et al. [104] reported the synthesis of N and S co-doped CDs followed by their incorporation onto MoS2 nanosheets via a hydrothermal method (200 °C, 24 h), as it is illustrated in Figure 11a. In this study, different mass ratios of N and S co-doped CDs to MoS2 were tested to evaluate their effect on the photocatalytic activity. Regardless of the mass ratio, the as-prepared hybrid materials exhibited flower-like morphologies. MB and Malachite Green (MG) were dissolved in water samples collected from Tianjin Lake and photodegradation experiments were carried out following the addition of the heterostructure. As is observed in Figure 11b,c, the increasing N and S co-doped CDs concentration enhanced the photocatalytic degradation of organic dyes. However, at the highest mass ratio of CDs to MoS2, a decline in the photocatalytic performance was observed. This may be attributed to possible aggregation resulting in active-site restriction. Herein, N and S co-doped CDs–MoS2 hybrid material achieved degradation of 99.8% of MB and 95.1% of MG in an actual water system. The recyclability of the hybrid material was also assessed. As it was observed, MB degradation remained high, reaching 88.9% after five cycles. Simultaneously, XRD and SEM characterization of the used photocatalyst confirmed that its crystal structure and morphology remained unchanged, indicating excellent stability. In a similar work, Li et al. [105] synthesized N-doped CDs modified MoS2 nanoflowers through a hydrothermal process, in an increased concentration of CDs and achieved 91.39% of MB dye photocatalytic degradation.
Recently, Zhang et al. [37] reported the utilization of CDs–MoS2 hybrid material as a photo-piezoelectric synergistic catalyst. At this point, it is important to mention that when a piezoelectric material is subjected to external stress or strain, spontaneous polarization for piezocatalysis is caused and may also enhance the separation of photogenerated e and h+. The carriers can migrate to the surface due to the internal electric field of the piezocrystal. In other words, they could easily capture oxygen/water to produce reactive oxygen species (ROS) for pollutant degradation. Thus, the development of a synergistic photo-piezocatalyst shows great potential in enhancing the activation of peroxymonosulfate (PMS) to generate sulfate radicals (SO4·) for pollutant degradation. MoS2 decoration with CDs could enhance its piezoelectric properties, due to the improved asymmetry, promote the electron transfer, and ensure the persistent PMS activation, as CDs possess a large π-electrons conjugated system. A CD–MoS2 photo-piezocatalyst was utilized for the degradation of TC. More specifically, N-doped CDs were successfully attached onto a MoS2 surface via a hydrothermal method (200 °C, 12 h), as it is illustrated by TEM (Figure 11d). Under visible light irradiation (Figure 11e), both MoS2 and N-doped CDs–MoS2, along with PMS, presented increased performance compared to only PMS, reaching TC degradation efficiency of 55.57% and 59.71%, respectively, after 60 min. Interestingly, under the synergistic visible light irradiation and ultrasonic vibration (Figure 11f) for 60 min in the presence of PMS, the N-doped CDs–MoS2 photo-piezocatalyst presented 99.44% TC degradation. The possible mechanism is described by Equations (17)–(25). More specifically, under the synergistic stimuli, the internal electric field of photo-piezocatalyst generates a large number of charges at its surface (Equation (17)). Afterwards, surface charges and h+ attack the O–O bonds of PMS, resulting not only in the formation of SO4· and ·OH (Equation (18)), but also the reduction of O2 to O2· (Equation (19)). In addition, h+ reacts with PMS and H2O forming SO5· (Equation (20)) and ·OH (Equation (21)). The two SO5· molecules form SO4· and 1O2 (Equation (22)) and then SO4· could easily react with H2O to produce ·OH (Equation (23)). Moreover, ·OH and O2· could further react to 1O2 (Equation (24)). The reactive substances degrade TC into inorganic compounds (Equation (25)).
N-doped CDs-MoS2 → e + h+
HSO5 + e → SO4·− + ·OH
O2 + e → O2·−
HSO5 + h+ → SO5·− + H+
H2O + h+ → ·OH + H+
SO5·− + SO5·− → 2SO4·− + 1O2
SO4·− + H2O → H+ + SO42− + ·OH
·OH + O2·− → OH + 1O2
1O2/O2·−/·OH/SO4·− + TC → CO2 + H2O + inorganic
Tungsten disulfide (WS2) is composed of an S–M–S multilayer structure, which contributes to extended light absorption and promotes efficient electron transfer in heterojunctions. Nevertheless, it still suffers from rapid charge carrier recombination, which limits its overall photocatalytic performance. Fatima et al. [106] reported the incorporation of CDs in a WS2–polyaniline (PANI) heterojunction by ultrasonication for 3 h (Figure 12a,b). The photocatalytic performance of WS2, WS2–PANI and CDs–WS2–PANI was evaluated for the degradation of estradiol (EST) and nitrofurantoin (NFT) pharmaceutical pollutants (Figure 12c,d). CDs–WS2–PANI presented the highest degradation efficiency, 96.63% for EST and 95.76% for NFT in 60 min. In contrast, WS2 and WS2/PANI showed degradation of 18.64% and 58.06% for EST and 28.19% and 70.54% for NFT. The electrons present in the CB of WS2 cannot be scavenged due to the lower oxidation potential, which results in rapid recombination of charge carriers. The conducting polymer PANI improves the degradation efficiency by reducing recombination and shifting the CB to favor ROS generation. In addition, CDs further improve the photocatalytic performance by extending light absorption, trapping electrons, and hindering e–h+ recombination. To assess the recyclability of the CDs–WS2–PANI hybrid material, photodegradation experiments were systematically performed under consistent conditions across each cycle. The results demonstrate that the photocatalyst maintained its degradation efficiency, showing no significant decline even after four cycles. Additionally, XRD analysis conducted before and after photocatalysis verified the structural stability of the hybrid material. These findings indicate that the synthesized photocatalyst is effective in the degradation of water pollutants and exhibits excellent stability.
Indium sulfide (In2S3) is a binary chalcogenide semiconductor with bandgap range of 2.0–2.5 eV. Moreover, it presents five different crystalline structures: a-phase defect cubic (α-In2S3), β-phase defect spinel (β-In2S3), γ-phase layered hexagonal (γ-In2S3), ϵ-phase rhombohedra structure (ϵ-In2S3) and Th3P4-phase defect cubic defect structure (Th3P4–In2S3). Due to its interesting photoelectric characteristics, stability, and diverse crystalline structure, In2S3 has become prevalent among researchers. However, the fast recombination of e–h+ pairs, which is attributed to the relatively narrow bandgap, diminishes its photocatalytic efficiency; as little as 10% of the photogenerated electrons are available for photoreduction. Hence, heterojunction fabrication is an effective strategy for enhancing charge separation and redox capability [107]. On this basis, Yao et al. [108] prepared CDs–In2S3 hybrid materials in various mass ratios through a simple one-pot hydrothermal route. Briefly, InCl3·4H2O and thioacetamide were dissolved into distilled water followed by glucose addition. The homogeneous solution was transferred into a Teflon container and heated to 180 °C for 12 h. GQDs were obtained via the polymerization reaction of glucose during the process; therefore, the mass ratio of In2S3 to GQDs was controlled by adjusting the glucose amount. The final hybrids with different mass ratios demonstrated flower-like structures, where GQDs with a diameter of 3 nm are deposited onto In2S3 nanosheets. The photocatalytic activity of both In2S3 and GQDs–In2S3 heterostructures was evaluated by the degradation of MB. Based on the UV–Vis results, GQDs–In2S3 presented enhanced photocatalytic performance compared to pristine In2S3 after 30 min of irradiation. More specifically, In2S3 achieved 24% MB degradation and the GQDs–In2S3 heterostructures reached 72%, 90% and 75% as the GQDs loading was increased, indicating that their concentration may play a key role in photocatalysis. In addition, the photogenerated charge carrier behaviors of In2S3 and GQDs–In2S3 were also evaluated. The results showed that the photocurrent density of the heterostructure is significantly higher than In2S3, which suggests the enhanced charge transfer and separation rate.
Zinc indium sulfide (ZnIn2S4) belongs to the family of ternary chalcogenides and has been recently the subject of various studies in photocatalytic splitting of water, conversion of carbon dioxide, and removal of water pollutants [109]. Similar to other metal sulfides, ZnIn2S4 suffers from fast recombination of photogenerated e–h+, while its CB and VB positions are not always sufficient for driving redox reactions. In the study of Li et al. [110], ZnIn2S4 modified with a N-doped CDs photocatalyst was developed for in situ H2O2 production and further combined with Fe2+ to assemble a photocatalysis-assisted self-Fenton system for degradation of antibiotics, such as levofloxacin (LEV) and AMX. The heterostructure was synthesized via mixing N-doped CDs and ZnIn2S4 precursors under heating at 80 °C for 2 h. The photocurrent intensity of N-doped CDs–ZnIn2S4 (0.34 μA/cm2) is higher than ZnIn2S4 (0.14 μA/cm2) indicating the larger amount of photogenerated carriers. Moreover, as it is discussed, N-doped CDs possess lower electrostatic potential value (4.762 eV) compared to ZnIn2S4 (5.978 eV); therefore, photogenerated electrons tend to migrate from ZnIn2S4 to N-doped CDs, promoting e–h+ separation. The photocatalytic performance of H2O2 production by different samples was firstly evaluated (Figure 13a); the combination of ZnIn2S4 with N-doped CDs dramatically increased the H2O2 production rate. In addition, as the amount of N-doped CDs in the hybrid material was increased (from 100 mg to 250 mg), H2O2 production raised from 139.4 μM to 152.8 μM, respectively. However, further increase of the dosage hinders the efficiency, which may be assigned to the excessive accumulation of N-doped CDs on the active site. Due to the large amounts of H2O2 produced, a photocatalysis–self-Fenton system of Fe2+/N-doped CDs–ZnIn2S4 was conducted for the degradation of different antibiotics under pH = 7 with 1 mg/L Fe2+. Figure 13b,c shows that over 95% of LEV and 81% of AMX were decomposed within 30 min. It is also interesting to note that the recyclability of the N-doped CDs–ZnIn2S4 hybrid material was evaluated through its application in the degradation of LEV over five cycles. The photocatalytic material presented high degradation rate (>90%) towards the antibiotic pollutant even after the last cycle. In addition, no noticeable change in the crystalline structure was observed, confirming the excellent structural stability of the N-doped CDs–ZnIn2S4 hybrid material.

2.5. CDs–Mxene Heterostructures

Mxenes are a new class of 2-D nanomaterials, discovered in 2011 [111], and include transition metal carbides, nitrides, or carbonitrides. They are produced from 3-D MAX starting materials, represented by the general formula M(n+1)AXn, where M is an early transition metal, n = 1–3, A is an A-group element (primarily groups 13 and 14), and X is carbon or/and nitrogen. During chemical etching of MAX phases, A-layers are removed resulting in the formation of the Mxenes. The general formula of Mxenes is Mn+1XnTx, where Tx represents hydroxyl, oxygen, or fluorine termination groups are on their surface. The production of Mxenes involves several chemical and physical methods, including solution-based etching, such as HF or LiF/HCl (Figure 14), and electrochemical etching, and anhydrous etching, such as molten salt etching or halogen etching. Direct synthesis and various intercalation and delamination techniques have been thoroughly examined since their inception [112].
Mxenes have emerged as promising photocatalytic materials due to high metallic conductivity, excellent hydrophilicity, strong molecular adsorption, and efficient charge transfer [113] and can serve as an exceptional platform for the synthesis of monatomic catalysts, demonstrating commendable performance [114] in different applications [115], including biomedicine, energy storage devices (battery, supercapacitor), sensors, catalysts, etc. Although Mxenes demonstrate exceptional properties, they are constrained by drawbacks such as limited surface characteristics, risks for accumulation, and variable surface charges, which may influence their efficacy in water treatment processes [116]. Additionally, even though Mxenes cannot be utilized directly as photocatalysts as they are not semiconducting materials [112], the usage of Mxene-based photocatalysts (oxidized Mxene-precursors and surface modified Mxenes) in environmental remediation and artificial photosynthesis has been on the rise. The notable distinction between the two categories of photocatalysts is that some components in the latter originate from the oxidation of Mxene precursors. Moreover, various photocatalysts and surface reconstructions [117] exert distinct influences on photocatalytic efficacy [118,119]. Consequently, the synthetic procedure of Mxenes and the related catalysts are crucial for material design.
Figure 14. Synthesis and characterization of Ti3C2Tx Mxene nanosheets. (a) Schematic illustrating the synthesis of Ti3C2Tx Mxene nanosheets. (Adopted from [112] with permission). SEM images of exfoliated Ti3AlC2 by fluoride salts show a fully exfoliated grain [120]. (b) LiF with HCl at 50 °C for 48 h. (Adopted from [121] with permission). (c) NaF with HCl at 40 °C for 48 h. (Adopted from [122] with permission).
Figure 14. Synthesis and characterization of Ti3C2Tx Mxene nanosheets. (a) Schematic illustrating the synthesis of Ti3C2Tx Mxene nanosheets. (Adopted from [112] with permission). SEM images of exfoliated Ti3AlC2 by fluoride salts show a fully exfoliated grain [120]. (b) LiF with HCl at 50 °C for 48 h. (Adopted from [121] with permission). (c) NaF with HCl at 40 °C for 48 h. (Adopted from [122] with permission).
Inorganics 13 00286 g014
Different surface engineering methods have been applied, resulting in significant enhancements in the stability of Mxenes in aqueous environments and a reduction in their receptivity to oxidative damage and deterioration. The incorporation of hydrophilic functional groups, such as -OH and -O, improves the resilience and long-term use of Mxenes in the processing of wastewater [123]. The development of Mxenes containing nanocomposites, which integrate Mxenes with polymers, CNTs, CQDs or metal oxides, has resulted in materials exhibiting complementary characteristics not found in their respective parts [124].
Mxene-CDs nanocomposites are currently establishing themselves as a focal point of research, particularly for energy applications. Their remarkable capacity to enhance electron transport and ion migration, along with their specific capacitance, facilitates rapid redox reactions [125]. Guo et al. [126] recently reported the manufacturing of Mxene (Ti3C2Tx), MnO2, and CDs-decorated lamellae through the substitution of CNTs in earlier nanocomposites with CDs. This resulted in an improved capacitance retention rate of the nanocomposite, which is essential for the development of flexible functional devices, including smart wearable electronics. Nguyen et al. [127] have prepared shell structured Mxene@carbon nanodots as highly active bifunctional catalysts both for electrochemical and photoelectrochemical solar-assisted water splitting.
Cao et al. [128] reported on the synergistic combination of a carbon matrix with a 2-D Mxene that offers significant potential for diverse applications of Mxene-based hybrid materials, particularly in areas such as electrochemical energy storage, photo- and electro-catalytic water splitting, electromagnetic wave absorption, and sensing [129]. The multidimensional carbon-based matrix, ranging from zero-dimensional (0-D) to three-dimensional (3-D) structures, facilitates the construction of diverse heterostructure configurations with 2-D Mxenes. This approach enables the optimization of an Mxene’s intrinsic strength, leveraging its layered structure and distinctive properties for various applications.
An interesting method to prepare CQDs and respective doped Mxenes has been reported by Chen and coworkers [130]. They used cationic polystyrene (CPS) covered with Ti3C2Tx–Mxene and pyrolyzed it at temperatures as low as 410 °C for 2 h in a nitrogen atmosphere. During the pyrolysis process, the Ti3C2Tx–Mxene shell served as a barrier, effectively retaining the release of CPS decomposition products and facilitating their penetration into the interlayers of the Mxene (Figure 15). The composite can be used as such but can also act as source for extraction of pure CQDs.
Modified Ti3C2 Mxenes have garnered significant interest due to their distinctive properties and are regarded as highly potential materials for pollutant removal (e.g., through active electro-adsorption processes) and water splitting [131]. Photo(electro)catalytic water decomposition for hydrogen production has become lots of attention in recent publications as Mxene-based and Mxene-derived photocatalysts have significant activity for environmental and renewable energy purposes, including the disintegration of water to generate H2.
Hui et al. [132] have prepared MoIn2S4 and CQDs–MoIn2S4 composite photocatalysts by applying an in situ hydrothermal process, in which citric acid was both the carbon source and surfactant. The optimized CQDs–MoIn2S4 photocatalyst had the maximum photocatalytic efficacy, attaining MB elimination rate of 94.4% after 60 min of simulated solar exposure. The enhanced performance was chiefly ascribed to the up-conversion photoluminescence effect of CQDs, which improved the light absorption of MoIn2S4. The few-layer design of MoIn2S4 as a matrix contained multiple active sites that effectively prevented the clustering of CQDs. Moreover, the CQDs functioned as electron acceptors and transport hubs, mitigating the fast recombination of photogenerated carriers and enhancing electron transmission. Moreover, it has been shown that superoxide radicals (O2·) are the principal active species responsible for photocatalytic destruction.
Other applications of Mxenes based nanocomposites in the wastewater treatment are more focused on their excellent adsorptive properties and thus mostly used as adsorption agents or as separation membranes [133].

