Next Article in Journal
High-Performance Self-Driven SnSe/Si Heterojunction Photovoltaic Photodetector
Previous Article in Journal
Bi2WO6@g-C3N4 Heterostructure for Cathodic Photoelectrochemical Dopamine Sensor
Previous Article in Special Issue
Organic Luminescent Sensor for Mercury(II) and Iron(III) Ions in Aqueous Solutions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Fluorescent Quantum Dots and Its Composites for Highly Sensitive Detection of Heavy Metal Ions and Pesticide Residues: A Review

1
National Center for International Research on Photoelectric and Energy Materials, School of Materials and Energy, Yunnan University, Kunming 650504, China
2
Yunnan Key Laboratory of Carbon Neutrality and Green Low-Carbon Technologies, Yunnan University, Kunming 650504, China
*
Author to whom correspondence should be addressed.
Chemosensors 2023, 11(7), 405; https://doi.org/10.3390/chemosensors11070405
Submission received: 26 May 2023 / Revised: 8 July 2023 / Accepted: 10 July 2023 / Published: 19 July 2023
(This article belongs to the Special Issue Advances in Nanocomposite Luminescent Sensors)

Abstract

:
Quantum dots nanomaterials have attracted extensive interest for fluorescence chemical sensors due their attributes, such as excellent optical characteristics, quantum size effects, interface effects, etc. Moreover, the fluorescence properties of quantum dots can be adjusted by changing their structure, size, morphology, composition, doping, and surface modification. In recent years, quantum dots nanomaterials have been considered the preferred sensing materials for the detection of heavy metal ions and pesticide residues by the interactions between quantum dots and various analytes, showing excellent sensitivity, selectivity, and interference, as well as reducing the cost of equipment compared with traditional measurement methods. In this review, the applications and sensing mechanisms of semiconductor quantum dots and carbon-based quantum dots are comprehensively discussed. The application of semiconductor quantum dots, carbon quantum dots, graphene quantum dots, and their nanocomposites that are utilized as fluorescence sensors are discussed in detailed, and the properties of various quantum dots for heavy metal ion and pesticide residue determination are also presented. The recent advances in and application perspectives regarding quantum dots and their composites are also summarized.

1. Introduction

Environmental pollution and food safety are becoming increasingly prominent issues with rapid economic and industrial development and population growth in recent years [1,2]. Mineral exploitation causes heavy metal pollution, while pesticide abuse leads to excessive pesticide residues. Due to their characteristics of non-degradability and long-term accumulation, heavy metal ions and excessive pesticide residues can cause severe damage to the human body through the food chain, including food poisoning, diarrhea, neurotoxicity, brain damage, renal injury, skeletal disorders, and even health issues for children and fetuses [3,4,5,6,7]. Hence, the detection and monitoring of heavy metal ions and pesticide residues are important measures to ensure safety to human lives and environmental protection. Moreover, the maximum concentration levels for some heavy metal ions and pesticide residues have been specified by many international organizations, such as the World Health Organization (WHO), European Union (EU), Codex Alimentarius Commission (CAC), and the United States Environmental Protection Agency (USEPA) [4,8]. Consequently, there is a growing need for simple, rapid, and accurate detection of heavy metal ions and pesticides in water and food.
Traditional detection methods, including atomic absorption spectrometry, surface plasmon resonance, atomic emission spectrometry, inductively coupled plasma, mass spectrometer, high performance liquid chromatography, colorimetric method, and surface-enhanced Raman scattering, can obtain relatively stable and accurate detection results, but the equipment and maintenance costs are high, and the sample pretreatment process is complicated, so these methods are still mainly used in the laboratory [9,10,11,12,13,14]. Fluorescence detection technology is an emerging technology from recent decades, which has the characteristics of cost-effectiveness, rapidity, simple operation, facile and easy detection techniques, and low detection limits. Among numerous fluorescent materials, QDs as zero dimensional nanomaterials offer greater potential and have been widely recognized as valuable fluorescent probes due to their unique optical and physicochemical properties, offering distinct superiority over organic fluorophores in chemical or biological applications [15].
Quantum dots (QDs) are nanoparticles with a size typically between 1 nm to 10 nm, confined inside three dimensions with quantized energy states because of their smaller size than the Bohr radius [16,17]. Generally, QDs are divided into two classes: one is composed of II-I, III-V, and IV-VI elements (e.g., ZnS, ZnSe, ZnO, CdS, InAs), and the second is based on group IV elements, such as carbon quantum dots (CQDs), graphene quantum dots (GQDs), silicon quantum dots (Si QDs), and germanium quantum dots (Ge QDs) [18,19]. In recent years, perovskite QDs (PQDs), as a new type of QD, have been increasingly recognized, studied, and applied [20,21,22,23,24,25,26]. The fluorescence properties of QDs can be adjusted by changing the size, morphology, composition, doping, and surface engineering. For these reasons, QDs have been applied successively to fluorescence sensors based on their excellent properties, such as good optical stability and easy surface modification and optical adjustment [27,28,29,30,31,32]. Other optical properties of QDs that are not reflected in conventional organic fluorophores include wide absorption spectra, narrow and symmetrical emission bands, large stoke shifts, and high optical stability, which also present powerful attractiveness in fluorescence sensing applications [27,28]. In addition, QDs present 20 times brighter than conventional organic fluorophores due to their higher quantum yield and molar absorption coefficient. QDs are thousands of times more stable against photobleaching than traditional organic dyes. Moreover, QDs can be surface functionalized easily by conjugation electrostatically and covalently, either directly or via a bridge.
In recent years, the surface modification and functionalization of QDs have attracted extensive attention by improving the stability, biocompatibility, sensitivity, and selectivity of QDs-based sensors in complex practical detection conditions. The fluorescence efficiency of QDs is sensitive to the presence of adsorbent on the surfaces of QDs due to the large number of energy states present in QDs. Compared with pure QDs, QDs functionalized by organic or inorganic ligands would provide significantly better fluorescence performance in terms of fluorescence intensity properties, chemical stability, and stability against photobleaching. QD-based sensors for heavy metal ions and pesticide residues in aqueous solution were prepared by modifying the surfaces of QDs with suitable ligand molecules, such as cysteine, glutathione, mercaptoacetic acid, L-aspartic acid, and thioglycolic acids [19,28,31]. In addition, inorganic materials contribute to the surface passivation of QDs resulting from the formation of the core-shell structures of QDs [19].
Recently, optical sensors have revealed their significant potential and application in the determination of heavy metal ions [33,34,35] and pesticide residues [36,37,38,39,40,41]. Based on the advantages of QDs and their potential application in developing fluorescence sensors and performing rapid in-situ detection, there is a need to advance the study status and discuss further research trends regarding QDs as fluorescence sensors. Thus, in this review, a detailed introduction is presented to the recent advances in QD-based fluorescence sensors, such as SCQDs, CQDs, GQDs, and their nanocomposites with various inorganic and organic materials, for the determination of multitudinous heavy metal ions and pesticide residues. Furthermore, the current problems and challenges in the application of QD-based fluorescence sensors are discussed.

2. QD-Based Fluorescence Sensors

Generally, the fluorescence of QDs is attributed to the recombination of excitons (electrons and holes). Changing the surface state or ligand composition of QDs will affect the recombination efficiency of excitons and thus influence the fluorescent efficiency. Currently, various QD-based fluorescent sensors have been studied based on the direct or indirect interaction between QDs and target analytes. The quantitative analysis of target analytes is achieved through the linear correlation between the concentration of the analyte and the fluorescence intensity changes of the QDs. Therefore, functionalized QDs could be obtained by changing the surface ligands of QDs, and then QD-based fluorescence sensors could be developed by utilizing the fluorescence changes resulting from direct physical adsorption or chelation between target analytes and small molecules or functional groups on the surfaces of QDs.
In 1987, Spanhel et al. [42] first proposed that Cd2+ could improve the photoluminescence efficiency by up to 50% in alkaline water solution. This effect could be ascribed to the formation of the Cd(OH)2 shell on the CdS core. Furthermore, the Cd(OH)2 shell would effectively weaken the non-radiative recombination of carriers. In 2002, Chen and Rosenzweig [43] synthesized CdS QDs coated with mercaptoglycerol and used as fluorescent probe for the detection of Cu2+ in aqueous solution, and the potential of fluorescence sensors based on QDs was determined. QDs have been widely used in new fluorescent probes for metal ions over the last decade. Usually, different compounds and special functional groups are introduced on the surfaces of QDs to form specific selective fluorescent probes for detecting different metal ions. CdS QDs, CdSe QDs, and CdTe QDs are often used for detecting metal ions, such as Hg2+ and Cu2+. Zhao’s group [44] synthesized CdTe QDs and CdTe QDs modified by bovine serum albumin (BSA). The fluorescence experiments showed that the fluorescence of QDs was significantly enhanced due to linking of BSA on the surfaces of QDs. When Hg2+ reacts with BSA-coated CdTe QDs, the fluorescence intensity of QDs can be effectively quenched. BSA-coated CdTe QD probes for Hg2+ showed a linear detection range from 0.001 to 1 μM. Uppa et al. [45] also prepared Hg2+ fluorescent probes relying on cysteamine capped CdS QDs modulated Ag+, and this probe shown excellent detection sensitivity. Han and coworkers [46] synthesized BSA functionalized CdS QDs by a stepwise procedure, and they were used as fluorescence sensors for the determination of Cu2+. This sensor exhibited good interference and linear responses in the range of 0–80 µM. Moreover, carbon-based QDs as fluorescence sensors are increasingly becoming the focus of attention due to their simple synthesis technology, high photostability, low cytotoxicity, and better biocompatibility [47,48,49].

3. Mechanisms of Fluorescence Sensors

As previously reported, various QD-based fluorescence sensors have achieved the identification and quantification of target analytes. Overall, the change in the fluorescence intensity of QDs is attributed to the alteration of energy transfer pathways. For different types of QDs, several mechanisms have been proposed.
Fluorescence quenching or enhancement is the change in the fluorescence quantum yield in QDs, which is caused by the interaction between QDs and intended analytes according to fluorescence resonance energy transfer (FRET), photoinduced electron transfer (PET), internal filtration effects (IFEs), aggregate effects (AEs), static quenching effects (SQEs), and dynamic quenching effects (DQEs) [50,51]. FRET, PET, and IFEs are the three main mechanisms that might be included in dynamic quenching. The quenching process requires sufficient contact between QDs and analytes. Diffusive encounters and collision contact can be attributed to the dynamic quenching process, while new complex forms without changes in fluorescence lifetime will be assigned to the static quenching process [52]. However, dynamic quenching (i.e., FRET, PET) can cause changes in the fluorescence lifetime of fluorescence QDs after adding analytes, while these changes cannot be observed in IFEs and SQEs [53]. In addition, the fluorescence enhancement of QDs caused by the interaction of QDs and analytes can be attributed to surface plasmon enhanced fluorescence (SPEF), which is also referred to as metal surface enhanced fluorescence (MEF) and chelate enhanced fluorescence (CHEF). MEF is defined as the fluorescence being enhanced when the distances between fluorophores and metal are within the range of 5 to 90 nm [54], while CHEF can be due to coordination of the surface functional groups with analytes [15].