2.6. CDs–Metal Oxyhalide Heterostructures

In recent years, growing interest in the combination of metal oxyhalides (MxOyXz, where X represents F, Cl, Br, or I) with CDs for photocatalytic water treatment applications and has been reflected in numerous studies. Among them, Bi-based oxyhalides (BixOyXz) have attracted considerable attention because of their unique layered structure, tunable bandgap, high chemical stability, and low toxicity [134].
BiOCl is one of the most popular Bi-based oxyhalide photocatalyst for water pollutants degradation. However, due to its wide bandgap (3.6 eV), it exhibits limited activity under visible light. In order to improve its performance under irradiation in a wide-spectrum of light, CDs–BiOCl heterostructures have been developed and studied by several groups. The concentration of CDs appears to be a key factor in the photocatalytic performance; thus, various studies focused on optimizing the synthesis process by adjusting the amount of CDs loaded onto BiOCl. For instance, Li et al. [135] prepared CDs–modified 3-D flower-like BiOCl structures (Figure 16a–d) with various CDs concentrations and evaluated their performance in the degradation of RhB dye under visible light. After 30 min, the degradation rates of CDs–BiOCl as the loading of CDs increases are 59.3%, 89.1%, 99% and 95%, while for pristine BiOCl it is 54.3% (Figure 16e). In all cases, CDs enhance the degradation efficiency of BiOCl as they improve light absorption, charge transfer, and charge carrier separation. Although, beyond a certain loading level, the performance slightly declines, which may be attributed to CDs shielding the active sites of BiOCl. Moreover, the photocurrent density of CDs–BiOCl is increased compared to BiOCl, which indicates that the photogenerated carriers are separated efficiently (Figure 16f). The recyclability of the optimized CDs–BiOCl hybrid material was also investigated. It is worth noting that the photocatalyst maintained a degradation efficiency of 95% despite undergoing four cycles. Furthermore, XRD analysis conducted before and after the recycling process confirmed the high structural stability of the CDs–BiOCl photocatalyst. The removal of RhB dye using CDs–BiOCl heterostructures was also conducted by other research groups. Similarly, in the study of Shi et al. [136], the photocatalytic performance of the hybrid materials decorated with different amounts of GQDs was evaluated and the results showed that under visible-light irradiation the degradation efficiency of BiOCl was enhanced as the GQDs loading increased. (Figure 16g). Wang et al. [137] presented the photocatalytic performance of the optimized Bi-doped CDs–BiOCl sample under 190–1100 nm broadband light irradiation, which exhibited 2.3 times higher RhB degradation than BiOCl. Additionally, other organic compounds, such as tetracycline hydrochloride (TCH) and bisphenol A (BPA), have been successfully degraded using CDs–BiOCl photocatalysts. Lu et al. [138] presented an in-depth study about synergistic role of CDs and Ovs in the enhancement of BiOCl photocatalytic properties (Figure 16h). The hybrid materials were prepared via a high-temperature hydrothermal process, during which CDs are combined with BiOCl via C=O bonding. ESR spectroscopy was employed to verify that CDs promote the formation of additional Ovs within the material. As shown in the results (Figure 16i), a significant signal with a g-value of 2.003 is detected in pristine and modified BiOCl. Notably, the signal intensity of CDs–BiOCl is stronger, and combined with XPS analysis, these results indicated that CDs enhances Ovs. As a result, the photocatalytic performance of CDs in the degradation of TCH under visible light irradiation was highly improved from 19.48% to 79.50%. More specifically, the interactions of BiOCl with CDs via C=O bonding, leads to the increase in the number of Ovs, caused by local lattice defects and dislocations, which ultimately create new energy levels between the CB and VB bands of BiOCl. The Ovs serve as electron traps, promoting the capture of light-induced electrons at their respective energy levels; trapped electrons react with O2 to form ·O2, while others are transferred to the CDs. Simultaneously, due to their up-conversion characteristics, the CDs extend absorption into visible light range, which results in the generation of more e–h+ pairs. Therefore, CDs–BiOCl combination is beneficial for the degradation of TCH. Among various CDs–BiOCl hybrid materials prepared with different amounts of CDs and tested for their photocatalytic performance in TCH degradation, the optimum photocatalyst was further investigated in terms of recyclability and structural stability. Based on the results, the hybrid material demonstrated excellent photocatalytic stability, maintaining over 91% TCH degradation efficiency after five cycles, with XRD analysis confirming its structural integrity. In a similar work [139], the synergistic role of functionalization with CDs and heteroatom doping of BiOCl was investigated. N-doping could substitute Cl in the lattice structure of BiOCl or attach to Bi, forming an intermediate energy level and shorten the electron transfer distance. Nevertheless, the fast in situ recombination of charge carriers hinders their photocatalytic properties. Inspired by this issue, they modified N-doped BiOCl with CDs through a calcination process. Based on the EIS Nyquist plots (Figure 16j), CDs–N-doped BiOCl presented the smallest curve radius indicating that both incorporation of CDs and N-doping reduce the electron migration energy barrier, thereby enhancing interfacial charge transfer within the photocatalyst. The photocatalytic performance of BiOCl, N-doped BiOCl, and CD–N-doped BiOCl in the degradation of BPA was 31.88%, 48.39%, and 90.65% under visible light irradiation for 20 min, respectively.
The enhancement of photocatalytic activity In other metal oxyhalides through functionalization with CDs has also been explored. In a recent study [140], CDs–Cu2Cl(OH)3 heterostructures were synthesized via hydrothermal treatment, using different concentrations of CDs, for the photocatalytic degradation of MB dye from aqueous environment. Based on the UV–Vis spectra, Cu2Cl(OH)3 exhibits a broad but low absorbance at 200–500 nm. After modification with CDs, the absorbance is highly increased until overload and a blue-shift at 300–400 nm is also observed. (Figure 16k) The degradation of MB dye using Cu2Cl(OH)3 and the optimized CDs–Cu2Cl(OH)3 heterostructure was 37.40% and 88.20%, respectively, suggesting that CDs could greatly enhance the photocatalytic efficiency, with significant cycling stability (Figure 16l,m).

2.7. CDs–Metal-Free Heterostructures

During the past few decades, metal-free semiconductors have attracted significant attention as a green alternative class of photocatalytic materials due to their appealing characteristics and properties, making them suitable for various photocatalytic reactions [141]. Among them, one of the most promising candidates is the 2-D graphitic carbon nitride (g–C3N4), due to its environmentally friendly nature and facile and low-cost synthesis procedure, as well as chemical stability. Although g–C3N4 possesses a moderate bandgap (~2.7 eV) and favorable band edge positions, further improvements are still necessary to enhance its photocatalytic performance. Thus, researchers continue to investigate several strategies in order to optimize its capability, including the increase of specific surface area via morphology modulation [142], tuning the energy band structure via heteroatom doping [143], and enhancing charge carrier separation through the development of nanoarchitectures and interfacial heterostructures [144,145]. The combination of g–C3N4 with CDs has been considered as an appealing strategy for the development of metal-free photocatalysts with enhanced properties and eco-friendly characteristics. CDs could function as a photosensitizer, improving the visible light absorption of g–C3N4 through their up-converted photoluminescence properties [146,147], while also stimulating g–C3N4 to generate electron–hole pairs and produce more reactive O2· and ·OH radicals, ultimately enhancing its photocatalytic ability. Additionally, CDs could serve as effective electron acceptors by capturing photoinduced electrons, thereby minimizing their recombination with holes in the valence band of g–C3N4. Herein, the most recent research efforts in the utilization of CDs–g–C3N4 binary systems in the degradation of water pollutants are presented.
To begin with, CD–g–C3N4 heterostructures have been successfully utilized in the removal of organic compounds, such as dyes and pharmaceuticals, from aqueous environment. As it is indicated in the study of Jourshabani et al. [148], N-doped CQDs, playing the role of electron acceptors, were anchored onto the surface of g–C3N4 nanosheets for RhB degradation. Briefly, g–C3N4 was synthesized via a two-step process, including hydrothermal treatment of cyanuric acid and a melamine mixture, followed by calcination, while N-doped CQDs were prepared by a micro-plasma technique. The above components were combined through ultrasonication, forming the final N-doped CQDs–g–C3N4 heterostructure. The effect of the N-doped CQDs concentration in the photocatalytic performance of the hybrid material was evaluated: 0.2 g of g–C3N4 were mixed with 1.0 mg, 2.0 mg and 3.0 mg of N-doped CQDs, respectively. Following this, RhB samples containing pristine g–C3N4 and N-doped CQDs–g–C3N4 hybrid materials, were exposed to simulated solar light for 2.5 h. Based on the results, g–C3N4 presented 65% degradation efficiency, while the heterostructure containing moderate amount of N-doped CQDs achieved a degradation efficiency of 91%. Further increase of N-doped CQDs in the heterostructure results in the decline of its photocatalytic performance, which may be attributed to the reduction of available active sites. In a similar study [149], N-doped GQDs–g–C3N4 hybrid material, which was successfully prepared via stirring for 12 h at room temperature (Figure 17a), presented excellent performance regarding the light-absorbing capacity, charge carrier separation, and visible-light driving. As a result, it degraded 91.2% of RhB within 90 min (Figure 17b) with a degradation rate 1.47 times higher than that of pristine g–C3N4 (Figure 17c). It is worth mentioning that the hybrid material retained its activity over 10 consecutive cycles with no significant decline in performance, highlighting its excellent photocatalytic stability. The structural stability of the N-doped GQD–g–C3N4 was also confirmed via XRD analysis of the recycled sample. In addition, Ishak et al. [150] presented the functionalization of g–C3N4 with microwave-synthesized CQDs for the efficient degradation of MB dye. In their noteworthy study, they address a critical issue that has attracted considerable attention from the scientific community: the presence of by-products in CDs synthesis. It is well known that during bottom-up synthesis of CDs, including methods such as microwave irradiation, a wide variety of by-products, agglomerates, or unreacted species may be formed. These components could potentially affect the photocatalytic activity of the resulting materials. To investigate this, the researchers isolated different fractions using column chromatography (Figure 17d). The first fraction contains 2-pyridone-based by-products, which are highly fluorescent molecules. These fluorophores were combined with g–C3N4 forming a hybrid material. Similarly, fluorophores-free CDs and as-prepared CDs were also attached onto g–C3N4 via a facile route involving 4 h of mixing, followed by 1 h of ultrasonication. The photocatalytic experiments showed that the best MB degradation performance was achieved by the purified CDs–g–C3N4 heterostructure, followed by as-prepared CDs–g–C3N4 > g–C3N4 > fluorophores-g–C3N4, indicating that by-products during CDs synthesis decrease the heterostructure’s photocatalytic activity.
In another study conducted by Awang et al. [151], the photocatalytic degradation of diclofenac by N-doped CQDs–g–C3N4 was evaluated. More specifically, the experiments were carried out for 180 min under visible light irradiation and the results showed that the hybrid material achieved higher photocatalytic efficiency (62%) in comparison with the bulk g–C3N4 (50%). N-doped CQDs improved visible light absorption, as well as charge carrier separation. Nevertheless, an excessive amount of N-doped CQDs could lead to a shield effect and hinder photocatalysis, which has been addressed in various publications. Recently, Yang et al. [152] investigated the crucial role of CDs surface functionality in the development of an effective CDs–g–C3N4 photocatalyst. Therefore, different CDs samples were produced, nominated as 1 N-CDs, 2 N-CDs and ca-CDs, which were assigned to CDs rich in pyrrolic N, graphitic N, and carboxyl groups (N-free), respectively (Figure 18). CDs were physically integrated with g–C3N4 to preserve their chemical functionalities. The efficiency of CDs–g–C3N4-catalyzed PMS in p-nitrophenol (PNP) degradation was further studied. The findings of this work prove that functional groups matter, as the pyrrolic-N and carboxyl groups promote electron trapping, improving charge separation and enhancing photocatalytic activity. On the other hand, the graphitic-N caused electron-rich regions that hinder charge separation, resulting in the reduction of the effectiveness. In particular, the 1 N-CDs–g–C3N4 photocatalyst, showed the greatest PNP degradation (95.9%), with the best charge separation and ROS generation. Furthermore, ca-CDs–g–C3N4 presented satisfactory performance (68.7%) due to surface carboxyl groups, whereas 2 N-CDs–g–C3N4 demonstrated poor performance (21.1%) due to electron accumulation barriers. The roles of reactive species involved in the PNP degradation process followed the order h+ > 1O2 > ·OH > O2·, while SO4· played a minor role due to quenching by carboxyl groups. The possible reaction routes of photocatalytic processes have been described in Equations (26)–(31):
HSO5 + e → ⋅OH + SO42−
O2 + e → O2
O2·− + 2 H2O → 1O2 + H2O2 + 2 OH
O2·− + H+ → H2O2 + 1O2
H2O2 + e → ⋅OH + OH
h+/1O2/·OH/O2·− + PNP → intermediates → CO2 + H2O
The CDs–g–C3N4 heterostructure has also been applied in water disinfection from bacteria. For instance, Yang et al. [153] synthesized a CDs–g–C3N4 photocatalyst through an electrostatic self-assembly method for the effective disinfection of water from methicillin-resistant Staphylococcus aureus (MRSA) (Figure 19a). In this work, O-doped g–C3N4 was prepared via hydrothermal treatment of bulk g–C3N4 in the presence of H2O2. The aim of O-doping was to hinder the recombination of photogenerated hole–electron pairs and promote the exposure of additional active centers. Simultaneously, CDs were synthesized via hydrothermal treatment of 2,5-diamino-phenyl sulfonic acid. The loading of CDs onto O-doped g–C3N4 enhances the generation of photogenerated electron–hole pairs by broadening the visible light absorption range and facilitates the separation of charge carriers through effective trapping of photogenerated electrons. Moreover, CDs incorporation alters the surface charge of the hybrid photocatalytic material from negative to positive, thereby improving the interaction between active species and bacterial cells. As it is presented in Figure 19b, the CDs–O-doped g–C3N4 heterostructure demonstrated a significantly enhanced ability to inactivate MRSA, achieving a 4.08-log reduction of cell density compared to 0.46-log and 1.13-log reduction of bulk g–C3N4 and O-doped g–C3N4, respectively. In addition, fluorescence-based live/dead staining assays were performed to verify the damaging effect of CDs–O-doped g–C3N4 towards MRSA (Figure 19c). In these assays, SYTO 9 dye penetrated the intact cell membrane and labelled live bacteria with green fluorescence, whereas PI dye selectively entered the damaged cell membrane and caused the dead bacteria to emit red fluorescence. As it was observed, the majority of bacteria exhibited red fluorescence, suggesting disruption of the cell membrane wall in the presence of the photocatalyst. The superior antibacterial performance of the heterostructure is primarily due to improved photoelectric conversion efficiency and modified surface charge characteristics. CDs not only extend the visible light absorption range but also facilitate the separation of photogenerated electron–hole pairs by acting as electron traps. Additionally, the positive surface charge of O-doped g–C3N4, after CDs incorporation, improves the electrostatic interactions among the photocatalyst and the bacteria, thereby enhancing the likelihood of active species coming into contact with bacterial cells. Reactive oxygen species, particularly superoxide radicals (O2·) and photogenerated holes, were identified as the dominant agents responsible for disinfection. These species compromise bacterial cells by damaging membranes, causing intracellular leakage, and ultimately leading to cell death.
The removal of heavy metal ions from water using CDs–g–C3N4 has been recently studied by Xu et al. [154]. Licorice was utilized as a carbon source for the green in situ formation of CQDs onto a g–C3N4 surface via a hydrothermal route. The as-prepared hybrid materials, containing various mass ratios of licorice:g–C3N4 were evaluated in the Cr (VI) photocatalytic reduction. Notably, after 20 min of visible light irradiation, the removal efficiency of Cr (VI) achieved by pristine g–C3N4 and the optimum sample of CQDs–g–C3N4 was 71.3% and 99.5%, respectively (Figure 19d). XRD analysis confirmed that the CQDs–g–C3N4 photocatalyst retains its original crystal structure after the Cr (VI) reduction, with no discernible changes compared to the initial sample. In addition, based on the pseudo-first-order kinetic model, the rate constant of hybrid material was 2.48 times higher than that of g–C3N4 (Figure 19e). It is also important to note that the main reactive species involved in Cr (VI) reduction were investigated via trapping experiments using benzoquinone, ammonium oxalate, and potassium bromate for the selective capture of superoxide radicals, photogenerated holes, and electrons, respectively. As it was noticed, the addition of benzoquinone had no significant effect in Cr (VI) removal. On the other hand, ammonium oxalate enhanced reduction, which could be ascribed to the fact that it traps holes, promoting the separation of photogenerated carriers and thereby producing more photogenerated electrons. In contrast, the addition of potassium bromate significantly suppressed Cr (VI) removal. Therefore, the results strongly indicate that the photogenerated electrons are the most crucial reactive species in the photocatalytic degradation of Cr (VI).
In addition, there are various noteworthy publications, from 2023 to the present, which address the photocatalytic degradation of water pollutants utilizing CDs–graphitic C3N4 heterostructures. A summary of these studies is provided in Table 3.