3.1. Fluorescence Resonance Energy Transfer (FRET)

Fluorescence resonance energy transfer (FRET) is based upon non-radiative energy transfer between proximal molecules. Generally, the FRET system consists of two fluorophores, which are defined as donor and acceptor molecules. The donor molecule (fluorescence QDs) absorbs a photon of a certain frequency and is excited to a higher electron energy state, and before the electron returns to the ground state, the energy transfer is realized to the neighboring receptor molecule through dipole interaction so that the donor fluorescence intensity is reduced, and the receptor can emit stronger fluorescence than its characteristic fluorescence (fluorescence sensitization) or no fluorescence (fluorescence quenching). This process is also accompanied by a corresponding reduction or extension of the fluorescence lifetime.
Recently, FRET-based sensors have been widely used in the field of detecting and quantizing analytes of interest, such as heavy metal ions and pesticides. Various fluorescence sensors based on the FRET theory have been designed and constructed, and the mechanism can be classified into the following categories.
(i)
Original FRET process destruction
In a FRET-based sensor, the donor and acceptor molecules are typically fluorescent nanomaterials that are engineered to be in close proximity to each other. In this system, QDs act as donors, the modified molecules or doped ions serve as accepters, and the analyte acts as a quencher. Upon incorporation of analyte, the FRET between the donor and acceptor could be destroyed due to the quenching of the QDs. Wang et al. [55] constructed a FRET system between cysteamine capped CdS QDs (Mea-CdS QDs) and fluorescein and applied it as a sensor for the content determination of Cu2+. The presence of Cu2+ could destroy the FRET process between Mea-CdS QDs and fluorescein, resulting in quenching of the Mea-CdS QDs’ fluorescence. The linear response range was 4–14 μM with an LOD of 0.17 μM.
(ii)
New FRET system formation
Due to the specific interactions between metal ions and functional groups or ligands, donors and acceptors can be connected together through additional analytes, forming a FRET system. Huang et al. [56] reported a strategy for Hg2+ detection using Mn-doped CdS/ZnS Core/Shell QDs (donor) and gold nanoparticles (Au NPs, acceptor) via DNA hybridization, which was generated when Hg2+ ions were present in aqueous solution. This sensor exhibited a linear range from 1 nM to 10 nM with a LOD of 0.49 nM. Wang and colleagues [57] reported a FRET system for Hg2+ detection using two-color CdTe quantum dots assisted by cetyltrimethylammonium bromide. Li et al. [58] used CdS QDs and Au NPs to structure a FRET method to detect thrombin.
(iii)
Analyte-based FRET system activation
This theory supposes the existence of potential FRET in the original system. After incorporation with analytes, FRET emerges between the QDs and the acceptor due to the acceptor being activated. Zhang’s group [59] fabricated, based on rhodamine derivative and Zn coped CdTe QDs (CdTe:Zn QDs), a fluorescent sensor for the detection of Hg2+. CdTe:Zn QDs/silica cores acted as donors, which were formed via reverse microemulsion method, and rhodamine derivative acted as the probe for Hg2+. The integration between Hg2+ and rhodamine derivative resulted in the formation of fluorescence of FRET that could be attributed to the activated reaction of the rhodamine moieties. In the case of the presence of Hg2+, CdTe:Zn QDs would not be quenched because of the existence of the aminos but rather transfer their excited energy to the complex of the probe/Hg2+ (acceptor). With the increasing of Hg2+ concentration, the emission peak at 521 nm gradually decreased, and the rhodamine B derivative emission peaked at 577 nm. Therefore, FRET-based sensing for Hg2+ was achieved, and the detection limit reached 0.5 µM. Wu and colleagues [60], grounded in the same strategy, constructed CdSe/ZnS QD-based ratiometric fluorescent sensors for Hg2+.

3.2. Photo-Induced Electron Transfer (PET)

The photo-induced electron transfer mechanism refers to the charge transfer between donor and acceptor molecules. There is a great tendency for heavy metal ions to form complexes with N, O, S-containing ligands, or groups. This property promotes the formation of electron-rich fluorophores based on donor–receptor interactions. Intramolecular charge transfer and charge separation occur in PET, and PET can be divided into reducing PET and oxidizing PET. In reducing PET, electrons are accepted from the electron donor, while in oxidizing PET, electrons are donated to the electron acceptor. Quenching efficiency and electron transfer present a positive correlation relationship. PET has the ability of fluorescence “turn-on” and “turn-off” for the determination of pesticide residues and heavy metal ions.

3.3. Internal Filtration Effects (IFEs)

IFEs are caused by a large concentration of quencher in the detection system; part of the excitation or emission light is absorbed by the quencher, which leads to the weakened excited light being accepted by the QDs, resulting in fluorescence quenching. Compared with other quenches, IFEs do not constitute a typical fluorescence quenching process, only involving the absorption of excitation or emission of light by the quencher. There is no energy or charge transfer between the quencher and QDs, so the fluorescence lifetime of the QDs remains unchanged after the addition of quencher. At the same time, no new absorption peaks are generated in the absorption spectrum because no new substances are formed.

3.4. Aggregation Effects (AEs)

The strong interaction between adsorbed molecules and the surface molecules of quantum dots usually leads to aggregation effects. The addition of quencher reduces the stability of fluorescent materials and leads to aggregation, resulting in the formation of large particles. Photo-induced electrons are captured by surface defects, forming non-radiative recombination, and agglomeration leads to the increase in the particle size of fluorescent materials and the weakening of quantum effects, resulting in fluorescence quenching, which can be called aggregation-caused quenching (ACQ).

4. QDs as Fluorescence Sensors

Due to their excellent optical properties and unique mechanism, QDs-based fluorescence sensors have achieved rapid development with high sensitivity, high selectivity, high photostability and low LOD. In addition, the sensitivity and selectivity of QD-based fluorescence sensors have been greatly improved on the basis of functionalization or integration with small molecular, nanomaterials and biomaterials. However, these sensors have limitations, such as toxicity, specificity, multiplex detection, and portability. In this section, we summarize the progress in the research on and application of QDs in the detection of metal ions and pesticide residues, and we discuss various sensing strategies.

4.1. QDs for Detection of Heavy Metal Ions

The easy surface functionalization of QDs allows for making them hydrophilic and increasing their biocompatibility. Currently, QDs as fluorescence sensors have been widely used in the detection of different heavy metal ions, including copper, mercury, iron, cobalt, lead, and chromium ions.

4.1.1. Copper(II) Ions

Copper, one of the essential elements for animals and humans, is found in many proteins and enzymes, and it also has a wide range of applications in industry and medicine. However, excessive Cu2+ can cause kidney damage, liver necrosis, brain tissue lesions, and DNA damage and even induce cell death, so it is important to analyze copper ion concentrations sensitively. In this section, the research work of our group based on SCQD sensors is introduced, and we discuss the current studies on non-metal-based QDs.
Regarding QD-based fluorescence sensor, our group previously used N-acetyl-L-cysteines acid modified CdS QDs (NALC-CdS QDs) for selective and sensitive detection of Cu2+ in the aqueous system with a limit of detection (LOD) of 0.48 µM and an excellent linear range from 0 to 25 µM (Figure 1a–c) [61]. The fluorescence of NALC-CdS QDs is effectively quenched by the introduction of Cu2+ due to PET and aggregate effects. Moreover, testing of real water samples confirms its availability and potential in practical application. CdS QDs (CA-CdS QDs) capped with citric acid, a specific Cu2+ chelator, as fluorescence sensors for Cu2+ were reported, and a LOD of 9.2 nM was achieved [62]. The fluorescence quenching condition is most often attributed to the adsorbed.
Cu2+ appears on the surface and is reduced to Cu+ by the S2− vacancies of QDs, resulting in the formation of non-radiative surface channels, which could be described using the following approach: CdS + Cu2+ → CdS+ + Cu+. In this study, a linear relationship was obtained within the range of 0.01–50 µM with a LOD of 9.2 nM. The selectivity of the sensor has been evaluated in the presence of various common metal ions. In addition, the sensor has been used for the detection of Cu2+ in tap water, suggesting that the proposed sensor has potential application for environmental monitoring.
ZnO QDs functionalized with (3-aminopropyl) triethoxysilane (APTEs) (NH2-ZnO QDs) were synthesized via the simple sol-gel method [63]. The photoluminescence of NH2-ZnO QDs was selectively quenched by addition of Cu2+ (Figure 1d), and two linear relationships were obtained in the ranges of 2–20 nM and 1–100 µM (Figure 1e,f). Moreover, this sensor can provide a low LOD of 1.72 nM. The detection mechanism is shown in Figure 1g. In addition, our group also studied fluorescence sensors based on ZnO nanoparticles and realized the quantitative detection of Cu2+ with a detection range of 10–1000 µM [64]. Interestingly, this sensor can provide a paper sensor, which was constructed for visual observation of Cu2+ by the naked eye. It appears that the sensing system might provide a fast and convenient platform for other analytes by applying visualization modules.
Non-metal-based QDs, such as CQDs, GQDs, and their nanocomposites, have also been researched as fluorescence sensors for the detection of various heavy metal ions in recent years. Liu et al. [65] developed quenching-based detection of Cu2+ using a carbon dot-based sensor. The study provided a new method for improving the sensitivity of fluorescence methods based on reducing surface defects and enhancing fluorescence intensity, and they proposed that the fluorescence quenching of CDs-BSA could be attributed to the multi-site coordination of Cu2+ with the -COOH and -NH2 on the surface of CDs-BSA. Dong et al. [66] investigated polyamine-functionalized CQDs (BPEI-CQDs) as sensitive Cu2+ fluorescent probes. They found that Cu2+ can be combined onto the surfaces of CQDs through amino groups and form cupric amine, leading to fluorescence quenching of the CQDs resulting from the IFEs, and the prepared sensor presented a linear response range of 10–1100 nM with an LOD of 6 nM (Figure 2). In this study, the proposed sensor’s accessibility was evaluated in a Min River water sample and provided a result agreeing with that obtained by the ICPMS method. As a result, the study-proposed sensors have promising applications in the detection of contaminants in environmental water samples. Wang et al. [67] designed CQDs@Cu-IIP as a fluorescence sensor for Cu2+ by combining fluorescence analysis technology with ion imprinting technology. In this sensing system, CQDs were grafted onto the surface of an SBA-15-NH2 via an amide reaction. Copper ion imprinted polymer (Cu(II)-IIP) was prepared by surface imprinting technique with SBA-15-NH2 as the substrate, copper ions as a template, 3-aminopropyl-3-ethoxysilane as a functional monomer, tetraethoxysilane as a crosslinker, ammonia water as an initiator, and water as a solvent throughout the process. In the study, the selectivity of a CQD-based sensor for Cu2+ was greatly improved due to ion imprinting polymers (IIPs) with specific recognition sites for target ions. Furthermore, the sensor has been applied to tap water and river water with recovery rates of 99.29–105.42%.
Recently, carbon-based materials have revealed many advantages in the design of fluorescence sensors. Tan’s group [68] designed a fluorescence “on–off–on” assay of Cu2+ and EDTA based on the structure of af-GQDs-PNA. In this study, the selectivity for Cu2+ is due to the specific reaction of Cu2+ with the introduced 1-(2-pyridylazo)-2-naphthol (PAN), rather than from direct interaction with GQDs. Specifically, the fluorescence of af-GQDs can be effectively quenched by PAN based on the FRET between af-GQDs and PAN, and the complex formation of Cu2+-PAN could improve the FRET efficiency, while EDTA would weaken the FRET effect and cause the fluorescence recovery of GQDs owing to its strong chelating ability. This proposed sensor exhibited a linear response for Cu2+ within a concentration range from 0.01–10 μM with an LOD of 0.87 nM, in addition to the selectivity of this sensor for detecting Cu2+ in the presence of various metal ions, such as Mn2+, Ni2+, Fe2+, Cd2+, Pb2+, Bi3+, Zn2+, Cr3+, and Fe3+. Finally, the proposed fluorescence “on–off–on” sensor was successfully used for the detection of Cu2+ in water and human serum. As a result, this sensor will provide a feasible strategy for design of sensors with high selectivity for and interference with different analytes. The previously reported QD-based sensors for the detection of Cu2+ are summarized in Table 1.