2.8. CDs-Based Ternary Heterostructures

Up to this point, the significance of hybrid materials has been well recognized. The growing interest reflects the potential of such materials to exhibit enhanced properties through the combination of distinct components. Concurrently, progress in materials science has led to the development of novel structures, often involving more than two constituents. These advanced multi-component systems demonstrate improved and synergistic functionalities, highlighting the evolving capabilities of modern material design. Therefore, the final section of this review presents CDs-based ternary heterostructures employed in the photocatalytic degradation of water pollutants.
For instance, in the recent study of Guan et al. [163], two different metal oxides were combined with CQDs, forming a CQDs–ZnO-TiO2 ternary heterostructure. The novel photocatalytic material was utilized for the effective degradation of MB dye. Although ZnO is a promising material for photocatalytic applications, its efficiency is limited because of the rapid electron–hole recombination. On the other hand, TiO2, as a n-type semiconductor, features numerous surface oxygen vacancies and highly polar Ti-O bonds, which facilitate water molecule dissociation and hydroxyl groups formation on its surface. As a result, combining ZnO with TiO2 is an effective modification strategy to enhance photocatalytic degradation performance and hydrogen production. In order to improve the visible light utilization of ZnO–TiO2 hybrid material, its functionalization with CQDs is an effective strategy. As it is presented in their work, the CQDs–ZnO–TiO2 heterostructure achieved a 96.70% MB degradation efficiency. In another study, Hao et al. [164] combined Cu2O and TiO2 metal oxides with N-doped CDs using hydrothermal, as well as chemical precipitation, processes. The as-prepared N-doped CDs–TiO2–Cu2O ternary heterojunction was tested in the photocatalytic degradation of MO. Based on the results, the photocatalyst exhibited a 99% MO degradation efficiency within 20 min.
Moreover, metal oxide–metal sulfide heterojunctions combined with CDs have emerged in recent years as a subject of research due to their promising photocatalytic properties. A dendritic TiO2–CdS heterostructure modified with CQDs was presented by Dou et al. [165]. The as-synthesized CQDs–TiO2–CdS photocatalyst was utilized for RhB degradation. More specifically, the TiO2–CdS hybrid material was prepared by a hydrothermal method, followed by CDs impregnation via soaking at room temperature (Figure 20a). Based on the photocatalytic experiments, it was found that CDs improved the efficiency of dendritic TiO2–CdS, as the RhB degradation was increased from 69.00% to 84.50% within 50 min of irradiation. A CDs–TiO2 heterojunction decorated with CDs lead to the increase of the absorption range in the visible region and improved the photo-generated electron–hole separation and carrier transfer rate when electrons are transferred from TiO2 to CdS. As is observed in Figure 20b, photogenerated electrons in the CB of TiO2 will recombine with holes in VB of CdS, resulting into the accumulation of electrons in CdS VB. Simultaneously, CDs act as electron trapping agents, preventing photogenerated electrons from recombination with holes. Under light, the excess holes in the VB of TiO2 can participate in the oxidation reaction of RhB dye, which makes the photocatalyst having excellent degradation activity. Recently, Shi et al. [166] developed a CDs-decorated ZnFe2O4–ZnIn2S4 novel material and investigated the effect of CDs position (Figure 20c) on the charge transfer of the heterojunction. More specifically, two cases were studied, in which CDs were located either on the exterior or within the interior of the ZnFe2O4–ZnIn2S4 core–shell structure. The experimental results in the photocatalytic degradation of tetracycline revealed that when CDs were located at the contact interface of core–shell material (interior), the heterojunction presented enhanced performance. This indicates that CDs act as a high-speed and smooth pathway for electron migration, effectively facilitating the separation of photo-excited electrons and holes. Since the photocatalyst’s stability is a crucial factor for its practical application, CDs-decorated ZnFe2O4–ZnIn2S4 was further characterized by XRD and SEM analysis after undergoing multiple photocatalytic cycles. The results verified that both its crystal structure and morphology remained unchanged throughout the reaction process.
Among various mediator candidates, metals (e.g., Ag, Au, Pt, and Pd) are often used in ternary heterojunctions due to their high ability to form Schottky junctions and charge transfer carriers. Wang et al. [167] reported the preparation of a g–C3N4–N-doped CDs–Ag ternary structure and its application for organic pollution degradation in aqueous solutions. More specifically, the g–C3N4–N-doped CDs–Ag photocatalysts were constructed using an interfacial electronic control mechanism with double doping of N–CDs and Ag NPs. In that case, the N-doped CDs act as a bridge for electrons, guiding the flow of photogenerated charge carriers. The Ag NPs act as photosensitizers and electron acceptors, enhancing the electron transport in the material. The cooperative effect between N-doped CDs and Ag NPs could broaden the absorption range of the visible spectrum of the g–C3N4–N-doped CDs–Ag catalyst, enhance the excitation of electrons, and reduce the photogenerated electron–hole pairs recombination rate. The experimental results showed that the degradation of several contaminants (MO, RhB, TCH, and chrysin hydrochloride (CHC)) can be achieved within 100 min using a 300 W xenon lamp as the light source (λ > 400 nm). The reaction rates of g–C3N4–N-doped CDs–Ag photocatalysts for the degradation of MO and RhB were 13.5 and 12.9 times higher than those of the conventional g–C3N4, respectively. It is also important to mention that this study revealed that the g–C3N4–N-doped CDs–Ag heterostructures remained stable after six cycles of use and the degradation rate continued to be higher than 96%.
Ternary heterostructures based on CDs, Bi-based oxyhalides, and g–C3N4 components have also been investigated in water treatment applications. Interestingly, Bi-based oxyhalides exhibit lower bandgap as the atomic number of the halogen increases, which enhances their ability to absorb visible light. This property enables the strategic combination of different halogens in Bi-based oxyhalides structures to tailor their fundamental properties and enhance their photocatalytic efficiency. Recently, Abdurahman et al. [168] investigated the combination of g–C3N4 decorated with CQDs and BiOClxBr1-x fabricated by calcination and hydrothermal methods. CQDs act as the electron reservoir, which could delay the photogenerated electron–hole pairs’ recombination rate, thus improving the overall photocatalytic performance. The obtained g–C3N4–CQDs–BiOClxBr1-x was examined in the photodegradation of tetracycline under visible light illumination, along with the effects of the CQDs content, halogen loading, and pH value. The morphology characterization of the synthesized photocatalyst revealed that CQDs and BiOClxBr1-x solid solutions were deposited onto a g–C3N4 surface. Based on the results, it was observed that synergistic effect of 1 wt% CQDs and BiOCl0.75Br0.25 significantly improved the interfacial charge transfer efficiency and light harvesting capacity of the hybrid material. The degradation rate of tetracycline (TC) over g–C3N4–CQDs–BiOCl0.75Br0.25 was 83.4% after 30 min and favorable stability with near-initial capacity under visible light irradiation. In this context, Zhu et al. [169] prepared a series of Bi7O9I3–g–C3N4 heterostructures (Figure 21a) modified by different contents of lignin-derived CQDs (0.2 wt%, 0.5 wt% and 1 wt%) for synchronous photocatalytic reduction of Cr (VI) and levofloxacin (LEV) degradation, under simulated light irradiation, using a 300 W Xe lamp. As is observed in Figure 21b, pristine CQDs, Bi7O9I3, and g–C3N4 photocatalysts exhibited limited Cr (VI) reduction efficiencies of 4.5%, 51.2%, and 36.8%, respectively, which may be due to the intensive photon scattering and rapid charge recombination. On the other hand, the binary structures of CQDs–Bi7O9I3, CQDs–g–C3N4, and Bi7O9I3/g–C3N4 presented higher Cr (VI) reduction performance of 64.1%, 41.6%, and 74.4%, due to the improved charge separation efficiency. Notably, the development of ternary CQDs–Bi7O9I3–g–C3N4 heterojunctions could further enhance the photocatalytic performance. More specifically, the ternary heterojunction containing 0.5 wt% of CQDs demonstrates optimal performance through broadened UV–Vis light absorption, highly efficient charge carrier separation, as well as, enhanced redox properties. As a result, up to 100% of Cr (VI) photoreduction efficiency was achieved under 60 min of light irradiation, while its reaction rate (0.08725 min−1) was about 4.8 times higher than that of Bi7O9I3-g–C3N4 (0.01825 min−1). Similarly, in Figure 21c the performance of the same photocatalysts in LEV degradation can be observed. CQDs-decorated Bi7O9I3-g–C3N4 reaches 94.8% after 60 min, where the reaction rate (0.03863 min−1) is 2.2 times higher than that of pristine Bi7O9I3-g–C3N4 (0.0174 min−1). Consequently, CQDs sandwiched with a Bi7O9I3-g–C3N4 heterogeneous interface act as an electron reservoir, facilitating the efficient separation of photogenerated electron–hole pairs and accelerating the charge transfer.
The list of CDs-based ternary heterostructures is exceptionally extensive due to the vast variety of available materials and the numerous possible combinations among them. By integrating CDs with various binary systems, researchers have been able to significantly enhance photocatalytic performance, particularly in terms of light absorption and charge carrier mobility, resulting in the development of next-generation photocatalysts. Given this rapidly expanding research topic, Table 4 provides a concise summary of the most recent and noteworthy advances regarding the utilization of novel CDs-based ternary heterostructures in the field of photocatalytic water treatment.
Figure 21. (a) Schematic illustration of the preparation of CQDs–Bi7O9I3–g–C3N4 hybrid material. Photoreduction of (b) Cr (VI) and (c) LEV by pristine CQDs, Bi7O9I3, g–C3N4, binary heterostructures Bi7O9I3–g–C3N4, CQDs–Bi7O9I3, CQDs–g–C3N4, and ternary heterostructures CQDs–Bi7O9I3–g–C3N4 containing different CQDs concentrations. (Adopted from [169] with permission).
Figure 21. (a) Schematic illustration of the preparation of CQDs–Bi7O9I3–g–C3N4 hybrid material. Photoreduction of (b) Cr (VI) and (c) LEV by pristine CQDs, Bi7O9I3, g–C3N4, binary heterostructures Bi7O9I3–g–C3N4, CQDs–Bi7O9I3, CQDs–g–C3N4, and ternary heterostructures CQDs–Bi7O9I3–g–C3N4 containing different CQDs concentrations. (Adopted from [169] with permission).
Inorganics 13 00286 g021
Table 4. Recent advances in photocatalytic water treatment using CDs-based ternary heterostructures.
Table 4. Recent advances in photocatalytic water treatment using CDs-based ternary heterostructures.
CD-Based Ternary HeterostructuresWater PollutantLight SourcePhotocatalytic DegradationReferences
N-doped GQDs–TiO2-Graphene OxideMethylene Blue
(10 ppm)
Crystal Violet
(10 ppm)
Basic Red 46
(10 ppm)
300 W Xenon lamp
>420 nm
96.60% after 150 min
82.40% after 150 min
>99.99% after 150 min
(50 mg/100 mL photocatalyst)
[170]
GQDs–ZnO–NiOMethylene Blue
(30 ppm)
Methyl Orange
(30 ppm)
1000 W Halogen lamp
UV and Vis
93.42% after 90 min
75.68% after 90 min
(100 mg photocatalyst)
[171]
CQDs–CdS–Ta3N5Levofloxacin
(15 ppm)
300 W Xenon lamp
>420 nm
91.90% after 60 min
(25 mg/100 mL photocatalyst)
[172]
CDs–CdS–g–C3N4Tetracycline
(20 ppm)
98.00% after 45 min
(40 mg/50 mL photocatalyst)
[173]
CDs–WO3–g–C3N4Malachite Green
(20 ppm)
Vis96.30% after 80 min
(50 mg/100 mL photocatalyst)
[174]
N-doped CDs–Co3O4–MoS2RhB
(30 ppm)
*97.75% within 5 min
(200 mg/L photocatalyst)
[175]
CQDs–Ag–MoS2Tartrazine
(20 ppm)
Vis>99.9% after 30 min
(300 mg/L photocatalyst)
[176]
CQDs–BiFeO3–BiOBrImidacloprid
(10 ppm)
500 W Xenon lamp
>400 nm
95.7% after 180 min[177]
CQDs–BiOBr–g–C3N4Tetracycline
(10 ppm)
500 W Xenon lamp
Vis
>99.9% after 60 min
(40 mg/40 mL photocatalyst)
[178]
CQDs–BiOBr-Ti3C2Moxifloxacin
(10 ppm)
500 W Xenon lamp
>420 nm
96.10% after 120 min
(25 mg/50 mL photocatalyst)
[179]
CQDs–BiOBr–W18O49Tetracycline
(20 ppm)
300 W Xenon lamp 97.30% after 45 min
(20 mg/50 mL photocatalyst)
[180]
CQDs–Bi2MoO6–CuS300 W Xenon lamp
>420 nm
96.98% after 60 min
(20 mg/100 mL photocatalyst)
[181]
N-doped GQDs–TiO2–g–C3N4Ciprofloxacin
(10 ppm)
>420 nm89.60% after 150 min
(50 mg/100 mL photocatalyst)
[182]
F-doped CDs–TiO2–g–C3N4RhB
(3 ppm)
500 W Xenon lamp
Vis
74.00% after 50 min
(10 mg/100 mL photocatalyst)
[183]
P-doped GQD–TiO2–AgIMethyl Orange
(10 ppm)
300 W Xenon lamp
>420 nm
78.20% after 60 min
(100 mg/100 mL photocatalyst)
[184]
N-doped CD–CuFe2O4–g–C3N4Tetracycline Hydrochloride
(10 ppm)
300 W Xenon lamp85.69% after 60 min
(certain mass of photocatalyst in 100 mL of water pollutant)
[185]
* not mentioned.

2.9. From Conventional Photocatalysts to Novel Heterostructures: The Added Value of CDs

There is no doubt that the modern world has shifted its focus towards hybrid materials due to their advanced properties, which make them attractive for a wide range of technological applications.
Herein, the combination of conventional photocatalytic materials with CDs, forming either binary or ternary heterostructures with enhanced degradation efficiency of water pollutants was investigated. According to the most recent literature, CDs-based hybrid materials exhibit significantly improved photocatalytic performance compared to common inorganic semiconductors. As researchers indicate, the novel heterostructures have demonstrated superior efficiency in the removal of organic dyes, pharmaceuticals, or heavy metal ions from water, often achieving removal rates exceeding 90–95% within a short irradiation time. These enhancements are generally attributed to the unique characteristics of CDs, including the up-conversion efficiency and the ability to act as electron donors or acceptors. Thus, the final hybrid structures present enhanced light absorption (from UV to Vis) and decreased charge carrier recombination, respectively. Moreover, the abundant and tunable surface functionalities of CDs could enhance the active sites of the photocatalyst towards specific pollutants. In addition, CDs integration decreases agglomeration phenomena and plausible toxicity.
The enhancement in the overall photocatalytic performance of the hybrid materials also pave the way for more sustainable and efficient applications in various fields beyond environmental remediation, including hydrogen production, solar energy conversion, sensors, self-cleaning coatings, etc. As research continues to explore novel CDs-based heterostructures and study their physicochemical properties, their role in the development of next-generation functional nanomaterials becomes increasingly indispensable.

3. Outlook

The growing burden of water pollution—driven by urbanization, industrialization, and climate-induced stressors—presents a formidable global challenge. Photocatalysis has already proved its potential to degrade water pollutants via a sustainable and efficient approach. However, the transition from laboratory-scale innovations to real-world impact remains complex. Based on this view, there is an urgent need to develop next-generation photocatalysts with efficiency upon visible light irradiation, robustness, scalability, and diversity to several environmental conditions and pollutants. Conventional photocatalysts, including metal oxides, metal chalcogenides, metal oxyhalides, etc., suffer from various inherent limitations that hinder their practical application in water remediation.
In particular, most of the traditional photocatalysts are activated only upon UV light, impeding their solar utilization efficiency. High rates of electron–hole recombination also constitute a common issue, leading to reduced photocatalytic performance. Photochemical instability and limited water solubility have been reported in the use of photocatalysts, including metal chalcogenides. Among other factors, surface chemistry has been proven to be a crucial parameter for photocatalysis. Materials like MXenes underline the importance of surface charges and features in water treatment procedures. Apart from surface functionalization, photocatalyst synthesis routes should be taken into account given the fact that some synthetic protocols possess high demands on temperature or energy. These challenges collectively underscore the need for novel photocatalytic materials with improved light-harvesting capabilities, superior charge separation, and environmental compatibility, which are characteristics that emerging CD-based systems are increasingly being designed to address.
The development of heterostructures of inorganic photocatalysts with CDs has emerged as a promising strategy to address these intrinsic limitations. CDs provide a unique combination of superb physicochemical properties and cost- and environmental-efficient synthesis and functionalization. In particular, CDs can be synthesized with a variety of techniques and precursors offering multifunctional characteristics. Up-conversion photoluminescence ability and CDs’ dual role either as electron acceptors or donors can provide new perspective on their photocatalytic performance. CDs surfaces are abundant in various functional groups, such as -COOH, -OH and -NH2, allowing the interactions with other semiconductor materials. The higher surface area of CDs, along with their porosity, provided abundant active surface for the adsorption of pollutants and thus increased photocatalytic efficiency. These features encourage the fabrication of CD-based heterostructures.
Given the fact that there is a rising demand for sustainable water treatment approaches, there has been a marked increase in research focused on CD-based photocatalytic heterostructures in recent years. Particularly, studies published from 2023 onward have demonstrated notable progress in the rational design, synthesis, and application of these hybrid materials for the degradation of common pollutants. Researchers are now exploring a wide range of inorganic hosts combined with CDs to construct tailored nanocomposites with enhanced photocatalytic efficiency and stability under visible light. These inorganic semiconductors, ranging from classic metal oxides (e.g., ZnO, TiO2, ferrites) and metal sulfides (e.g., CuS, MoS2, WS2, In2S3, ZnIn2S4) to more novel systems, including MXenes and metal oxyhalides (e.g., BiOCl, Cu2Cl(OH)3), can serve as versatile platforms due to their tunable band structures, unique morphologies, and varying catalytic properties. In addition, CDs-based ternary heterostructures have gathered research interest aiming to exploit the synergistic effect among the components.
For instance, CDs have been successfully incorporated as co-catalysts or sensitizers into various metal oxide matrices. Among them, TiO2 is a popular option in photocatalytic research and its modification with CDs further expands its application. Especially, functionalization of TiO2 with CDs derived from natural sources provides a sustainable route to expand visible-light activity and improve charge separation. Future efforts should prioritize the rational design of these heterostructures with tailored morphologies, surface functionalities, and dopant engineering to further optimize efficiency in water treatment performance. Similar to TiO2, ZnO photocatalysts have been modified with CDs overcoming common limitations of ZnO’s photocatalytic application, including wide bang gap and fast recombination of excitons. The tuning of physicochemical properties of CDs–ZnO photocatalysts could secure improved performance in the degradation of pollutants in complex water matrices. Moreover, CDs–ferrite heterostructures combine magnetic properties and improved photocatalytic action in the visible-light spectrum, offering a practical and sustainable approach for water remediation. Moving forward, the fabrication of this type of heterostructure, with optimum surface chemistry, magnetic recyclability and optical properties, holds strong promise for scalable and selective pollutant removal.
Metal chalcogenides, especially transition metal sulfides (e.g., MoS2, CdS, ZnS, CuS), have aroused research interest in photocatalysis owing to their narrow band gaps, strong light-harvesting capability, and tunable electronic properties. However, their practical application is often hindered by rapid recombination of photogenerated charge carriers and photo-corrosion, particularly in the case of CdS under visible light. This review already underscored the potential of metal chalcogenides photocatalysts integrated CDs to address these limitations and provide new insights in water treatment. The addition of CDs can expand absorption properties to the visible region, as well as facilitate the separation of photogenerated electron–hole pairs, thereby suppressing recombination and improving photocatalytic efficiency. Moreover, the surface functional groups on CDs can form strong interfacial interactions with the metal sulfide matrix, promoting stable hybrid structures.
For instance, CDs addition to MoS2 has been shown to increase the specific surface area and number of active sites while accelerating charge transfer at the interface, leading to significantly improved degradation rates of organic pollutants such as methylene blue and tetracycline. Similarly, CDs–CdS heterostructures showed superior photocatalytic activity under visible light, where CDs not only improve charge carrier dynamics but also alleviate photo-corrosion of CdS by scavenging photogenerated holes.
Additionally, CuS and MoS2 nanosheets modified with doped CDs have demonstrated remarkable improvements in Fenton-like and piezo-photocatalytic degradation of tetracycline, owing to enhanced charge separation, radical generation, and internal electric field effects. Similarly, WS2–PANI–CDs composites exploit the conducting nature of PANI and the electron-trapping ability of CDs to mitigate charge recombination and achieve nearly complete degradation of pharmaceutical pollutants. In ternary chalcogenides like ZnIn2S4, CDs facilitate not only enhanced photogenerated carrier separation but also in situ H2O2 production, further enabling advanced oxidation processes such as self-Fenton degradation of antibiotics. The synergistic effects between CDs and metal sulfides reinforce the generation of ROS, which are primarily responsible for the oxidative degradation of various organic contaminants in wastewater. This integration also enhances stability and reusability, making CDs–metal sulfide composites promising candidates for sustainable water purification technologies. Beyond chalcogenides, In2S3 heterostructures either with or without CDs exhibited flower-like morphologies, where the incorporation of CDs substantially boosts light absorption and suppresses carrier recombination.
Furthermore, the modification of MXenes with CDs not only improves their environmental stability and dispersion but also opens new avenues in photo(electro)catalytic water treatment and hydrogen evolution. In another promising direction, the fabrication of photocatalysts based on metal oxyhalides, especially BiOCl and CDs has garnered research interest, since extended light absorption and facile charge migration has been highlighted in the above presented studies. CDs also have a contributory role to these systems by triggering OVs in BiOCl and supporting ROS generation and therefore pollutant degradation.
Another promising strategy highlighted in this review was the development of metal-free heterostructures based on CDs, especially those with g–C3N4. Recent advancements clearly demonstrate that the incorporation of CDs into g–C3N4 architectures offers several improvements. CDs serve not only as photosensitizers, enhancing visible-light harvesting via up-conversion photoluminescence, but also as electron acceptors, significantly suppressing electron–hole recombination, therefore increasing ROS formation. Additionally, surface engineering of CDs has been shown to play a crucial role in tuning the photocatalytic behavior of the resulting heterostructures in terms of heteroatom doping or tuning with functional groups. Tailoring these surface functionalities can significantly influence charge separation efficiency, ROS generation, and pollutant-specific interactions, as demonstrated in several recent mechanistic studies. Among other parameters, the role of by products in the final performance has been investigated, pointing out that fluorescent by-products or unreacted precursor compounds can mitigate the photocatalyst’s efficiency. So, it is of the utmost importance to develop purification protocols that ensure purity and reproducibility.
The development of CDs-based ternary heterostructures has significantly advanced photocatalytic water purification by enabling improved charge separation, broader light absorption, and enhanced redox activity. However, despite these promising achievements, several challenges must be addressed to transition these materials from laboratory research to real-world applications. Stability and reusability of CDs-based ternary heterostructures under real, complex environmental matrices also require further examination. Photocatalysts often face challenges such as photo-corrosion, leaching of metal components, and performance loss over multiple cycles. Future research should aim to develop more robust and environmentally benign materials with long-term operational stability.
Independent from the choice of the inorganic host, research studies are focused on the implementation of doped CDs, which further highlights the importance of surface chemistry to optical properties and photocatalytic performance of the as-synthesized platforms. Other studies propose an eco-friendly approach for CDs synthesis by using natural materials as precursors, including crocus cancellatus, mushrooms, fruit waste, etc. Moreover, CDs content in the final heterostructures has been a critical factor for the evaluation of the proposed novel photocatalysts. A rise in CDs concentration led to improved action, but an excessive amount of CDs accounts for inhibition of photocatalysts’ active sites. However, a deeper understanding of how CDs properties such as size, surface chemistry, and doping configuration govern the photocatalytic mechanisms in heterostructures is needed.
Altogether, these examples highlight a growing research trend toward rational design of semiconductor-CDs hybrids, where the selection of the inorganic host is guided by its complementary interaction with CDs to enhance light harvesting, charge dynamics, and photocatalytic redox activity. The diversity of these hosts underscores the adaptability of CDs and their potential to act as universal enhancers across a broad spectrum of catalytic materials.

4. Conclusions

This review highlighted the recent advancements in CDs-based photocatalytic heterostructures for water purification, with a particular emphasis on their synthesis strategies, structural features, and mechanisms of action. CDs have emerged as versatile and efficient co-catalysts or sensitizers due to their unique optical and electronic properties, low toxicity, and ease of functionalization, as well as low cost and environmentally benign synthesis. Their ability to enhance charge carrier separation, extend light absorption into the visible region, and interact with a wide range of inorganic conventional photocatalysts positions them as key components in the development of innovative photocatalytic platforms.
Through a detailed examination of binary and ternary composites, the present review article underscored the influence of material design in the photocatalytic performance, since several factors, including pH, temperature, synthesis, and post-synthesis modifications affect their performance. Among them, CDs surface functionalization and loading amount have aroused research interest due to their influence on their interaction with the host photocatalyst and therefore their photocatalytic activity. These hybrid systems have demonstrated excellent performance in the degradation of various organic pollutants, including pharmaceuticals, dyes, and industrial chemicals, under visible light irradiation.
Apart from improved photocatalytic efficiency, CDs addition provides reinforced stability, environmental adaptability, and ROS generation pathways. There is growing alignment among photocatalytic application and sustainability driven by the utilization of biomass-derived precursors and green chemistry approaches for the fabrication of these heterostructures.
Collectively, CDs-based heterostructures represent a promising class of materials for addressing the global challenge of water contamination. Future research should prioritize a deeper mechanistic understanding, precise optimization of several parameters, potential for scale-up synthesis, and thorough evaluation of long-term stability under realistic environmental conditions to facilitate the practical application of CDs-based photocatalytic platforms.

Author Contributions

Conceptualization, C.A., A.N. and A.Z.; methodology, A.Z. and A.N.; validation, C.A., A.Z., A.N., N.A. (Nikolaos Argirusis) and G.S.; investigation, A.Z., A.N., N.A. (Nikolaos Argirusis), N.A. (Niyaz Alizadeh) and M.A.; writing—original draft preparation, A.Z., A.N., A.K., C.A. and M.A.; writing—review and editing, A.Z., A.N., A.K., C.A., G.S. and K.V.K.; visualization, A.Z.; supervision, A.N., C.A., K.V.K. and G.S.; and project administration, C.A., K.V.K. and G.S. All authors have read and agreed to the published version of the manuscript.

Funding

A.K. acknowledges the Special Account for Research Funding (E.L.K.E.) of the National Technical University of Athens (N.T.U.A.) of Greece for funding. This research received no additional external funding.

Data Availability Statement

No new data has been produced and included in the present manuscript.

Conflicts of Interest

Author Nikolaos Argirusis and Niyaz Alizadeh were employed by the company mat4nrg GmbH. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
AMXamoxicillin
BPAbisphenol A
CBconduction band
CDscarbon dots
CHCchrysin hydrochloride
CNDscarbon nanodots
CPDscarbon polymer dots
CPScationic polystyrene
CPXciprofloxacin
CQDscarbon quantum dots
GQDsgraphene quantum dots
g–C3N4graphitic nitride
IUPACInternational Union of Pure and Applied Chemistry
EISelectrochemical impedance spectroscopy
ESRelectron spin resonance
ESTestradiol
FT-IRFourier transform infrared spectroscopy
LEVlevofloxacin
MBmethylene blue
MGmalachite green
MOmethyl orange
MRSAmethicillin-resistant staphylococcus aureus
NFTnitrofurantoin
NPsnanoparticles
NTsnanotubes
OFXofloxacin
OVsoxygen vacancies
PLphotoluminescence
PANIpolyaniline
PMSperoxymonosulfate
PNPp-nitrophenol
RR141reactive red azo dye
Redoxreduction/oxidation
RhBrhodamine B
ROSreactive oxygen species
SEMscanning electron microscopy
TEMtransmission electron microscopy
TCtetracycline
TCHtetracycline hydrochloride
UVultraviolet
VBvalence band
Visvisible
XRDX-ray diffraction
XPSX-ray photoelectron spectroscopy