4.1.2. Mercury Ions

The detection of mercury ions is necessary because they are currently the most toxic heavy metal ions and can accumulate in the body and cause irreversible health damage. Moreover, they can exert harmful effects on all living organisms even at low concentrations. Therefore, it is of inestimable significance to establish a rapid and accurate qualitative and quantitative method to detect mercury ions for human life activities. Currently, many sensors based on functionalized QDs for detection of Hg2+ have been reported. Generally, for the types of metal chalcogenide QDs, including CdS, ZnS, CdSe, ZnSe, and CdTe, a precipitate of HgX (X = S, Se, Te) occurs when Hg2+ is added due to the lower solubility product (Ksp) values with corresponding thioides, which can replace the cations of QDs by cation exchange. As a consequence, the formation of precipitation will lead to the generation of surface defects, in turn resulting in non-radiative recombination and leading to the fluorescence of QDs being quenched. While based on functionalized GQDs or CQDs, the detection mechanisms usually depend on Hg2+ quenching fluorescence due to the electron transfer occurring between the QDs and Hg2+ based on the interactions with surface groups, either dynamic or static. Generally, the detection of Hg2+ is realized according to the linear or nonlinear relationship between the observed attenuation of fluorescence intensity of quantum dots and the concentration of Hg2+.
Choudhary and Nageswaran [85] proposed mercapto acid (3-MIBA) coated CdTe QDs as sensor for the selective detection of Hg2+. There are two reasons for the fluorescence quenching of the QDs by Hg2+, including the strong affinity of Hg2+ to S2- leading to the deactivation of the capping agents, resulting in the aggregation of the quantum dots, and the excitonic electron transfer from the LUMO of QDs to LUMO of Hg2+, preventing the emission. This sensor is based on a linear relationship between the fluorescence intensity and the concentration of Hg2+ in the range of 1.5–100 nM with an LOD of 1.5 ± 0.5 nM. Based on the same mechanisms, Singhal’s group [86] developed CdS QDs functionalized by glutathione (GSH@CdS QDs) as fluorescence probes for sensing of Hg2+. This probe could attain trace determination of Hg2+ with a low LOD of 0.54 nM in two successive linear ranges of 0–1000 nM and 1–20 µM. The authors found that the interaction between Hg2+ and the surfaces of QDs varies with different concentrations of Hg2+. At the lower concentration of Hg2+, the fluorescence quenching may be due to the interaction between Hg2+ ions and the thiol group of the GSH capping layer, causing the dissociation of the GSH capping layer from the surfaces of QDs and resulting in the aggregation of CdS QDs, while at higher concentrations of Hg2+, the quenching would be interpreted as the binding between Hg2+ and S2- and formatting the HgS on the surfaces of CdS QDs, which can induce effective non-radiative electron transfer. You and coworkers [87] synthesized polyethyleneimine passivated Ag2S QDs (PEI-Ag2S QDs) that were utilized for fluorescence detection of Hg2+ with a low LOD of 0.5 nM based on the metal-induced aggregation strategy. The interference of the sensor has been examined in the presence of common metal ions and anions, suggesting that other interfering ions were unaffected by the fluorescence of the proposed sensor. Finally, the sensor has been used for the detection of Hg2+ in tap, river, and lake water. Recently, a similar investigation was proposed by Granados-Oliveros et al. [88]. They obtained two structures of fluorescent probes for Hg2+ based on functionalized CdSe/ZnS QDs by oleic acid (OA) and L-glutathione. The designed fluorescent probes demonstrated excellent selectivity to Hg2+ due to the lower Ksp of HgSe and HgS compared to the other metal ions, and they established good linearity with the LODs of 74.8 nM and 54.8 nM for CdSe/ZnS/OA and CdSe/ZnS/GSH, respectively.
In recent years, CDs, CQDs, and GQDs has been widely reported for the detection of Hg2+ owing to their simple preparation, high selectivity and sensitivity, and affordability. Our groups investigated fluorescent probes, especially GQD-based materials, as sensors for Hg2+. N-doped GQDs (N-GQDs) [89], N, S-GQDs [90] (Figure 3), and rhodamine B assisted GQDs (RhB-GQDs) [91] were successfully synthesized by a simply hydrothermal approach, and they emerged as highly sensitive and selective in response to Hg2+. The fluorescence quenching of the N-GQDs occurred upon the incorporation of Hg2+ owing to efficient non-radiative electron transfer from N-GQDs to the excited state of Hg2+ and the coordination effect of N-GQDs with Hg2+. From this observation, a two-stage linear relationship between the fluorescence intensity of N-GQDs and the concentration of Hg2+ was discussed in the range of 1–1000 nM. The LODs were obtained as 0.45 nM and 67.3 nM The selectivity of the sensor has been assessed in the presence of various cations, suggesting that the existence of interfering cations does not affect the detection of Hg2+.These results clarified that the proposed sensor has high selectivity for Hg2+ detection in complex mixtures.
Subsequently, the N, S-GQDs were also studied as a fluorescent probe for Hg2+. In this study, a good linear response was achieved according to the fluorescence “turn off” system in the range of 1–30 nM with a low LOD of 0.27 nM. Furthermore, the fluorescence quenching mechanisms were explored by density functional theory (DFT) simulation (Figure 3a,b). The present of an S atom breaks the electrical neutrality of the C atom, which benefits the generation of charge favorable sites for the interaction of Hg2+ and N, S-GQDs. Thereafter, as indicated in Figure 3c, the fluorescence quenching for Hg2+ is mainly attributed to the dynamic quenching effects (DQEs), which can be ascribed to the adsorption of Hg2+ altering the charge distribution of N, S-GQDs resulting from the interaction of N, S-GQDs and Hg2+. However, there is interference from Fe3+. Next, RhB-GQDs were investigated for improving the selectivity of Hg2+. In this study, rhodamine B could not only act as a nitrogen source during the formation of RhB-GQDs but also form a protective layer on RhB-GQDs through surface adsorption. The excellent selectivity for Hg2+ is mainly attributed to the establishment of a rhodamine B protective shell on the surfaces of GQDs, further reducing the interference of other metal ions with weak affinity for the N and O functional groups. The strong affinity between Hg2+ and the -COOH groups of RhB-GQDs promote the transfer of non-radiating electrons and energy, resulting in dynamic fluorescence quenching. The designed sensor displayed a good linear relationship within the ranges of 0–10 nM and 50–1000 nM with the outstanding LODs of 0.16 nM and 0.48 nM, respectively. Moreover, this sensor was applied to the analysis of drinking water.
Table 2 summarizes various QD-based fluorescence sensors for Hg2+ detection in recent years. Of these studies, the majority focused on the detection of Hg2+ in water. Only a few have been used in the analysis of cells, food, or beverages.