References

  1. Babuji, P.; Thirumalaisamy, S.; Duraisamy, K.; Periyasamy, G. Human Health Risks Due to Exposure to Water Pollution: A Review. Water 2023, 15, 2532. [Google Scholar] [CrossRef]
  2. Singh, V.; Ahmed, G.; Vedika, S.; Kumar, P.; Chaturvedi, S.K.; Rai, S.N.; Vamanu, E.; Kumar, A. Toxic Heavy Metal Ions Contamination in Water and Their Sustainable Reduction by Eco-Friendly Methods: Isotherms, Thermodynamics and Kinetics Study. Sci. Rep. 2024, 14, 7595. [Google Scholar] [CrossRef]
  3. AbuQamar, S.F.; El-Saadony, M.T.; Alkafaas, S.S.; Elsalahaty, M.I.; Elkafas, S.S.; Mathew, B.T.; Aljasmi, A.N.; Alhammadi, H.S.; Salem, H.M.; Abd El-Mageed, T.A.; et al. Ecological Impacts and Management Strategies of Pesticide Pollution on Aquatic Life and Human Beings. Mar. Pollut. Bull. 2024, 206, 116613. [Google Scholar] [CrossRef]
  4. Wada, O.Z.; Olawade, D.B. Recent Occurrence of Pharmaceuticals in Freshwater, Emerging Treatment Technologies, and Future Considerations: A Review. Chemosphere 2025, 374, 144153. [Google Scholar] [CrossRef]
  5. Dutta, S.; Adhikary, S.; Bhattacharya, S.; Roy, D.; Chatterjee, S.; Chakraborty, A.; Banerjee, D.; Ganguly, A.; Nanda, S.; Rajak, P. Contamination of Textile Dyes in Aquatic Environment: Adverse Impacts on Aquatic Ecosystem and Human Health, and Its Management Using Bioremediation. J. Environ. Manag. 2024, 353, 120103. [Google Scholar] [CrossRef] [PubMed]
  6. Vijayanand, M.; Ramakrishnan, A.; Subramanian, R.; Issac, P.K.; Nasr, M.; Khoo, K.S.; Rajagopal, R.; Greff, B.; Wan Azelee, N.I.; Jeon, B.-H.; et al. Polyaromatic Hydrocarbons (PAHs) in the Water Environment: A Review on Toxicity, Microbial Biodegradation, Systematic Biological Advancements, and Environmental Fate. Environ. Res. 2023, 227, 115716. [Google Scholar] [CrossRef]
  7. Kye, H.; Kim, J.; Ju, S.; Lee, J.; Lim, C.; Yoon, Y. Microplastics in Water Systems: A Review of Their Impacts on the Environment and Their Potential Hazards. Heliyon 2023, 9, e14359. [Google Scholar] [CrossRef] [PubMed]
  8. Mishra, S.; Sundaram, B. A Review of the Photocatalysis Process Used for Wastewater Treatment. Mater. Today Proc. 2024, 102, 393–409. [Google Scholar] [CrossRef]
  9. Li, B.; Qi, B.; Guo, Z.; Wang, D.; Jiao, T. Recent Developments in the Application of Membrane Separation Technology and Its Challenges in Oil-Water Separation: A Review. Chemosphere 2023, 327, 138528. [Google Scholar] [CrossRef] [PubMed]
  10. Tahraoui, H.; Toumi, S.; Boudoukhani, M.; Touzout, N.; Sid, A.N.E.H.; Amrane, A.; Belhadj, A.-E.; Hadjadj, M.; Laichi, Y.; Aboumustapha, M.; et al. Evaluating the Effectiveness of Coagulation–Flocculation Treatment Using Aluminum Sulfate on a Polluted Surface Water Source: A Year-Long Study. Water 2024, 16, 400. [Google Scholar] [CrossRef]
  11. Satyam, S.; Patra, S. Innovations and Challenges in Adsorption-Based Wastewater Remediation: A Comprehensive Review. Heliyon 2024, 10, e29573. [Google Scholar] [CrossRef]
  12. Zhou, S.; Jia, Y.; Fang, H.; Jin, C.; Mo, Y.; Xiao, Z.; Zhang, N.; Sun, L.; Lu, H. A New Understanding on the Prerequisite of Antibiotic Biodegradation in Wastewater Treatment: Adhesive Behavior between Antibiotic-Degrading Bacteria and Ciprofloxacin. Water Res. 2024, 252, 121226. [Google Scholar] [CrossRef] [PubMed]
  13. Mohamadpour, F.; Mohammad Amani, A. Photocatalytic Systems: Reactions, Mechanism, and Applications. RSC Adv. 2024, 14, 20609–20645. [Google Scholar] [CrossRef] [PubMed]
  14. Ren, G.; Han, H.; Wang, Y.; Liu, S.; Zhao, J.; Meng, X.; Li, Z. Recent Advances of Photocatalytic Application in Water Treatment: A Review. Nanomaterials 2021, 11, 1804. [Google Scholar] [CrossRef] [PubMed]
  15. Mohd Zaki, R.S.R.; Jusoh, R.; Chanakaewsomboon, I.; Setiabudi, H.D. Recent Advances in Metal Oxide Photocatalysts for Photocatalytic Degradation of Organic Pollutants: A Review on Photocatalysts Modification Strategies. Mater. Today Proc. 2024, 107, 59–67. [Google Scholar] [CrossRef]
  16. Zhang, S.; Ou, X.; Xiang, Q.; Carabineiro, S.A.C.; Fan, J.; Lv, K. Research Progress in Metal Sulfides for Photocatalysis: From Activity to Stability. Chemosphere 2022, 303, 135085. [Google Scholar] [CrossRef]
  17. Adabala, S.; Dutta, D.P. A Review on Recent Advances in Metal Chalcogenide-Based Photocatalysts for CO2 Reduction. J. Environ. Chem. Eng. 2022, 10, 107763. [Google Scholar] [CrossRef]
  18. Suresh, R.; Rajendran, S.; Kumar, P.S.; Hoang, T.K.A.; Soto-Moscoso, M. Halides and Oxyhalides-Based Photocatalysts for Abatement of Organic Water Contaminants—An Overview. Environ. Res. 2022, 212, 113149. [Google Scholar] [CrossRef]
  19. Li, N.; Huo, J.; Zhang, Y.; Ye, B.; Chen, X.; Li, X.; Xu, S.; He, J.; Chen, X.; Tang, Y.; et al. Transition Metal Carbides/Nitrides (MXenes): Properties, Synthesis, Functional Modification and Photocatalytic Application. Sep. Purif. Technol. 2024, 330, 125325. [Google Scholar] [CrossRef]
  20. Bhanderi, D.; Lakhani, P.; Modi, C.K. Graphitic Carbon Nitride (g-C3N4) as an Emerging Photocatalyst for Sustainable Environmental Applications: A Comprehensive Review. RSC Sustain. 2024, 2, 265–287. [Google Scholar] [CrossRef]
  21. Noureen, L.; Wang, Q.; Humayun, M.; Shah, W.A.; Xu, Q.; Wang, X. Recent Advances in Structural Engineering of Photocatalysts for Environmental Remediation. Environ. Res. 2023, 219, 115084. [Google Scholar] [CrossRef]
  22. Kozak, M.; Mazierski, P.; Żebrowska, J.; Klimczuk, T.; Lisowski, W.; Żak, A.M.; Skowron, P.M.; Zaleska-Medynska, A. Detailed Insight into Photocatalytic Inactivation of Pathogenic Bacteria in the Presence of Visible-Light-Active Multicomponent Photocatalysts. Nanomaterials 2024, 14, 409. [Google Scholar] [CrossRef]
  23. Wei, Y.; Wu, Q.; Meng, H.; Zhang, Y.; Cao, C. Recent Advances in Photocatalytic Self-Cleaning Performances of TiO2-Based Building Materials. RSC Adv. 2023, 13, 20584–20597. [Google Scholar] [CrossRef]
  24. Abhishek, B.; Jayarama, A.; Rao, A.S.; Nagarkar, S.S.; Dutta, A.; Duttagupta, S.P.; Prabhu, S.S.; Pinto, R. Challenges in Photocatalytic Hydrogen Evolution: Importance of Photocatalysts and Photocatalytic Reactors. Int. J. Hydrogen Energy 2024, 81, 1442–1466. [Google Scholar] [CrossRef]
  25. Kumagai, H.; Tamaki, Y.; Ishitani, O. Photocatalytic Systems for CO2 Reduction: Metal-Complex Photocatalysts and Their Hybrids with Photofunctional Solid Materials. Acc. Chem. Res. 2022, 55, 978–990. [Google Scholar] [CrossRef]
  26. Ansari, A.S.; Azzahra, G.; Nugroho, F.G.; Mujtaba, M.M.; Ahmed, A.T.A. Oxides and Metal Oxide/Carbon Hybrid Materials for Efficient Photocatalytic Organic Pollutant Removal. Catalysts 2025, 15, 134. [Google Scholar] [CrossRef]
  27. Akbari, M.; Rasouli, J.; Rasouli, K.; Ghaedi, S.; Mohammadi, M.; Rajabi, H.; Sabbaghi, S. MXene-Based Composite Photocatalysts for Efficient Degradation of Antibiotics in Wastewater. Sci. Rep. 2024, 14, 31498. [Google Scholar] [CrossRef] [PubMed]
  28. Abdelfattah, I.; El-Shamy, A.M. A Comparative Study for Optimizing Photocatalytic Activity of TiO2-Based Composites with ZrO2, ZnO, Ta2O5, SnO, Fe2O3, and CuO Additives. Sci. Rep. 2024, 14, 27175. [Google Scholar] [CrossRef] [PubMed]
  29. Baig, A.; Siddique, M.; Panchal, S. A Review of Visible-Light-Active Zinc Oxide Photocatalysts for Environmental Application. Catalysts 2025, 15, 100. [Google Scholar] [CrossRef]
  30. He, X.; Wang, A.; Wu, P.; Tang, S.; Zhang, Y.; Li, L.; Ding, P. Photocatalytic Degradation of Microcystin-LR by Modified TiO2 Photocatalysis: A Review. Sci. Total Environ. 2020, 743, 140694. [Google Scholar] [CrossRef] [PubMed]
  31. Jie, L.; Gao, X.; Cao, X.; Wu, S.; Long, X.; Ma, Q.; Su, J. A Review of CdS Photocatalytic Nanomaterials: Morphology, Synthesis Methods, and Applications. Mater. Sci. Semicond. Process. 2024, 176, 108288. [Google Scholar] [CrossRef]
  32. Wang, C.-Y.; Zhang, X.; Yu, H.-Q. Bismuth Oxyhalide Photocatalysts for Water Purification: Progress and Challenges. Coord. Chem. Rev. 2023, 493, 215339. [Google Scholar] [CrossRef]
  33. Iqbal, A.; Hong, J.; Ko, T.Y.; Koo, C.M. Improving Oxidation Stability of 2D MXenes: Synthesis, Storage Media, and Conditions. Nano Converg. 2021, 8, 9. [Google Scholar] [CrossRef]
  34. Keneshbekova, A.; Smagulova, G.; Kaidar, B.; Imash, A.; Ilyanov, A.; Kazhdanbekov, R.; Yensep, E.; Lesbayev, A. MXene/Carbon Nanocomposites for Water Treatment. Membranes 2024, 14, 184. [Google Scholar] [CrossRef]
  35. Asrami, M.R.; Jourshabani, M.; Park, M.H.; Shin, D.; Lee, B. A Unique and Well-Designed 2D Graphitic Carbon Nitride with Sponge-like Architecture for Enhanced Visible-Light Photocatalytic Activity. J. Mater. Sci. Technol. 2023, 159, 99–111. [Google Scholar] [CrossRef]
  36. Mohtar, S.S.; Aziz, F.; Ismail, A.F.; Sambudi, N.S.; Abdullah, H.; Rosli, A.N.; Ohtani, B. Impact of Doping and Additive Applications on Photocatalyst Textural Properties in Removing Organic Pollutants: A Review. Catalysts 2021, 11, 1160. [Google Scholar] [CrossRef]
  37. Zhang, S.; Xu, Z.; Ji, T.; Guan, P.; Weng, Y. Carbon Quantum Dots Modified MoS2 for High-Efficiency and Long-Endurance Persulfate Activation: Enhanced Electron Transfer and Piezoelectricity. Sep. Purif. Technol. 2025, 353, 128148. [Google Scholar] [CrossRef]
  38. Tran, V.V.; Nu, T.T.V.; Jung, H.-R.; Chang, M. Advanced Photocatalysts Based on Conducting Polymer/Metal Oxide Composites for Environmental Applications. Polymers 2021, 13, 3031. [Google Scholar] [CrossRef]
  39. Yadav, A.N.; Kumar, R.; Jaiswal, R.K.; Singh, A.K.; Kumar, P.; Singh, K. Surface Modification of CdS Quantum Dots: An Effective Approach for Improving Biocompatibility. Mater. Res. Express 2019, 6, 055002. [Google Scholar] [CrossRef]
  40. Zhang, Y.; Chen, Q.; Xiao, Q.; Shi, L.; Zhao, Z.; Wang, H. Enhancement of CdS Resistance to Photocorrosion and Photocatalytic Removal of Uranyl by Complexation with N-Deficient g-C3N4 under Aerobic Conditions. Chemosphere 2023, 335, 139022. [Google Scholar] [CrossRef] [PubMed]
  41. Zeng, Q.; Zhou, X.; Shen, L.; Zhao, D.L.; Kong, N.; Li, Y.; Qiu, X.; Chen, C.; Teng, J.; Xu, Y.; et al. Exceptional Self-Cleaning MXene-Based Membrane for Highly Efficient Oil/Water Separation. J. Membr. Sci. 2024, 700, 122691. [Google Scholar] [CrossRef]
  42. Falara, P.P.; Zourou, A.; Kordatos, K.V. Recent Advances in Carbon Dots/2-D Hybrid Materials. Carbon 2022, 195, 219–245. [Google Scholar] [CrossRef]
  43. Xu, X.; Ray, R.; Gu, Y.; Ploehn, H.J.; Gearheart, L.; Raker, K.; Scrivens, W.A. Electrophoretic Analysis and Purification of Fluorescent Single-Walled Carbon Nanotube Fragments. J. Am. Chem. Soc. 2004, 126, 12736–12737. [Google Scholar] [CrossRef]
  44. Xia, C.; Zhu, S.; Feng, T.; Yang, M.; Yang, B. Evolution and Synthesis of Carbon Dots: From Carbon Dots to Carbonized Polymer Dots. Adv. Sci. 2019, 6, 1901316. [Google Scholar] [CrossRef]
  45. Zhai, Y.; Zhang, B.; Shi, R.; Zhang, S.; Liu, Y.; Wang, B.; Zhang, K.; Waterhouse, G.I.N.; Zhang, T.; Lu, S. Carbon Dots as New Building Blocks for Electrochemical Energy Storage and Electrocatalysis. Adv. Energy Mater. 2022, 12, 2103426. [Google Scholar] [CrossRef]
  46. Gengan, S.; Ananda Murthy, H.C.; Sillanpää, M.; Nhat, T. Carbon Dots and Their Application as Photocatalyst in Dye Degradation Studies—Mini Review. Results Chem. 2022, 4, 100674. [Google Scholar] [CrossRef]
  47. Ai, L.; Yang, Y.; Wang, B.; Chang, J.; Tang, Z.; Yang, B.; Lu, S. Insights into Photoluminescence Mechanisms of Carbon Dots: Advances and Perspectives. Sci. Bull. 2021, 66, 839–856. [Google Scholar] [CrossRef] [PubMed]
  48. Minervini, G.; Panniello, A.; Madonia, A.; Carbonaro, C.M.; Mocci, F.; Sibillano, T.; Giannini, C.; Comparelli, R.; Ingrosso, C.; Depalo, N.; et al. Photostable Carbon Dots with Intense Green Emission in an Open Reactor Synthesis. Carbon 2022, 198, 230–243. [Google Scholar] [CrossRef]
  49. Soledad-Flores, O.; Bailón-Ruiz, S.J.; Román-Velázquez, F. Rapid Synthesis of Non-Toxic, Water-Stable Carbon Dots Using Microwave Irradiation. Micro 2024, 4, 659–669. [Google Scholar] [CrossRef]
  50. de Oliveira, B.P.; da Silva Abreu, F.O.M. Carbon Quantum Dots Synthesis from Waste and By-Products: Perspectives and Challenges. Mater. Lett. 2021, 282, 128764. [Google Scholar] [CrossRef]
  51. Kaczmarek, A.; Hoffman, J.; Morgiel, J.; Mościcki, T.; Stobiński, L.; Szymański, Z.; Małolepszy, A. Luminescent Carbon Dots Synthesized by the Laser Ablation of Graphite in Polyethylenimine and Ethylenediamine. Materials 2021, 14, 729. [Google Scholar] [CrossRef] [PubMed]
  52. Khayal, A.; Dawane, V.; Amin, M.A.; Tirth, V.; Yadav, V.K.; Algahtani, A.; Khan, S.H.; Islam, S.; Yadav, K.K.; Jeon, B.-H. Advances in the Methods for the Synthesis of Carbon Dots and Their Emerging Applications. Polymers 2021, 13, 3190. [Google Scholar] [CrossRef] [PubMed]
  53. Zourou, A.; Ntziouni, A.; Roman, T.; Tampaxis, C.; Steriotis, T.; Gkouzia, G.; Alff, L.; Sanchez, D.E.; Terrones, M.; Kordatos, K.V. Synthesis and Characterization of Magnetic Carbon Dots (CDs)-Based Hybrid Material as an Adsorbent for the Removal of Organic Dye from Water. Carbon 2024, 230, 119612. [Google Scholar] [CrossRef]
  54. Wang, C.; Yang, M.; Shi, H.; Yao, Z.; Liu, E.; Hu, X.; Guo, P.; Xue, W.; Fan, J. Carbon Quantum Dots Prepared by Pyrolysis: Investigation of the Luminescence Mechanism and Application as Fluorescent Probes. Dyes Pigments 2022, 204, 110431. [Google Scholar] [CrossRef]
  55. Nazar, M.; Hasan, M.; Wirjosentono, B.; Gani, B.A.; Nada, C.E. Microwave Synthesis of Carbon Quantum Dots from Arabica Coffee Ground for Fluorescence Detection of Fe3+, Pb2+, and Cr3+. ACS Omega 2024, 9, 20571–20581. [Google Scholar] [CrossRef]
  56. Zhao, D.; Liu, X.; Wei, C.; Qu, Y.; Xiao, X.; Cheng, H. One-Step Synthesis of Red-Emitting Carbon Dots via a Solvothermal Method and Its Application in the Detection of Methylene Blue. RSC Adv. 2019, 9, 29533–29540. [Google Scholar] [CrossRef]
  57. Kaur, I.; Batra, V.; Bogireddy, N.K.R.; Baveja, J.; Kumar, Y.; Agarwal, V. Chemical- and Green-Precursor-Derived Carbon Dots for Photocatalytic Degradation of Dyes. iScience 2024, 27, 108920. [Google Scholar] [CrossRef]
  58. Habibi-Yangjeh, A.; Pournemati, K.; Ahmasdi, Z.; Khatase, A. Decoration of Carbon Dots on Oxygen-Vacancy-Enriched S-Scheme TiO2 Quantum Dots/TiO2Oxygen Vacancies Photocatalysts: Impressive Quantum-Dot Sized Photocatalysts for Remediation of Antibiotics, Bacteria, and Dyes. Langmuir 2024, 40, 8503–8519. [Google Scholar] [CrossRef]
  59. Camilli, E.; Foglia, M.L.; Bravo, J.P.; Copello, G.J.; Villanueva, M.E. Comparison among OH, N and P functionalized carbon quantum dots/TiO2 nanocomposites for food industry wastewater remediation. Diam. Relat. Mater. 2024, 145, 111103. [Google Scholar] [CrossRef]
  60. Sendao, R.M.S.; Algarra, M.; Lazaro-Martinez, J.; Brandao, A.T.S.C.; Gil, A.; Pereira, C.; Esteves de Silva, J.C.G.; Pinto da Silva, L. Visible-light-driven photocatalytic degradation of organic dyes using a TiO2 and waste-based carbon dots nanocomposite. Colloids Surf. A Physicochem. Eng. Asp. 2025, 713, 136475. [Google Scholar] [CrossRef]
  61. Ma, R.; Xu, Z.; Zhang, C.; Li, H.; Chen, J.; Fan, J.; Shi, Q. Highly efficient reduction of Cr(VI) from industries sewage using novel biomass-driven carbon dots modified TiO2 under sunlight. Chem. Eng. J. 2024, 500, 157480. [Google Scholar] [CrossRef]
  62. Rawat, J.; Sharma, H.; Dwivedi, C. Microwave-assisted synthesis of carbon quantum dots and their integration with TiO2 nanotubes for enhanced photocatalytic degradation. Diam. Relat. Mater. 2024, 144, 111050. [Google Scholar] [CrossRef]
  63. Bui, B.-C.; Vu, N.-N.; Nemamchs, H.-E.; Nguyen, H.T.; Nguyen, V.-A.; Nguyen-Tri, P. Single nickel atoms doped into TiO2 decorating carbon quantum dots for boosting photodegradation of ciprofloxacin. J. Water Process Eng. 2025, 70, 106904. [Google Scholar] [CrossRef]
  64. Taghiloo, B.; Shahnazi, A.; Nabid, M.R. Construction of nanocomposite hydrogel by TiO2-carbon quantum dots encapsulated in alginate with a highly efficient adsorption and photodegradation of dye pollutants. J. Alloys Compd. 2024, 1005, 175859. [Google Scholar] [CrossRef]
  65. Ayu, D.G.; Gea, S.; Andriayani; Telaumbanua, D.J.; Piliang, A.F.R.; Harahap, M.; Yen, Z.; Goei, R.; Tok, A.I.Y. Photocatalytic Degradation of Methylene Blue Using NDoped ZnO/Carbon Dot (N-ZnO/CD) Nanocomposites Derived from Organic Soybean. ACS Omega 2023, 8, 14965–14984. [Google Scholar] [CrossRef]
  66. Hidayat, N.; Widiyandari, H.; Parasdila, H.; Prilita, O.; Astuti, Y.; Mufti, N.; Ogi, T. Green synthesis of ZnO photocatalyst composited carbon quantum dots (CQDs) from lime (Citrus aurantifolia). Catal. Commun. 2024, 187, 106888. [Google Scholar] [CrossRef]
  67. Parveen, S.; Latif, N.; Chotana, G.A.; Kanwal, A.; Hussain, S.; Habila, M.A.; Iqbal, A.; Manavalan, R.K.; Farooq, N. ZnO/carbon quantum dots nanocomposites derived from Moringa oleifera gum: An improved catalytic vitiation of methylene blue dye. Mater. Res. Bull. 2025, 181, 113106. [Google Scholar] [CrossRef]
  68. Kalifathullah, S.K.; Sundaramurthy, D. Exploration of biological activities of green NCarbon Quantum Dots and photocatalytic studies of ZnO@NCQDs. Emergent Mater. 2024, 7, 2755–2766. [Google Scholar] [CrossRef]
  69. Xu, J.-J.; Lu, Y.-N.; Tao, F.-F.; Liang, P.-F.; Zhang, P.-A. ZnO Nanoparticles Modified by Carbon Quantum Dots for the Photocatalytic Removal of Synthetic Pigment Pollutants. ACS Omega 2023, 8, 7845–7857. [Google Scholar] [CrossRef] [PubMed]
  70. Bhavsar, F.S.; Mapari, M.G.; Tivalekar, S.R.; Naz, E.G.; Babar, D.G. A Nanocomposite of ZnO and N,P-Co-Doped Carbon Dotsfor the Photocatalytic Degradation of Various Dyes and their Kinetic Study. ChemistrySelect 2023, 8, e202301821. [Google Scholar] [CrossRef]
  71. Karaca, C.; Eroğlu, Z.; Karaca, S. Anthraquinone-Rich Rheum ribes L. as a Source of Nitrogen-Doped Carbon Quantum Dots for ZnO-Based S-Scheme Heterojunction Photocatalysts in Tetracycline Degradation. J. Environ. Chem. Eng. 2025, 13, 115999. [Google Scholar] [CrossRef]
  72. Nugroho, D.; Wannakan, K.; Nanan, S.; Benchawattananon, R. The Synthesis of carbon dots//zincoxide (CDs/ZnOH400) by using hydrothermal methods for degradation of ofloxacin antibiotics and reactive red azo dye (RR141). Sci. Rep. 2024, 14, 2455. [Google Scholar] [CrossRef]
  73. Nugroho, D.; Khoris, I.M.; Yoskamtorn, T.; Nanan, S.; Lee, J.; Benchawattananon, R. Hybrid nanostructure carbon dots/zinc oxide from Rosa indica for photodegradation of various pharmaceuticals pollutants. J. Water Process Eng. 2025, 70, 106828. [Google Scholar] [CrossRef]
  74. Jin, Y.; Tang, W.; Wang, J.; Ren, F.; Chen, Z.; Sun, Z. Construction of biomass derived carbon quantum dots modified TiO2 photocatalysts with superior photocatalytic activity for methylene blue degradation. J. Alloys Compd. 2023, 932, 167627. [Google Scholar] [CrossRef]
  75. Prabhakaran, P.K.; Balu, S.; Sridharan, G.; Ganapathy, D.; Sundramoorthy, A.K. Development of Eco-friendly CQDs/TiO2 nanocomposite for enhanced photocatalytic degradation of methyl orange dye. Eng. Res. Express 2025, 7, 015002. [Google Scholar] [CrossRef]
  76. Hsieh, M.-L.; Juang, R.-S.; Gansomi, Y.A.; Fu, C.-c.; Hsieh, C.-t.; Liu, W.-R. Synthesis and characterization of high performance ZnO/graphene quantum dot composites for photocatalytic degradation of metronidazole. J. Taiwan Inst. Chem. Eng. 2022, 131, 104180. [Google Scholar] [CrossRef]
  77. Preethi, G.; Pillai, R.; Ramdas, B.; Ramamoorthy, S.; Patil, B.; Lekshmi, I.C.; Mohan Kumar, P.; Rangaraj, L. Role of Carbon Quantum Dots in a Strategic Approach to Prepare Pristine Zn2SnO4 and Enhance Photocatalytic Activity under Direct Sunlight. Diam. Relat. Mater. 2023, 131, 109554. [Google Scholar] [CrossRef]
  78. Liu, X.; Lin, Q.; Zhao, L.; Fang, J.; Qi, J.; Fan, H.; Yue, X.; Li, G.; Qian, Y.; Li, H. Wood-Supported Nitrogen-Doped Carbon Quantum Dot @Cu2O Composites for Efficient Photocatalytic Degradation of Dye Wastewater. Cellulose 2024, 31, 7587–7600. [Google Scholar] [CrossRef]
  79. Wang, J.; Li, S.; Ma, P.; Guo, Z.; Ma, Q.; Zhao, Q.; Guo, Y.; Zhao, J.; Guan, G. Carbon Quantum Dots/ Cu2O S-Scheme Heterojunction for Enhanced Photocatalytic Degradation of Tetracycline. Colloids Surf. Physicochem. Eng. Asp. 2024, 690, 133779. [Google Scholar] [CrossRef]
  80. Chen, X.; Chen, C.; Zang, J. Bi2MoO6 Nanoflower-like Microsphere Photocatalyst Modified by Boron Doped Carbon Quantum Dots: Improving the Photocatalytic Degradation Performance of BPA in All Directions. J. Alloys Compd. 2023, 962, 171167. [Google Scholar] [CrossRef]
  81. Wang, X.; Chen, C.; Wang, Q.; Dai, K. Graphene Quantum Dot Modified Bi2MoO6 Nanoflower for Efficient Degradation of BPA under Visible Light. Chin. J. Struct. Chem. 2024, 43, 100473. [Google Scholar] [CrossRef]
  82. Liu, J.; Ji, L.; He, Q.; Zang, S.; Sun, J.; Yang, H.; Dong, T.; Liu, T.; Wu, H.; Chen, X.; et al. Algal Carbon Quantum Dots/Bi2MoO6 S-Scheme Heterojunction with Enhanced Visible-Light Photocatalytic Degradation for Ciprofloxacin. Sep. Purif. Technol. 2025, 363, 132196. [Google Scholar] [CrossRef]
  83. Hu, C.; Chen, Q.; Tian, M.; Wang, W.; Yu, J.; Chen, L. Efficient Combination of Carbon Quantum Dots and BiVO4 for Significantly Enhanced Photocatalytic Activities. Catalysts 2023, 13, 463. [Google Scholar] [CrossRef]
  84. Jiteshwaran, T.; Steffy, J.P.; Janani, B.; Syed, A.; Elgorban, A.M.; Abid, I.; Wong, L.S.; Khan, S.S. In Situ Growth of Carbon Quantum Dots on Acid/Base 3D Co2VO4 Nanoplates to Regulate Photocatalysis and Peroxymonosulfate Activation towards Highly Efficient Degradation of Ciprofloxacin. J. Water Process Eng. 2025, 71, 107336. [Google Scholar] [CrossRef]
  85. Wang, Q.; Wen, G.; Yang, Z.; Guo, Q.; Zhang, B.; Nie, Y.; Wang, D. Preparation and Study of CDs-WO3 Composites with Enhanced Photocatalytic Antimicrobial Properties and Degradation of Dyes. Biochem. Eng. J. 2025, 217, 109670. [Google Scholar] [CrossRef]
  86. Aloni, P.; Venkatesan, P.; Sundaresan, A.P.; Roy, D.; Ranjan, R.K.; Sharma, A.; Doong, R.-A.; Clament Sagaya Selvam, N. Unveiling the Impact of Nitrogen-Doped Graphene Quantum Dots on Improving the Photocatalytic Performance of CuWO4 Nanocomposite. Appl. Surf. Sci. 2025, 686, 162130. [Google Scholar] [CrossRef]
  87. Sarwar, A.; Razzaq, A.; Zafar, M.; Idrees, I.; Rehman, F.; Kim, W.Y. Copper Tungstate (CuWO4)/Graphene Quantum Dots (GQDs) Composite Photocatalyst for Enhanced Degradation of Phenol under Visible Light Irradiation. Results Phys. 2023, 45, 106253. [Google Scholar] [CrossRef]
  88. Khan, A.; Valicsek, Z.; Horváth, O.; Khan, M.M.; Wafi, A. Ferrite-Based Photocatalysts: Synthesis, Modifications, and Key Parameters in Photocatalytic-Related Applications. Mater. Today Commun. 2024, 40, 109556. [Google Scholar] [CrossRef]
  89. Esmail, L.A.; Jabbar, H.S.; Salih, S.K. Synthesis of a New Carbon Dot Magnetic Nanocomposite (CDs@Fe3O4) from Crocus Cancellatus: Characterization and Its Photocatalytic Degradation of Fluorescein Dye. Inorg. Chem. Commun. 2024, 159, 111823. [Google Scholar] [CrossRef]
  90. Joga, S.B.; Korabandi, D.; Lakkaboyana, S.K.; Kumar, V. Synthesis of Iron Nanoparticles on Lemon Peel Carbon Dots (LP-CDs@Fe3O4) Applied in Photo-Catalysis, Antioxidant, Antidiabetic, and Hemolytic Activity. Inorg. Chem. Commun. 2025, 174, 113960. [Google Scholar] [CrossRef]
  91. Sari, E.K.; Tumbelaka, R.M.; Ardiyanti, H.; Istiqomah, N.I.; Chotimah; Suharyadi, E. Green Synthesis of Magnetically Separable and Reusable Fe3O4/Cdots Nanocomposites Photocatalyst Utilizing Moringa oleifera Extract and Watermelon Peel for Rapid Dye Degradation. Carbon Resour. Convers. 2023, 6, 274–286. [Google Scholar] [CrossRef]
  92. Monje, D.S.; Mercado, D.F.; Mesa, G.A.P.; Valencia, G.C. Carbon Dots Decorated Magnetite Nanocomposite Obtained Using Yerba Mate Useful for Remediation of Textile Wastewater through a Photo-Fenton Treatment: Ilex Paraguariensis as a Platform of Environmental Interest-Part 2. Environ. Sci. Pollut. Res. Int. 2023, 30, 3070–3087. [Google Scholar] [CrossRef]
  93. Akhter, T.; Aslam, M. Carbon Dot-Modified Ferrite (CDs@ZF): Efficient Visible-Light Photocatalyst for Dye Degradation and Antibacterial Activity. Inorg. Chem. Commun. 2025, 173, 113777. [Google Scholar] [CrossRef]
  94. Mmelesi, O.K.; Ammar-Merah, S.; Nkambule, T.T.I.; Kefeni, K.K.; Kuvarega, A.T. Synergistic Role of N-Doped Carbon Quantum Dots on Zn-Doped Cobalt Ferrite (N-CQDs/ZnCF) for the Enhanced Photodegradation of Oxytetracycline under Visible Light. Mater. Sci. Eng. B 2023, 294, 116538. [Google Scholar] [CrossRef]
  95. Anh, V.C.N.; Nhi, L.T.T.; Dung, L.T.K.; Hoa, D.T.N.; Son, N.T.; Uyen, N.T.T.; Thu, N.N.U.; Son, L.V.T.; Hieu, L.T.; Tuyen, T.N.; et al. Photocatalytic Degradation of Methylene Blue under Visible Light by Cobalt Ferrite Nanoparticles/Graphene Quantum Dots. Beilstein J. Nanotechnol. 2024, 15, 475–489. [Google Scholar] [CrossRef]
  96. Renu; Nidhi; Kaur, P.; Komal; Minakshi; Paulik, C.; Kaushik, A.; Singhal, S. Rational Design of Boerhavia diffusa Derived CoFe2O4-Carbon dots@Boehmite Platform for Photocatalysis and Ultra Trace Monitoring of Hazardous Pesticide and UO22+ Ions. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2025, 325, 125111. [Google Scholar] [CrossRef] [PubMed]
  97. Mmelesi, O.K.; Ammar-Merah, S.; Nkambule, T.T.I.; Nkosi, B.; Liu, X.; Kefeni, K.K.; Kuvarega, A.T. The Photodegradation of Naproxen in an Aqueous Solution Employing a Cobalt Ferrite-Carbon Quantum Dots (CF/N-CQDs) Nanocomposite, Synthesized via Microwave Approach. J. Water Process Eng. 2024, 59, 104968. [Google Scholar] [CrossRef]
  98. Malitha, M.D.; Molla, M.T.H.; Bashar, M.A.; Chandra, D.; Ahsan, M.S. Fabrication of a Reusable Carbon Quantum Dots (CQDs) Modified Nanocomposite with Enhanced Visible Light Photocatalytic Activity. Sci. Rep. 2024, 14, 17976. [Google Scholar] [CrossRef]
  99. Naghash-Hamed, S.; Arsalani, N.; Mousavi, S.B. Facile Fabrication of CuFe2O4 Coated with Carbon Quantum Dots Nanocomposite as an Efficient Heterogeneous Catalyst toward the Reduction of Nitroaniline Compounds for Management of Aquatic Resources. J. Photochem. Photobiol. Chem. 2023, 443, 114822. [Google Scholar] [CrossRef]
  100. Jamal, F.; Rafique, A.; Moeen, S.; Haider, J.; Nabgan, W.; Haider, A.; Imran, M.; Nazir, G.; Alhassan, M.; Ikram, M.; et al. Review of Metal Sulfide Nanostructures and Their Applications. ACS Appl. Nano Mater. 2023, 6, 7077–7106. [Google Scholar] [CrossRef]
  101. Choudhary, M.; Saini, P.; Chakinala, N.; Surolia, P.K.; Gupta Chakinala, A. Carbon Dots Decorated Cadmium Sulfide Nanomaterials for Boosting Photocatalytic Activity for Ciprofloxacin Degradation. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2024, 319, 124572. [Google Scholar] [CrossRef]
  102. Orak, C.; Oğuz, T.; Horoz, S. Facile Synthesis of Mn-Doped CdS Nanoparticles on Carbon Quantum Dots: Towards Efficient Photocatalysis. J. Aust. Ceram. Soc. 2024, 60, 1657–1667. [Google Scholar] [CrossRef]
  103. Cheng, Y.; Deng, L.; Wang, D.; Wang, X.; Ji, C.; Zhou, Y.-H. CuS@Cu-CD Composites as Efficient Heterogeneous Fenton-like Catalysts for the Photodegradation of Tetracycline. Environ. Sci. Adv. 2023, 2, 495–507. [Google Scholar] [CrossRef]
  104. Qu, Y.; Li, X.; Cui, M.; Huang, R.; Ma, W.; Wang, Y.; Su, R.; Qi, W. Synergetic Assembly of a Molybdenum Disulfide/Carbon Quantum Dot Heterojunction with Enhanced Light Absorption and Electron Transfer Di-Functional Properties for Photocatalysis. Mater. Res. Bull. 2024, 171, 112627. [Google Scholar] [CrossRef]
  105. Li, J.; Li, C.; Chen, L.; Li, T.; Gao, F.; Chen, X.; Zhao, T.; Wang, F.; Jiang, Y. A Gradient Photothermal Hydrogel with Carbon-Dots Deposited Molybdenum Disulfide Nanoflowers as Photothermal Centers for Efficient Solar-Driven Water Evaporation and Treatment. Surf. Interfaces 2024, 55, 105433. [Google Scholar] [CrossRef]
  106. Fatima, T.; Husain, S.; Khanuja, M. Novel Ternary Z Scheme Carbon Quantum Dots (CQDs) Decorated WS2/PANI ((CQDs@WS2/PANI):0D:2D:1D) Nanocomposite for the Photocatalytic Degradation and Electrochemical Detection of Pharmaceutical Drugs. Nano Mater. Sci. 2025, 7, 259–275. [Google Scholar] [CrossRef]
  107. Mishra, S.R.; Gadore, V.; Ahmaruzzaman, M. An Overview of In2S3, In2S3-Based Photocatalyst: Characteristics Synthesis Modifications Design Strategies Catalytic Environmental Application. J. Environ. Chem. Eng. 2024, 12, 113449. [Google Scholar] [CrossRef]
  108. Yao, Y.; Zhao, Z.; Cui, H.; Dong, W.; Li, Z.; Liao, G. Graphene Quantum Dots Sensitized In2S3 Nanohybrids for Improved Photocatalytic Activity. J. Mater. Sci. Mater. Electron. 2023, 34, 1923. [Google Scholar] [CrossRef]
  109. Zhang, G.; Wu, H.; Chen, D.; Li, N.; Xu, Q.; Li, H.; He, J.; Lu, J. A Mini-Review on ZnIn2S4-Based Photocatalysts for Energy and Environmental Application. Green Energy Environ. 2022, 7, 176–204. [Google Scholar] [CrossRef]
  110. Li, X.; Liu, Y.; Huang, H.; Cheng, J. A Photocatalysis-Self-Fenton System Based on NCDs@ZnIn2S4 Composites at Neutral pH and Low Amount of Fe2+ for the Effective Degradation of Antibiotics. J. Environ. Manag. 2024, 370, 122580. [Google Scholar] [CrossRef]
  111. Naguib, M.; Kurtoglu, M.; Presser, V.; Lu, J.; Niu, J.; Heon, M.; Hultman, L.; Gogotsi, Y.; Barsoum, M.W. Two-Dimensional Nanocrystals Produced by Exfoliation of Ti3AlC2. Adv. Mater. 2011, 23, 4248–4253. [Google Scholar] [CrossRef]
  112. Bai, W.; Shi, L.; Li, Z.; Liu, D.; Liang, Y.; Han, B.; Qi, J.; Li, Y. Recent progress on the preparation and application in photocatalysis of 2D MXene-based materials. Mater. Today Energy 2024, 41, 101547. [Google Scholar] [CrossRef]
  113. Seling, T.R.; Songsart-Power, M.; Shringi, A.K.; Paudyal, J.; Yan, F.; Limbu, T.B. Ti3C2Tx MXene-Based Hybrid Photocatalysts in Organic Dye Degradation: A Review. Molecules 2025, 30, 1463. [Google Scholar] [CrossRef] [PubMed]
  114. Yan, Z.; Yin, K.; Xu, M.; Fang, N.; Yu, W.; Chu, Y.; Shu, S. Photocatalysis for synergistic water remediation H2 production: A review. Chem. Eng. J. 2023, 61, 145066. [Google Scholar] [CrossRef]
  115. Liu, Y.; Zhang, W.; Zheng, W. Qantum Dots Compete at the Acme of MXene Family for the Optimal Catalysis. Nano-Micro Lett. 2022, 14, 158. [Google Scholar] [CrossRef] [PubMed]
  116. Tawalbeh, M.; Mohammed, S.; Al-Othman, A.; Yusuf, M.; Mofijur, M.; Kamyab, H. MXenes and MXene-based materials for removal of pharmaceutical compounds from wastewater: Critical review. Environ. Res. 2023, 228, 115919. [Google Scholar] [CrossRef]
  117. Peng, J.; Chen, X.; Ong, W.-J.; Zhao, X.; Li, N. Surface and Heterointerface Engineering of 2D MXenes and Their Nanocomposites: Insights into Electro- and Photocatalysis. Chem 2019, 5, 18–50. [Google Scholar] [CrossRef]
  118. Jiang, S.; Liu, J.; Zhao, K.; Cui, D.; Liu, P.; Yin, H.; Al-Mamun, M.; Lowe, S.E.; Zhang, W.; Zhong, Y.L. Ru(bpy)32+-sensitized {001} facets LiCoO2 nanosheets catalyzed CO2 reduction reaction with 100% carbonaceous products. Nano Res. 2022, 15, 1061–1068. [Google Scholar] [CrossRef]
  119. Sun, Y.; Meng, X.; Dall’Agnese, Y.; Dall’Agnese, C.; Duan, S.; Gao, Y.; Chen, G.; Wang, X.-F. 2D MXenes as Co-Catalysts in Photocatalysis: Synthetic Methods. Nano-Micro Lett. 2019, 11, 79. [Google Scholar] [CrossRef]
  120. Zhu, Y.; Liu, J.; Guo, T.; Wang, J.J.; Nicolosi, V. Multifunctional Ti3C2Tx MXene composite hydrogels with strain sensitivity toward absorption-dominated electromagnetic-interference shielding. ACS Nano. 2021, 15, 1465–1474. [Google Scholar] [CrossRef]
  121. Liu, F.; Zhou, A.; Chen, J.; Jia, J.; Zhou, W.; Wang, L.; Hu, Q. Preparation of Ti3C2 and Ti2C MXenes by fluoride salts etching and methane adsorptive properties. Appl. Surf. Sci. 2017, 416, 781–789. [Google Scholar] [CrossRef]
  122. Li, M.; Lu, J.; Luo, K.; Li, Y.; Chang, K.; Chen, K.; Zhou, J.; Rosen, J.; Hultman, L.; Eklund, P. Element replacement approach by reaction with Lewis acidic molten salts to synthesize nanolaminated MAXphases MXenes. J. Am. Chem. Soc. 2019, 141, 4730–4737. [Google Scholar] [CrossRef]
  123. Miri-Jahromi, A.; Didandeh, M.; Shekarsokhan, S. Capability of MXene 2Dmaterial as an amoxicillin ampicillin cloxacillin adsorbent in wastewater. J. Mol. Liq. 2022, 351, 118545. [Google Scholar] [CrossRef]
  124. Alyasi, H.; Wahib, S.; Alcantara Gomez, T.; Rasool, K.; Mahmoud, K.A. The power of MXene-based materials for emerging contaminant removal from water—A review. Desalination 2024, 586, 117913. [Google Scholar] [CrossRef]
  125. Elemike, E.E.; Adeyemi, J.; Onwudiwe, D.C.; Wei, L.; Oyedeji, A.O. The future of energy materials: A case of MXenes-carbon dots nanocomposites. J. Energy Storage 2022, 50, 104711. [Google Scholar] [CrossRef]
  126. Guo, Z.; Li, Y.; Lu, Z.; Chao, Y.; Liu, W. High-performance MnO2@MXene/carbon nanotube fiber electrodes with internal external construction for supercapacitors. J. Mater. Sci 2022, 57, 3613–3628. [Google Scholar] [CrossRef]
  127. Nguyen, D.N.; Gund, G.S.; Jung, M.G.; Roh, S.H.; Park, J.; Kim, J.K.; Park, H.S. Core-shell structured MXene@carbon nanodots as bifunctional catalysts for solar-assisted water splitting. ACS Nano 2020, 14, 17615–17625. [Google Scholar] [CrossRef]
  128. Cao, J.-M.; Zatovsky, I.V.; Gu, Z.-Y.; Yang, J.-L.; Zhao, X.-X.; Guo, J.-Z.; Xu, H.; Wu, X.-L. Two-dimensional MXene with multidimensional carbonaceous matrix: A platform for general-purpose functional materials. Prog. Mater. Sci. 2023, 135, 101105. [Google Scholar] [CrossRef]
  129. Amor, A.B.; Hemmami, H.; Amor, I.B.; Zeghoud, S.; Alhamad, A.A.; Belkacem, M.; Nair, N.S.; Sruthimol, A.B. Advances in carbon quantum dot applications:Catalysis, sensing, and biomedical innovations. Mater. Sci. Semicond. Process. 2025, 185, 108945. [Google Scholar] [CrossRef]
  130. Chen, H.-R.; Meng, W.-M.; Wang, R.-Y.; Chen, F.-L.; Li, T.; Wang, D.-D.; Wang, F.; Zhu, S.-E.; Wei, C.-X.; Lu, H.-D.; et al. Engineering highly graphitic carbon quantum dots by catalytic dehydrogenation and carbonization of Ti3C2Tx-MXene wrapped polystyrene spheres. Carbon 2022, 190, 319–328. [Google Scholar] [CrossRef]
  131. Dey, A.; Varagnolo, S.; Power, N.P.; Vangapally, N.; Elias, Y.; Damptey, L.; Jaato, B.N.; Gopalan, S.; Golrokhi, Z.; Sonar, P.; et al. Doped MXenes—A new paradigm in 2D systems: Synthesis, properties and applications. Prog. Mater. Sci. 2023, 139, 101166. [Google Scholar] [CrossRef]
  132. Hui, J.; Wu, R.; Zhu, Y.; Zhang, Z.; Wei, S.; Ouyang, F. Citric acid-assisted in situ preparation of MoIn2S4/CQDs with few-layer promotes charge transfer and enhances photocatalytic activity. Appl. Surf. Sci. 2024, 667, 160420. [Google Scholar] [CrossRef]
  133. Lim, G.P.; Soon, C.F.; Al-Gheethi, A.A.; Morsin, M.; Tee, K.S. Recent progress and new perspective of MXene-based membranes for water purification: A review. Ceram. Int. 2022, 48, 16477–16491. [Google Scholar] [CrossRef]
  134. Wei, X.; Akbar, M.U.; Raza, A.; Li, G. A Review on Bismuth Oxyhalide Based Materials for Photocatalysis. Nanoscale Adv. 2021, 3, 3353–3372. [Google Scholar] [CrossRef] [PubMed]
  135. Li, J.; Li, X.; Li, X. Carbon Quantum Dots Modified 3D Flower-like BiOCl Nanostructures with Enhanced Visible Light Photocatalytic Degradation of Rhodamine B. Discov. Mater. 2025, 5, 30. [Google Scholar] [CrossRef]
  136. Shi, Z.; Chen, W.; Hu, Y.; Zhang, F.; Wang, L.; Zhou, D.; Chen, X.; Meng, S. Boosting Visible-Light Photocatalytic Activity of BiOCl Nanosheets via Synergetic Effect of Oxygen Vacancy Engineering and Graphene Quantum Dots-Sensitization. Molecules 2024, 29, 1362. [Google Scholar] [CrossRef]
  137. Wang, H.; Zhang, X.; Zhu, H.; Xiang, G. Robust Bi-Anchoring Carbon Dot/BiOCl Sheet Heterojunction Photocatalysts toward Superior Photocatalytic Activity. Nanoscale 2024, 16, 12670–12679. [Google Scholar] [CrossRef]
  138. Lu, T.; Huang, H.; Lv, G.; Meng, Z.; Zhu, L. Aerogel-Derived Carbon Quantum Dots Modified BiOCl Nanocomposites with Augmented Oxygen Vacancies for Enhanced Photodegradation of Antibiotics. J. Ind. Eng. Chem. 2024, 138, 586–600. [Google Scholar] [CrossRef]
  139. Zhang, J.; Li, Z.; Lei, Q.; Zhong, D.; Ke, Y.; Liu, W.; Yang, L. Significantly Activated Persulfate by Novel Carbon Quantum Dots-Modified N-BiOCl for Complete Degradation of Bisphenol-A under Visible Light Irradiation. Sci. Total Environ. 2023, 870, 161804. [Google Scholar] [CrossRef]
  140. Cui, M.-J.; Jiang, J.-Z.; Song, Z.-H.; Ren, T.-Z. Enhanced Photocatalytic Degradation of Methylene Blue Using Carbon Dots-Modified Copper Chloride Hydroxide Nanocomposite. Inorg. Chem. Commun. 2023, 158, 111559. [Google Scholar] [CrossRef]
  141. Li, C.; Xu, Y.; Tu, W.; Chen, G.; Xu, R. Metal-Free Photocatalysts for Various Applications in Energy Conversion and Environmental Purification. Green Chem. 2017, 19, 882–899. [Google Scholar] [CrossRef]
  142. Li, Y.-Y.; Si, Y.; Zhou, B.-X.; Huang, T.; Huang, W.-Q.; Hu, W.; Pan, A.; Fan, X.; Huang, G.-F. Interfacial charge modulation: Carbon quantum dot implanted carbon nitride double-deck nanoframes for robust visible-light photocatalytic tetracycline degradation. Nanoscale 2020, 12, 3135. [Google Scholar] [CrossRef]
  143. Wang, F.; Chen, P.; Feng, Y.; Xie, Z.; Liu, Y.; Su, Y.; Zhang, Q.; Wang, Y.; Ya, K.; Lv, W.; et al. Facile synthesis of N-doped carbon dots/g-C3N4 photocatalyst with enhanced visible-light photocatalytic activity for the degradation of indomethacin. Appl. Catal. B 2017, 207, 103. [Google Scholar] [CrossRef]
  144. Li, B.; Fang, Q.; Si, Y.; Huang, T.; Huang, W.-Q.; Hu, W.; Pan, A.; Fan, X.; Huang, G.-F. Ultra-thin tubular graphitic carbon Nitride-Carbon Dot lateral heterostructures: One-Step synthesis highly efficient catalytic hydrogen generation. Chem. Eng. J. 2020, 397, 125470. [Google Scholar] [CrossRef]
  145. Duan, Y.; Deng, L.; Shi, Z.; Liu, X.; Zeng, H.; Zhang, H.; Crittenden, J. Efficient sulfadiazine degradation via in-situ epitaxial grow of Graphitic Carbon Nitride (g-C3N4) on carbon dots heterostructures under visible light irradiation: Synthesis mechanisms toxicity evaluation. J. Colloid Interface Sci. 2020, 561, 696. [Google Scholar] [CrossRef] [PubMed]
  146. Xu, T.F.; Wang, D.N.; Dong, L.L.; Shen, H.B.; Lu, W.Y.; Chen, W.X. Graphitic carbon nitride co-modified by zinc phthalocyanine and graphene quantum dots for the efficient photocatalytic degradation of refractory contaminants. Appl. Catal. B Environ. 2019, 244, 96–106. [Google Scholar] [CrossRef]
  147. Yuan, A.L.; Lei, H.; Xi, F.N.; Liu, J.Y.; Qin, L.S.; Chen, Z.; Dong, X.P. Graphene quantum dots decorated graphitic carbon nitride nanorods for photocatalytic removal of antibiotics. J. Colloid Interface Sci. 2019, 548, 56–65. [Google Scholar] [CrossRef]
  148. Jourshabani, M.; Long, N.V.D.; Asrami, M.R.; Pho, Q.H.; Lee, B.-K.; Hessel, V. Nitrogen-Doped Carbon Quantum Dot as Electron Acceptor Anchored on Graphitic Carbon Nitride Nanosheet for Improving Rhodamine B Degradation. Mater. Sci. Eng. B 2024, 305, 117417. [Google Scholar] [CrossRef]
  149. Han, H.; Wang, B.; Tang, Q.; Jia, S.; Liu, J.; Li, H.; Wang, C.; Xu, H.; Hua, Y. Non-Metallic Nitrogen-Doped Graphene Quantum Dots Coupled with g-C3N4 Achieve Efficient Photocatalytic Performance. Appl. Surf. Sci. 2024, 649, 159171. [Google Scholar] [CrossRef]
  150. Ishak, N.; Galář, P.; Mekkat, R.; Grandcolas, M.; Šoóš, M. Fine-Tuning Photoluminescence and Photocatalysis: Exploring the Effects of Carbon Quantum Dots Synthesis and Purification on g-C3N4. Colloids Surf. Physicochem. Eng. Asp. 2025, 706, 135789. [Google Scholar] [CrossRef]
  151. 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]
  152. Yang, P.; Shen, A.; Zhu, Z.; Wang, L.; Tang, R.; Yang, K.; Chen, M.; Dai, H.; Zhou, X. Construction of carbon nitride-based heterojunction as photocatalyst for peroxymonosulfate activation: Important role of carbon dots in enhancing photocatalytic activity. Chem. Eng. J. 2023, 464, 142724. [Google Scholar] [CrossRef]
  153. Yang, X.; Sun, J.; Sheng, L.; Wang, Z.; Ye, Y.; Zheng, J.; Fan, M.; Zhang, Y.; Sun, X. Carbon dots cooperatively modulating photocatalytic performance and surface charge of O-doped g-C3N4 for efficient water disinfection. J. Colloid Interface Sci. 2023, 631, 25–34. [Google Scholar] [CrossRef]
  154. Xu, Z.; Su, X.; Yang, P.; Zhong, J.; Li, M.; Burda, C.; Dou, L. Efficient Photocatalytic CO2 and Cr(VI) Reduction on Carbon Quantum Dots/Carbon Nitride Heterojunctions. Fuel 2025, 381, 133285. [Google Scholar] [CrossRef]
  155. Cheng, K.; Shao, W.; Li, H.; Guo, W.; Bian, H.; Han, J.; Wu, G.; Xing, W. Biomass Derived Carbon Dots Mediated Exciton Dissociation in Rose Flower-like Carbon Nitride for Boosting Photocatalytic Performance. Ind. Crops Prod. 2023, 192, 116086. [Google Scholar] [CrossRef]
  156. Zhang, Y.; Yuan, J.; Ding, Y.; Zhang, B.; Zhang, S.; Liu, B. Metal-Free N-GQDs/P-g-C3N4 Photocatalyst with Broad-Spectrum Response: Enhanced Exciton Dissociation and Charge Migration for Promoting H2 Evolution and Tetracycline Degradation. Sep. Purif. Technol. 2023, 304, 122297. [Google Scholar] [CrossRef]
  157. Hu, J.; Chen, C.; Tan, C.; Fan, H.; Lu, J.; Li, Y.; Hu, H. Novel In-Situ Composites of Chlorine-Doped CQDs and g-C3N4 with Improved Interfacial Connectivity and Accelerated Charge Transfer to Enhance Photocatalytic Degradation of Tetracycline and Hydrogen Evolution. Appl. Surf. Sci. 2023, 638, 158060. [Google Scholar] [CrossRef]
  158. Aygun, A.; Tiri, R.N.E.; Bayat, R.; Sen, F. Hydrothermal Synthesis of BCQD@g-C3N4 Nanocomposites Supporting Environmental Sustainability: Organic Dye Removal and Bacterial Inactivation. J. Hazard. Mater. Adv. 2024, 16, 100464. [Google Scholar] [CrossRef]
  159. Sunil, S.; Mandal, B.K. Facile Synthesis of CQD/g-C3N4 as a Highly Effective Metal-Free Photocatalyst for the Degradation of Carmoisine and Indigo Carmine Dye. Inorg. Chem. Commun. 2025, 171, 113545. [Google Scholar] [CrossRef]
  160. Xu, N.; Liu, S.; Xu, Q.; Yuan, P.; Zhang, P.; Zhuo, S.; Zhu, C.; Du, J. Facile Synthesis of Porous Graphitic Carbon Nitride Modulated by Up-Conversion Carbon Quantum Dots for Visible Light-Triggered Photocatalysis towards Bacteria Inactivation. Appl. Catal. Gen. 2024, 673, 119586. [Google Scholar] [CrossRef]
  161. Wang, Y.; Bai, Y.; Han, C.; Li, Z.; Lun, X.; Zhang, C. Photocatalysis-PMS Oxidation System Based on CQDs-Doped Carbon Nitride Nanosheets for Degradation of Residual Drugs in Water. Environ. Sci. Pollut. Res. 2023, 30, 108538–108552. [Google Scholar] [CrossRef]
  162. Chen, R.; Hu, R.; Gao, L.; Wei, H.; Zhao, S.; Liu, X.; Li, X.; Xu, A. Synergistic Photocatalytic Activation of Permanganate by Carbon Quantum Dots-Doped g-C3N4 under Visible Light Irradiation. J. Environ. Chem. Eng. 2025, 13, 117162. [Google Scholar] [CrossRef]
  163. Guan, J.; Liu, X.; Bai, N.; Wang, F.; Yang, Z.; Zhang, J.; Gao, F.; Zhang, P.; Wei, Z. Luminescence Properties of CQDs and Photocatalytic Properties of TiO2/ZnO/CQDs Ternary Composites. J. Mater. Sci. Mater. Electron. 2023, 34, 2169. [Google Scholar] [CrossRef]
  164. Hao, L.; Liang, Z.; Yu, Y.; Houy, H.; Min, D. Nano-island-like heterojunction of nitrogen-doped carbon quantum dots anchoring on titanium oxide nano sheets@Cu2O for enhancing visible light-driving photocatalytic degradation of methyl orange. J. Water Process Eng. 2024, 57, 104675. [Google Scholar] [CrossRef]
  165. Dou, S.; Wang, D.; Shang, Q.; Kong, X.; Fang, Y. Carbon Dots Modified Dendritic TiO2-CdS Heterojunction for Enhanced Photodegradation of Rhodamine and Hydrogen Evolution. Diam. Relat. Mater. 2023, 137, 110115. [Google Scholar] [CrossRef]
  166. Shi, W.; Hao, C.; Shi, Y.; Guo, F.; Tang, Y. Effect of different carbon dots positions on the transfer of photo-induced charges in type I heterojunction for significantly enhanced photocatalytic activity. Sep. Purif. Technol. 2023, 304, 122337. [Google Scholar] [CrossRef]
  167. Wang, G.; Li, X. Hao P, Liu W, Zhan H, Bi S, g-C3N4/Nitrogen-Doped Carbon Dot/Silver Nanoparticle-Based Ternary Photocatalyst for Water Pollutant Treatment. ACS Appl. Nano Mater. 2023, 6, 5747–5758. [Google Scholar] [CrossRef]
  168. Abdurahman, M.-H.; Zuhairi Abdullah, A.; Da Oh, W.; Fazliani Shopware, N.; Faisal Gasim, M.; Okoye, P.; Ul-Hamid, A.; Rahman Mohamed, A. Tunable band structure of synthesized carbon dots modified graphitic carbon nitride/bismuth oxychlorobromide heterojunction for photocatalytic degradation of tetracycline in water. J. Colloid Interface Sci. 2023, 629, 189–205. [Google Scholar] [CrossRef]