4.1.3. Iron(III) Ions

Iron, an indispensable trace element, plays a vital role in basic biological processes, including the synthesis of hemoglobin, myoglobin, and DNA; oxygen transport; and cellular metabolism in living bodies, which is especially important for women [99,100,101]. The insufficiency or excess consumption of Fe3+ can cause a variety of diseases, such as anemia, hemochromatosis, diabetes, liver damage, arthritis, diabetes, heart failure, Parkinson, and ever cancer. Therefore, a variety of QD-based sensors have been reported for the sensitive detection of Fe3+.
Moreover, Zhou and coworkers [102] developed a ratiometric fluorescence detection method for Fe3+ based on CdTe QDs capped by thioglycolic acid (TGA) and N-acetyl-L-cysteine (NAC). The synthesized red emissive NAC capped QDs showed excellent selective quenching response for Fe3+, while the green emissive TGA capped QDs exhibited stable fluorescence emission toward all the tested metal ions. The quenching of NAC-CdS QDs might be attributable to ligand desorption, leading to the particle growth and agglomeration of QDs. The two types of CdTe QDs were incorporated into hydrogel optical fibers and achieved the ratiometric detection of Fe3+ based on the ratio of the fluorescence intensities at the two emission wavelengths of 520 nm and 628 nm. Additionally, the fluorescence probe achieved quantitative detection for Fe3+ within the limit of 0–3.5 μM with an LOD of 14 nM. According to similar principles, Kuang et al. [103] synthesized SiO2-embedded CdTe QDs (CdTe@SiO2 QDs) functionalized with rhodamine derivative to design the ratiometric fluorescence sensor for Fe3+ detection and achieved visual detection of Fe3+. Moreover, Srivastava’s group [104] fabricated functionalized SnS2 QDs and applied them in the detection for Fe3+. They found fluorescence quenching caused by both static and dynamic quenching effects (Figure 4). The static quenching process results from the formation of a non-fluorescent ground state complex (such as iron hydroxide), while the dynamic process is due to coordination between electron-rich NH2- and electron-deficient Fe3+.
However, the majority of GQDs that detect Fe3+ depend on high affinity for nitrogen and O-containing surface groups for quenching the fluorescence of GQDs [90,105,106]. Currently, sensors based on GQD utilize precursors that are rich in O-containing functional groups to form nitrogen and oxygen groups on the surfaces of QDs. The surface groups can increase the number of functional groups on oxygen-containing surfaces and improve the selectivity for Fe3+ by transferring electrons from excited state to metal ions and resulting in the formation of a ground state complex without fluorescence. Lou and coworkers [107] proposed poly(ethylene glycol) (PEG) passivated GQDs as fluorescence sensors for the detection of Fe3+ based on fluorescence quenching. In this study, the selection of PEG-GQDs for Fe3+ have been well improved owing to ethylenediaminetetraacetic acid (EDTA) being used as the masking agent. Compared with GQDs, the -OH on the surface of PEG-GQDs is propitious to forming metal hydroxides, which could contribute to better sensitivity for the detection of Fe3+. Similarly, Zhou et al. [108] developed N-doped GQDs (N-GQDs) via hydrothermal methods and used them as a probe for analysis of Fe3+. Herein, N-GQDs had higher fluorescence intensity and a smaller size than GQDs. The smaller size could provide more active sites, making them more prone to electronic transition. The presence of Fe3+, N-GQDs will bind to Fe3+, causing the formation of a non-fluorescent complex, which could suppress the radiative recombination of excited electron–hole pairs in N-GQDs and thus the fluorescence quenching of N-GQDs. The linear range was obtained from 0 to 100 µM with an LOD of 0.74 µM. Notably, the successful application of N-GQDs in in vitro bioimaging indicates the further practicability of N-GQDs.
Additionally, based on the same theory, Sun’s group [109] synthesized a CQD-based hydrogel (HG) nanocomposite material (CQDsHG) as a probe for the determination of Fe3+. The CQDsHG showed a sensitively response toward Fe3+ within the linear detection range of 0 to 150 µM and achieved an LOD of 0.24 µM. In brief, the CQDs were prepared using a microwave-assisted hydrothermal method and then loaded into HG through the sol-gel method to obtain the functional CQDsHG. The proposed CQDsHG revealed high adsorption (31.94 mg/g) and a selective quenching response for Fe3+. The linear response range was established as 0–150 µM with an LOD of 0.24 µM. Furthermore, the application of CQDsHG was demonstrated in tap water and lake water. As a result, the CQDsHG provides a bifunctional platform for removing and qualitatively detecting Cu2+.
Huang and Tong [110] structured a B-CQDs/CdTe-Eu3+ composite as a dual-emission ratiometric fluorescence sensor for tetracycline (TC) and Fe3+. Herein, green-emitting CdTe QDs were modified by Eu3+, and blue-emitting B-CQDs were synthesized using a typical hydrothermal method. The fluorescence of CdTe QDs was quenched by Eu3+ due to aggregation, while the addition of TC could restore the fluorescence of CdTe QDs. On this basis, the study introduced the blue-emitting B-CQDs with specific responses toward Fe3+ and fabricated B-CQDs/CdTe-Eu3+ ratiometric fluorescence sensors for TC and Fe3+. In this study, CdTe-Eu3+ only had a response to TC and not to Fe3+, while B-CQDs only had a response to Fe3+ and not to TC. As a result, this proposed sensor provides a visual and semi-quantitative method for the detection of Fe3+ and TC based on the different emissions. Furthermore, the influence of interfering ions, including various common cations, anions, and organic compounds, has been examined. In conclusion, the proposed sensor provides a highly sensitive platform for Fe3+ sensing. Finally, the actual application of the platform for selective and capable analysis of Fe3+ and TC has been demonstrated in river and lake waters.
For further enhancement of the sensitivity and selectivity of QD-based fluorescence sensors to Fe3+, a variety of sensing systems have been proposed in recent years. Table 3 summarizes the QD-based nanomaterials as fluorescence sensors for Fe3+ detection.