  169. Zhu, L.; Shen, D.; Zhang, H.; Hong Luo, K.; Li, C. Fabrication of Z-scheme Bi7O9I3/g-C3N4 heterojunction modified by carbon quantum dots for synchronous photocatalytic removal of Cr (Ⅵ) and organic pollutants. J. Hazard. Mater. 2023, 446, 130663. [Google Scholar] [CrossRef]
  170. Tao, Y.; Mou, Z.; Lei, W.; Li, Y.; Shangguan, L.; Zhu, W.; Gong, J. Improving adsorption and visible light photocatalytic performance of TiO2 via synergistic effect of nitrogen-doped graphene quantum dots and reduced mildly oxidized graphene oxide for cationic dye pollutants removal. Surf. Interfaces 2024, 46, 104150. [Google Scholar] [CrossRef]
  171. Joulaee, S.; Mirzaei, M.; Hassanpour, A.; Safardoust-Hojaghan, H.; Khani, A. Efficient removal of anionic and cationic dyes from waste water using green ZnO/NiO/graphene quantum dots nano photocatalyst. Opt.-Int. J. Light Electron Opt. 2023, 290, 171324. [Google Scholar] [CrossRef]
  172. 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]
  173. Guo, Q.; Li, X.; Liu, X.; Wen, M.; Wang, G.; Zhan, H.; Chen, X.; Li, H.; Ma, J.; Liu, W. Bifunctional Nitrogen-Doped Carbon Quantum Dot-Modified Graphitic Carbon Nitride/Cadmium Sulfide Nanomaterials for Antibiotic Selective Detection and Photocatalytic Degradation. Mater. Today Chem. 2024, 38, 102123. [Google Scholar] [CrossRef]
  174. Li, J.; Cheng, X.; Zhang, Q.; Zhang, L.; Qi, Z. Fabrication of Z-Scheme Heterojunctions of CDs@WO3/g-C3N4 Nanocomposite Photocatalyst with Enhanced Visible-Light Photocatalytic Degradation of Malachite Green. J. Mater. Sci. Mater. Electron. 2024, 35, 598. [Google Scholar] [CrossRef]
  175. Jiang, R.; Zhong, D.; Xu, Y.; Chang, H.; He, Y.; Zhang, J.; Liao, P. Chitosan Derived N-Doped Carbon Anchored Co3O4-Doped MoS2 Nanosheets as an Efficient Peroxymonosulfate Activator for Degradation of Dyes. Int. J. Biol. Macromol. 2024, 265, 130519. [Google Scholar] [CrossRef]
  176. Teymourinia, H.; Rtimi, S.; Ghalkhani, M.; Ramazani, A.; Aminabhavi, T.M. Flower-like Nanocomposite of Carbon Quantum Dots, MoS2, and Dendritic Ag-Based Z-Scheme Type Photocatalysts for Effective Tartrazine Degradation. Chem. Eng. J. 2023, 473, 145239. [Google Scholar] [CrossRef]
  177. Bai, S.; Lv, T.; Chen, M.; Li, C.; Wang, Z.; Yang, X.; Xia, T. Carbon Quantum Dots Assisted BiFeO3@BiOBr S-Scheme Heterojunction Enhanced Peroxymonosulfate Activation for the Photocatalytic Degradation of Imidacloprid under Visible Light: Performance, Mechanism and Biotoxicity. Sci. Total Environ. 2024, 915, 170029. [Google Scholar] [CrossRef]
  178. Chen, J.; Qin, S.; Yang, X.; Wang, Y.; Yang, T.; Que, M.; Ma, Y.; Li, Y. Synthesis of Highly Conductive Carbon Quantum Dot-Enhanced Z-Scheme BiOBr/g-C3N4 Heterojunction for Effective Photocatalytic Degradation of Tetracycline Hydrochloride. J. Phys. Chem. Solids 2024, 189, 111957. [Google Scholar] [CrossRef]
  179. Wang, L.; Sun, J.; Shi, J.; Huang, T.; Liu, K.; Tong, Z.; Zhang, H. Reinforced Built-in Electric Field and Mediated Schottky Barrier Height via Carbon Quantum Dots Electronic Bridges on BiOBr/Ti3C2 for Efficient Photocatalytic Quinolone Antibiotics Degradation. Chem. Eng. J. 2024, 500, 157168. [Google Scholar] [CrossRef]
  180. Tie, W.; Bhattacharyya, S.S.; Ma, T.; Yuan, S.; Chen, M.; He, W.; Lee, S.H. Improving Photoexcited Carrier Separation through Z-Scheme W18O49/BiOBr Heterostructure Coupling Carbon Quantum Dots for Efficient Photoelectric Response and Tetracycline Photodegradation. Carbon 2025, 231, 119707. [Google Scholar] [CrossRef]
  181. Xu, D.; Yu, C.; Peng, X.; Yan, H.; Zhang, Y. CQDs Modified Bi2MoO6/CuS p–n Heterojunction Photocatalytic Efficient Degradation of Tetracycline. Res. Chem. Intermed. 2024, 50, 2477–2499. [Google Scholar] [CrossRef]
  182. Yuan, J.; Zhang, Y.; Zhang, X.; Zhang, J.; Zhang, S. N-Doped Graphene Quantum Dot-Decorated N-TiO2/P-Doped Porous Hollow g-C3N4 Nanotube Composite Photocatalysts for Antibiotic Photodegradation and H2 Production. Int. J. Miner. Metall. Mater. 2024, 31, 165–178. [Google Scholar] [CrossRef]
  183. Yuan, S.; Yin, G.; Zhao, T.; Zhang, J.; Wei, S.; Zhang, H.; Liu, Z.; Zhang, J.; Lu, Q.; Sun, M. Preparation and Photocatalytic Degradation Properties of Z-Scheme Si-TiO2/g-C3N4 Heterojunction Modified with F-CDs. Solid State Sci. 2025, 160, 107832. [Google Scholar] [CrossRef]
  184. Shan, C.; Liu, Z.; Li, F.; Jia, W.; Cai, G.; Li, M.; Tang, T.; Li, S.; Wen, J.; Hu, G.; et al. Ternary TiO2/P-GQDs/AgI Nanocomposites with n-p-n Heterojunctions for Enhanced Visible Photocatalysis. J. Nanoparticle Res. 2023, 25, 128. [Google Scholar] [CrossRef]
  185. Zhang, J.; Shao, C.; Zhang, A.; Zhang, Y.; Zhang, L.; Bai, H. Nitrogen-Doped Carbon Quantum Dots Modified Dual-Vacancy Z-Scheme CuFe2O4/g-C3N4 Photocatalyst Synergistic with Fenton Technique for Photothermal Degradation of Antibiotics. Chem. Eng. J. 2024, 497, 154497. [Google Scholar] [CrossRef]
Figure 1. Water pollutants degradation mechanism over a heterogeneous photocatalyst. (Adopted from [14] with permission).
Figure 1. Water pollutants degradation mechanism over a heterogeneous photocatalyst. (Adopted from [14] with permission).
Inorganics 13 00286 g001
Figure 2. (a) CDs consisting of a carbon core and a rich array of chemical functionalities. (Adopted from [42] with permission). (b) Classification of CDs based on their structure. (Adopted from [45] with permission).
Figure 2. (a) CDs consisting of a carbon core and a rich array of chemical functionalities. (Adopted from [42] with permission). (b) Classification of CDs based on their structure. (Adopted from [45] with permission).
Inorganics 13 00286 g002
Figure 4. Synthetic strategies for CDs-based heterostructures: ex situ vs. in situ approaches.
Figure 4. Synthetic strategies for CDs-based heterostructures: ex situ vs. in situ approaches.
Inorganics 13 00286 g004
Figure 5. (a) Band gap energies of several metal oxides [26]. (b) Number of publications in Scopus between 2015 and 2025 on the topic “carbon dots” and “metal oxides”.
Figure 5. (a) Band gap energies of several metal oxides [26]. (b) Number of publications in Scopus between 2015 and 2025 on the topic “carbon dots” and “metal oxides”.
Inorganics 13 00286 g005
Figure 6. (a) Schematic representation of TiO2 decoration with CDs prepared by peanut-shells. (b) Photocatalytic performance of TiO2 and CDs–TiO2 hybrid materials, containing different amounts of CDs, in the reduction of Cr (VI). (c) The as-proposed photocatalytic mechanism. (Adopted from [61] with permission).
Figure 6. (a) Schematic representation of TiO2 decoration with CDs prepared by peanut-shells. (b) Photocatalytic performance of TiO2 and CDs–TiO2 hybrid materials, containing different amounts of CDs, in the reduction of Cr (VI). (c) The as-proposed photocatalytic mechanism. (Adopted from [61] with permission).
Inorganics 13 00286 g006
Figure 7. (a) Synthesis procedure of CD–TiO2 NT photocatalytic material. (b) Reduction of the band gap of TiO2 after CD modification. (Adopted from [62] with permission).
Figure 7. (a) Synthesis procedure of CD–TiO2 NT photocatalytic material. (b) Reduction of the band gap of TiO2 after CD modification. (Adopted from [62] with permission).
Inorganics 13 00286 g007
Figure 8. (a) Zero point of charge for N–CQDs–ZnO hybrid material. Impact of (b) pH value, (c) photocatalyst concentration, and (d) initial tetracycline concentration in the photocatalytic performance of the N–CQDs–ZnO heterostructure. In all cases the lowest values are represented by red color. (Adopted from [71] with permission).
Figure 8. (a) Zero point of charge for N–CQDs–ZnO hybrid material. Impact of (b) pH value, (c) photocatalyst concentration, and (d) initial tetracycline concentration in the photocatalytic performance of the N–CQDs–ZnO heterostructure. In all cases the lowest values are represented by red color. (Adopted from [71] with permission).
Inorganics 13 00286 g008
Figure 9. Number of publications in Scopus between 2015 and 2025 on the topic “carbon dots” and “metal sulfides”.
Figure 9. Number of publications in Scopus between 2015 and 2025 on the topic “carbon dots” and “metal sulfides”.
Inorganics 13 00286 g009
Figure 10. (a) Combination of CdS with CDs prepared by neem leaves. (b) CPX degradation of CDs, CdS and CDs–CdS hybrid materials. (Adopted from [101] with permission). (c,d) Morphological characterization of Cu-doped CDs–CdS via SEM and TEM analysis, respectively. (Cu-doped CDs are visible in the area highlighted in red). (e) Degradation of TC in different reaction systems, with and without catalysts. (f) The proposed mechanism of TC photocatalytic degradation by Cu-doped CDs–CdS. (g) Photocurrent response of CuS, CDs–CdS and Cu-doped CDs–CdS. (Adopted from [103] with permission).
Figure 10. (a) Combination of CdS with CDs prepared by neem leaves. (b) CPX degradation of CDs, CdS and CDs–CdS hybrid materials. (Adopted from [101] with permission). (c,d) Morphological characterization of Cu-doped CDs–CdS via SEM and TEM analysis, respectively. (Cu-doped CDs are visible in the area highlighted in red). (e) Degradation of TC in different reaction systems, with and without catalysts. (f) The proposed mechanism of TC photocatalytic degradation by Cu-doped CDs–CdS. (g) Photocurrent response of CuS, CDs–CdS and Cu-doped CDs–CdS. (Adopted from [103] with permission).
Inorganics 13 00286 g010aInorganics 13 00286 g010b
Figure 11. (a) Hydrothermal synthesis of N and S co-doped CDs–MoS2 prepared by Qu et al. [104] (b,c) The photocatalytic performance of MoS2 and N and S co-doped CDs–MoS2 containing different mass ratios in the degradation of MB and MG, respectively. (Adopted from [104] with permission). (d) TEM image of N-doped CDs–MoS2 prepared by Zhang et al. [37] (e,f) Degradation of TC by N-doped CDs–MoS2 and PMS under visible light irradiation and synergistic visible light irradiation and ultrasonic vibration, respectively. (Adopted from [37] with permission).
Figure 11. (a) Hydrothermal synthesis of N and S co-doped CDs–MoS2 prepared by Qu et al. [104] (b,c) The photocatalytic performance of MoS2 and N and S co-doped CDs–MoS2 containing different mass ratios in the degradation of MB and MG, respectively. (Adopted from [104] with permission). (d) TEM image of N-doped CDs–MoS2 prepared by Zhang et al. [37] (e,f) Degradation of TC by N-doped CDs–MoS2 and PMS under visible light irradiation and synergistic visible light irradiation and ultrasonic vibration, respectively. (Adopted from [37] with permission).
Inorganics 13 00286 g011aInorganics 13 00286 g011b
Figure 12. (a) Decoration of WS2–PANI with CDs via ultrasonication. (b) TEM image of the as-prepared hybrid material. (c,d) Photocatalytic degradation of pharmaceutical compounds EST and NFT, respectively, using WS2 (black), WS2–PANI (red), CDs–WS2–PANI (blue), CDs–WS2 (dark red) and CDs–PANI (green) as photocatalytic materials [106].
Figure 12. (a) Decoration of WS2–PANI with CDs via ultrasonication. (b) TEM image of the as-prepared hybrid material. (c,d) Photocatalytic degradation of pharmaceutical compounds EST and NFT, respectively, using WS2 (black), WS2–PANI (red), CDs–WS2–PANI (blue), CDs–WS2 (dark red) and CDs–PANI (green) as photocatalytic materials [106].
Inorganics 13 00286 g012
Figure 13. (a) The photocatalytic performance of H2O2 production using different samples, including the N-doped CDs–ZnIn2S4 heterostructure. (b,c) Photocatalytic application of self-Fenton system of Fe2+/N-doped CDs–ZnIn2S4 in the degradation of antibiotics LEV and AMX, respectively. (Adopted from [110] with permission).
Figure 13. (a) The photocatalytic performance of H2O2 production using different samples, including the N-doped CDs–ZnIn2S4 heterostructure. (b,c) Photocatalytic application of self-Fenton system of Fe2+/N-doped CDs–ZnIn2S4 in the degradation of antibiotics LEV and AMX, respectively. (Adopted from [110] with permission).
Inorganics 13 00286 g013
Figure 15. Preparation route of CQDs via a sacrificial template method (schematically). (Adopted from [130] with permission).
Figure 15. Preparation route of CQDs via a sacrificial template method (schematically). (Adopted from [130] with permission).
Inorganics 13 00286 g015
Figure 16. (a,b) SEM and (c,d) TEM images of CDs-modified 3-D flower-like BiOCl structures, presenting improved (e) RhB photocatalytic degradation and (f) photocurrent density, compared to pristine BiOCl [135]. (g) Photocatalytic performance of BiOCl and GQDs–BiOCl containing varying GQDs loading in the degradation of RhB under visible-light [136]. (h) CDs promote OV formation in BiOCl, resulting in improved photocatalytic performance. (i) ESR results of BiOCl and CDs–BiOCl. (Adopted from [138] with permission). (j) EIS Nyquist plots of BiOCl, N-doped BiOCl and CDs–N-doped BiOCl. (Adopted from [139] with permission). (k) UV–Vis spectra of Cu2Cl(OH)3 and CDs–Cu2Cl(OH)3 hybrids, containing various amounts of CDs. (l) Photodegradation of MB dye and (m) cycling stability tests using the optimized heterostructure. (Adopted from [140] with permission).
Figure 16. (a,b) SEM and (c,d) TEM images of CDs-modified 3-D flower-like BiOCl structures, presenting improved (e) RhB photocatalytic degradation and (f) photocurrent density, compared to pristine BiOCl [135]. (g) Photocatalytic performance of BiOCl and GQDs–BiOCl containing varying GQDs loading in the degradation of RhB under visible-light [136]. (h) CDs promote OV formation in BiOCl, resulting in improved photocatalytic performance. (i) ESR results of BiOCl and CDs–BiOCl. (Adopted from [138] with permission). (j) EIS Nyquist plots of BiOCl, N-doped BiOCl and CDs–N-doped BiOCl. (Adopted from [139] with permission). (k) UV–Vis spectra of Cu2Cl(OH)3 and CDs–Cu2Cl(OH)3 hybrids, containing various amounts of CDs. (l) Photodegradation of MB dye and (m) cycling stability tests using the optimized heterostructure. (Adopted from [140] with permission).
Inorganics 13 00286 g016aInorganics 13 00286 g016b
Figure 17. (a) Synthesis procedure of a N-doped GQDs–g–C3N4 hybrid material. (b) Photocatalytic degradation of RhB using pristine g–C3N4 and N-doped GQDs–g–C3N4, containing various N-doped GQDs concentrations. (c) The apparent rate constants. (Adopted by [149] with permission). (d) Purification process of CDs via column chromatography. (Adopted by [150] with permission).
Figure 17. (a) Synthesis procedure of a N-doped GQDs–g–C3N4 hybrid material. (b) Photocatalytic degradation of RhB using pristine g–C3N4 and N-doped GQDs–g–C3N4, containing various N-doped GQDs concentrations. (c) The apparent rate constants. (Adopted by [149] with permission). (d) Purification process of CDs via column chromatography. (Adopted by [150] with permission).
Inorganics 13 00286 g017aInorganics 13 00286 g017b
Figure 18. Study of the morphology, structure, and chemical properties of CDs. (ac) TEM images of CDs rich in pyrrolic-N, rich in graphitic-N, and rich in carboxyl groups (N-free), respectively, (d) particle size distributions, (e) XRD patterns, and (f) FT-IR spectra of all the synthesized CDs samples. (Adopted from [152] with permission).
Figure 18. Study of the morphology, structure, and chemical properties of CDs. (ac) TEM images of CDs rich in pyrrolic-N, rich in graphitic-N, and rich in carboxyl groups (N-free), respectively, (d) particle size distributions, (e) XRD patterns, and (f) FT-IR spectra of all the synthesized CDs samples. (Adopted from [152] with permission).
Inorganics 13 00286 g018
Figure 19. (a) Utilization of a CDs–g–C3N4 heterostructure in water disinfection from bacteria. (b) Inactivation efficiency of bulk g–C3N4, O-doped g–C3N4, and CDs–O-doped g–C3N4 photocatalysts toward MRSA under visible-light irradiation. (c) Fluorescent images of live (green)/dead (red) MRSA cells. (Adopted by [153] with permission). (d) Degradation kinetic curves of Cr (VI) under visible light irradiation and (e) the pseudo-first-order kinetics curves of Cr(VI) reduction, using both pristine g–C3N4 and CQDs–g–C3N4 hybrid material. (Adopted by [154] with permission).
Figure 19. (a) Utilization of a CDs–g–C3N4 heterostructure in water disinfection from bacteria. (b) Inactivation efficiency of bulk g–C3N4, O-doped g–C3N4, and CDs–O-doped g–C3N4 photocatalysts toward MRSA under visible-light irradiation. (c) Fluorescent images of live (green)/dead (red) MRSA cells. (Adopted by [153] with permission). (d) Degradation kinetic curves of Cr (VI) under visible light irradiation and (e) the pseudo-first-order kinetics curves of Cr(VI) reduction, using both pristine g–C3N4 and CQDs–g–C3N4 hybrid material. (Adopted by [154] with permission).
Inorganics 13 00286 g019
Figure 20. (a) Schematic representation of the preparation process of CDs–TiO2–CdS ternary heterostructure and (b) the as-proposed photocatalytic mechanism. (Adopted by [165] with permission) (c) Different CDs positions in the ZnFe2O4–ZnIn2S4 core–shell structure and its effect on charge transfer. (Adopted by [166] with permission).
Figure 20. (a) Schematic representation of the preparation process of CDs–TiO2–CdS ternary heterostructure and (b) the as-proposed photocatalytic mechanism. (Adopted by [165] with permission) (c) Different CDs positions in the ZnFe2O4–ZnIn2S4 core–shell structure and its effect on charge transfer. (Adopted by [166] with permission).
Inorganics 13 00286 g020
Table 1. Recent advances on photocatalytic water treatment using CDs–metal oxide heterostructures.
Table 1. Recent advances on photocatalytic water treatment using CDs–metal oxide heterostructures.
CDs–Metal OxidesWater Pollutant Light SourcePhotocatalytic
Degradation
References
N-doped CQDs–TiO2Methylene Blue
(10 ppm)
300 W Xenon lamp93.10% after 60 min
(10 mg/100 mL photocatalyst)
[74]
CQDs–TiO2Methyl Orange
(25 ppm)
300 W Xenon lamp
>400 nm
85.00% after 130 min
(20 mg/50 mL photocatalyst)
[75]
GQDs–ZnOMetronidazole
(100 ppm)
UV>99.99% after 30 min
(100 mg/100 mL photocatalyst)
[76]
CQDs–Zn2SnO4Methylene Blue
(3 ppm)
Natural Sunlight95.00% after 150 min
(200 mg/200 mL photocatalyst)
[77]
N-doped CDs–Cu2OMethylene Blue
(10 ppm)
500 W Xenon lamp96.40% after 120 min
*
[78]
CQDs–Cu2OTetracycline
(10 ppm)
Xenon lamp
Vis
92.49% after 100 min
(100 mg/90 mL photocatalyst)
[79]
B-doped CQDs–Bi2MoO6Bisphenol A
(20 ppm)
300 W Xenon lamp100% after 120 min
(20 mg/60 mL photocatalyst)
[80]
GQDs–Bi2MoO6Bisphenol A (20 ppm)300 W Xenon lamp
Vis
95.00% after 120 min
(50 mg/60 mL photocatalyst)
[81]
Sargassum horneri-derived CDs–Bi2MoO6Ciprofloxacin (20 ppm)300 W Xenon lamp97.70% after 180 min
(50 mg/100 mL photocatalyst)
[82]
CQDs–BiVO4Benzyl Paraben (10 ppm)300 W Xenon lamp
>420 nm
85.40% after 150 min
(100 mg/100 mL photocatalyst)
[83]
CQDs–Co2VO4Ciprofloxacin (20 ppm)500 W Halogen lamp
Vis
80.75% after 13 min
(50 mg/L photocatalyst and peroxymonosulphate)
[84]
CDs–WO3Methylene Blue (10 ppm)
Malachite Green (10 ppm)
10 W Xenon lamp 400–800 nm87.00% after 120 min
88.04% after 120 min
(10 mg photocatalyst)
[85]
N-doped GQDs–CuWO4Tetracycline (15 ppm)300 W Xenon lamp
300–800 nm
99.00% after 90 min
(10 mg photocatalyst)
[86]
GQDs–CuWO4Phenol (100 ppm)5 W LED light53.41% after 90 min
(200 mg/100 mL photocatalyst)
[87]
* not mentioned.
Table 2. Recent advances on photocatalytic water treatment using magnetic CDs–ferrites heterostructures.
Table 2. Recent advances on photocatalytic water treatment using magnetic CDs–ferrites heterostructures.
CDs–Ferrites Water PollutantLight SourcePhotocatalytic Degradation References
crocus cancellatus-derived CDs–Fe3O4Fluorescein
(45 μΜ)
Natural Sunlight94.4% after 60 min
(15 mg of photocatalyst and 0.05 M H2O2)
[89]
lemon peel-derived
CDs–Fe3O4
Methylene Blue
(50 ppm)
99.24% after 30 s
(100 mg of photocatalyst and 0.045 M H2O2)
[90]
watermelon peel derived CDs–Fe3O4Methylene Blue
(7 ppm)
UV98.00% after 30 min
(50 mg of photocatalyst)
[91]
yerba mate derived
CDs–Fe3O4
Methyl Orange
(8.5 ppm)
>400 nm97.70% after 7 h
(100 ppm photocatalyst and 150 mM H2O2)
[92]
mushrooms derived
CDs–ZnFe2O4
Methylene Blue
(10 ppm)
RhB
(10 ppm)
200 W Xenon lamp
Vis
94.25% after 30 min
96.10% after 60 min
(10 mg/L photocatalyst)
[93]
N-doped CDs–Zn-doped CoFe2O4Oxytetracycline
(10 ppm)
250 W HPMVL lamp
Vis
98.00% after 100 min
(200 mg/L photocatalyst)
[94]
GQDs–CoFe2O4Methylene Blue
(10 ppm)
160 W bulb~90% after 120 min
(50 mg/100 mL photocatalyst)
[95]
Boerhavia diffusa derived CDs–CoFe2O4Tetracycline
(50 ppm)
150 W Xenon lamp 92% after 120 min
(50 mg/100 mL photocatalyst and H2O2)
[96]
N-doped CDs–CoFe2O4Naproxen
(10 ppm)
Vis89.50% after 100 min
(20 mg/100 mL photocatalyst)
[97]
mango peel derived CDs–Co0,5Zn0,5Fe2O4Reactive Blue 222
(50 ppm)
Reactive Yellow 145 (50 ppm)
500 W Halogen lamp
Vis
~95% after 25 min
(1 g/L photocatalyst)
[98]
CDs–CuFe2O42-Nitroaniline
(200 ppm)
4-Nitroaniline
(200 ppm)
*96.70% after 45 s
96.50% after 15 s
(7 mg/5 mL photocatalyst and NaBH4)
[99]
* not mentioned.
Table 3. Recent advances on photocatalytic water treatment using CDs–graphitic C3N4 heterostructures.
Table 3. Recent advances on photocatalytic water treatment using CDs–graphitic C3N4 heterostructures.
CD–g–C3N4Water PollutantLight SourcePhotocatalytic Degradation References
ginkgo leaves-derived CDs–g–C3N4Rhodamine B
(10 ppm)
Vis>99.99% after 60 min
(* photocatalyst powder/50 mL)
[155]
N-doped GQDs–
P-doped g–C3N4
Tetracycline
(20 ppm)
>420 nm89.90% after 60 min
(50 mg/100 mL photocatalyst)
[156]
Cl-doped CQDs–
g–C3N4
91.70% after 120 min
(40 mg/40 mL photocatalyst)
[157]
B-doped CQDs–
g–C3N4
Rhodamine B
(1 g/L)
Methyl orange
(1 g/L)
Vis65.58% after 120 min
73.56% after 120 min
(10 mg/100 mL photocatalyst)
[158]
CQDs–g–C3N4Indigo Carmine
(10 ppm)
Carmoisine
(10 ppm)
UV and Vis96.00% after 60 min
93.50% after 60 min
(1.0 mg/mL photocatalyst)
[159]
E. coli (200 ppm)
S. aureu (300 ppm)
300 W Xenon lamp
>420 nm
** MIC of 12.5 ppm
MIC of 40 ppm
[160]
Meloxicam
(10 ppm)
Tetracycline
(10 ppm)
Vis99.90% after 30 min
95.97% after 45 min
(20 mg photocatalyst and 20 mg peroxymonosulfate/50 mL)
[161]
Sulfadiazine
(40 µM)
300 W Xenon lamp
>400 nm
98.00% after 30 min
(20 mg photocatalyst and
KMnO4/50 mL)
[162]
* not mentioned, ** MIC: minimum inhibitory concentration.
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