4.1.4. Lead(II) Ions

Lead, one of the most dangerous hazardous heavy metal elements, has high toxicity and non-degradability, and it can be absorbed by the human body via direct contact or the food chain [117,118]. In drinking water, the nontoxic allowable limit for Pb is 10 µg/L (48 nM), which is stipulated by the WHO [119]. Lead has been studied widely due to its substantial health hazards for humans and animals. Until now, various QD-based fluorescent sensors have been widely explored based on the specific coordination between Pb2+ and the function groups or ligands on the surfaces of QDs. Generally, the interactions between QDs and Pb2+ are mostly related to direct fluorescence quenching or an energy transfer processes (i.e., FRET, PET, etc.).
Wu et al. [120] prepared thiol-capped CdTe QDs for the detection of Pb2+. They concluded that the presence of Pb2+ could change the surface state of CdTe QDs, resulting in fluorescence quenching. The sensor showed a linear range of 2–100 µM with a LOD of 270 nM. Finally, feasibility analysis is conducted, the proposed sensor was successfully applied to the analysis of Pb2+ in food, and the results revealed that it holds enormous potential for applications in food safe monitoring. Li’s group [121] developed a fluorescence sensor for the detection of Pb2+ according to the fluorescence quenching of TGA-CdTe QDs, and this sensor was successfully used for the determination of Pb2+ in food. In this study, smaller-sized QDs exhibited a detection limit of 4.7 nM within the linear range of 1.96–15.9 nM, while larger QDs showed a higher LOD of 0.22 µM in the detection range of 4.98 nM to 0.285 µM. Recently, ZnSeS/Cu:ZnS/ZnS QDs capped by 3-mercaptopropionic acid (3-MPA) with a core/shell/shell structure were developed for the detection of Pb2+ [122]. Fluorescence quenching could be caused by the collision interaction between QDs and Pb2+, resulting in the partial displacement of the 3-MPA ligand on the QD surface, which could change the QD surface properties and generate surface defects. In addition, the energy transduction between QDs and Pb2+-3-MPA complexes may also lead to fluorescence quenching. In this study, the sensor displayed a linear response between 0.04 and 6 µM with a low LOD of 21 nM. The selectivity of the sensor revealed strong interference ability in the presence of interfering elements, such as coexisting cations, various amino acids, and relevant molecules. Moreover, this sensor was used in the analysis of real samples and exhibited LODs of 21, 76, and 28 nM in ultrapure water, mineral water, and lake water, respectively.
Furthermore, fluorescence enhancement is a rare case in the detection of Pb2+. Cheng and coworkers [123] showed a fluorescence turn-on mechanism based on 3-MPA/GSH-Ag2S QDs for the detection of Pb2+ due to the signal intensity of Ag2S QDs via aggregation-induced enhanced emission (AIEE). In this case system, the probe presented a linear response range from 0.05 to 20 µM and an LOD calculated to be 15.5 nM. Recently, several carbon-based materials (e. g., CDs, CQDs, GQDs, and their nanocomposites) have been explored for the development of Pb2+ sensors based on changes in fluorescence. Highly fluorescent CDs based on the functionalization of 3-MPA were used as a fluorescent sensor for Pb2+ relying on a “turn-on” model [124]. The sensor exhibited high sensitivity and selectivity based on fluorescence-enhancing attributes to the chelation action between Pb2+ and -COOH groups present in 3-MPA, resulting in AIEE. The relationship between the fluorescence intensity of QDs and the concentration of Pb2+ showed a linear relation, and an LOD of 0.051 nM was obtained by calculation. In this study, a turbidimetric method was developed for the detection of Pb2+ with an LOD of 13.2 nM with the naked eye owing to the formation of milky white precipitates resulting from Pb2+-induced aggregation of QDs.
Babu et al. [125] synthesized CQDs from Delonix regia that were also applied in the construction of Pb2+ sensors. The developed sensors for Pb2+ detection showed an LOD of 3.3 nM within the linear response range of 10–180 µM (Figure 5a–c). The mechanisms of this sensor are shown in Figure 5d. With this sensor, the fluorescence quenching of CQDs is largely due to FRET between the CQDs and Pb2+. Herein, Pb2+ acts as a non-fluorescent acceptor, and CQDs act as a donor fluorophore. In addition, Pb2+ is prone to be adsorbed on the surfaces of CQDs to form a non-fluorescent complex due to electrostatic interactions between CQDs and Pb2+. The formation of non-fluorescent complexes also leads to the fluorescence quenching of CQDs. Additionally, the sensor was studied for its applications in tap, river, and lake waters with recovery values of 99–101%. Likewise, the proposed sensor was also used in industrial effluents, yielding recovery rates of 98–100%. As a result, the proposed sensor exhibited great potential in environmental monitoring.
Along with CQDs, GQDs have also been extensively used for the detection for Pb2+. Qi et al. [126] reported functionalized GQDs by 3,9-dithia-6-monoazaundecane (DMA) as a fluorescence probe for Pb2+, relying on the GQD–DMA–tryptophan compound system. Interference tests have been evaluated using various common metal ions. It has been revealed that Pb2+ has a strong effect on the fluorescence quenching of GQDs. This sensor provides a linear response within the range of 10–1000 pM with a low LOD of 9 pM. Significantly, the sensor has been applied in the cerebrospinal fluid (striatum) of rats. The results revealed that the proposed sensor has high sensitivity and selectivity for the detection of Pb2+ in complex mixtures. As a result, the current study demonstrated the sensors have potential application as a highly selective and sensitive Pb2+ sensor in the biology and medicine environments.
Later, Feng’s team [127] developed a fluorescence “turn-on” type for Pb2+ detection based on a PET process between rGQDs and GO. rGQDs were prepared via the reduction of GQDs, GO was prepared through oxidation by concentration, and aptamer-rGQDs were synthesized through condensation reactions with EDC. In this study, a sensing mechanism was involved, based on the photoinduced electron transfer between rGQDs and GO. The aptamer–rGQDs could bind to the surface of GO due to electrostatic attraction and π–π stacking interactions. As an effect, GO induces significant fluorescence quenching of the rGQDs’ attributes due to photoinduced electron transfer between rGQDs and GO. The formation of assembly aptamer–rGQDs/GO acts as a quencher and effectively “turns off” the fluorescence. Furthermore, Pb2+ can strongly coordinate with the aptamer on the surfaces of rGQDs, resulting in the forming of the G-quadruplex/Pb2+ complex, which can break up electrostatic attraction and π–π stacking interactions between aptamer–rGQDs and GO and effectively “turn on” the fluorescence. As a result, the sensor established a broad linear response within the range of 9.9–435 nM and found an LOD of 0.6 nM. Furthermore, the selectivity of the sensor has been investigated for Pb2+ detection. It has been found that there is no obvious change in fluorescence intensity. Therefore, this proposed sensor has high sensitivity and good reproducibility for sensing Pb2+.
Recently, dithienopyrrole derivative functionalized GQDs (DTPAN-fn-GQDs) were used as fluorescence probes for improving selectivity and sensitivity of Pb2+ [128]. The fluorescence quenching of GQDs may result from the existence of Pb2+, which will destroy the electron transfer that occurs from the HOMO of DTPAN to the LUMO of GQD. Moreover, this sensor was also fabricated as paper strips for rapid visual determination of Pb2+. The linearity range for Pb2+ sensing was obtained from 0.5 to 40 nM (R2 = 0.98) with a low LOD of 0.25 nM. The selectivity and interference were evaluated by the interference of common metal ions and organic molecules. It was found that there was no significant change in fluorescence intensity upon the addition of the interference analyte. Finally, the applicability of DTPAN-fn-GQD has been verified by applying it in real water samples and developing paper strips.
Table 4 summarizes the various fluorescent sensors for the detection of Pb2+, including the sensing mechanism, type of the sensor, sensitivity, and LOD.