Zourou, A.; Ntziouni, A.; Karagianni, A.; Alizadeh, N.; Argirusis, N.; Antoniadou, M.; Sourkouni, G.; Kordatos, K.V.; Argirusis, C. Recent Advances in Carbon Dots-Based Photocatalysts for Water Treatment Applications. Inorganics 2025, 13, 286. https://doi.org/10.3390/inorganics13090286

AMA Style

Zourou A, Ntziouni A, Karagianni A, Alizadeh N, Argirusis N, Antoniadou M, Sourkouni G, Kordatos KV, Argirusis C. Recent Advances in Carbon Dots-Based Photocatalysts for Water Treatment Applications. Inorganics. 2025; 13(9):286. https://doi.org/10.3390/inorganics13090286

Chicago/Turabian Style

Zourou, Adamantia, Afrodite Ntziouni, Alexandra Karagianni, Niyaz Alizadeh, Nikolaos Argirusis, Maria Antoniadou, Georgia Sourkouni, Konstantinos V. Kordatos, and Christos Argirusis. 2025. "Recent Advances in Carbon Dots-Based Photocatalysts for Water Treatment Applications" Inorganics 13, no. 9: 286. https://doi.org/10.3390/inorganics13090286

APA Style

Zourou, A., Ntziouni, A., Karagianni, A., Alizadeh, N., Argirusis, N., Antoniadou, M., Sourkouni, G., Kordatos, K. V., & Argirusis, C. (2025). Recent Advances in Carbon Dots-Based Photocatalysts for Water Treatment Applications. Inorganics, 13(9), 286. https://doi.org/10.3390/inorganics13090286

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