4.1.5. Other Metal Ions

The surface functionalization of QDs ensures their specificity of detection, providing a feasibility strategy for the design of QD-based sensors. Until now, various groups, such as -COOH, -NH2, and -OH, have been used to improve the selectivity and sensitivity of QD-based sensors for other metal ions, including Co2+, Zn2+, Cd2+, Cr6+, Cr3+, Ag+, Al3+, etc.
For example, our group synthesized CdS QDs, ZnS QDs, ZnSe QDs, ZnO QDs, and ZnS/ZnO QDs, which have been used in Co2+, Cr3+, Al3+, Cr6+, and Ag+ detection [99,139,140,141,142,143,144]. Wang’s group [144] presented L-aspartic acid capped CdS QDs (L-Asp@CdS QDs) as sensors for the detection of Ag+. Among various common metal ions, the fluorescence intensity of QDs could only be increased by incorporating Ag+, which can be attributed to the formation of Ag2S on the L-Asp@CdS QDs surface, and then reducing the non-radiative electron/hole recombination process. Xia and coworkers [145] observed the fluorescence responses of CdTe QDs toward Ag+, which could be related to the particle size of QDs. In this study, for small sizes, the fluorescence enhancement was observed after addition a lower concentrations of Ag+, but with the further increase of Ag+ concentration, it showed significant fluorescence quenching. However, for larger sizes, the quenching was only observed with Ag+ incorporation. The above phenomena can be attributed to the smaller QDs typically having more surface defects than larger QDs, and the addition of Ag+ showed surface passivation actions toward QDs. Moreover, for larger QDs, the non-radiative recombination dominated the whole process because they had fewer surface defects. Similarly, Chen et al. [146] developed based-CdTe QDs sensors as a portable device for Ag+ detection relying on surface passivation and electron transfer, resulting from the interaction between Ag+ and Te2−.
Furthermore, fluorescence enhancement can also occur with the detection of Cd2+, Al3+, and Zn2+. Liu et al. [147] designed a “turn on” ratiometric fluorescent probe based on AgInZnS QDs and NGQDs and used them in the detection for Cd2+. This sensor showed a linear range of 0.5–100 µM with an LOD of 28.6 nM. The fluorescence enhancement of QDs could be due to their adsorption on the surfaces of QDs and the formation of a CdS passivation layer, which can be ascribed to the electrostatic interaction and strong affinity between Cd2+ and S2- on the QD surface. Simultaneously, a sensitive and selective probe, developed by Xu et al. and based on CdTe QDs, was used for the detection of Zn2+ and Cd2+. [148]. ZnS/ZnO QDs were synthesized as fluorescent probes for the detection of Al3+ based on the “turn-on” fluorescence sensing type [143]. With this sensor, Al3+ can be adsorbed on the surfaces of QDs owing to the interaction between Al3+ and the -OH of L-Cys, which regulates the charge transfer between ZnO QDs and L-Cys, thus enhancing the fluorescence of ZnS/ZnO QDs.
Yan et al. [149] reported a rapid and sensitive “turn-on” fluorescence sensor for detecting Cd2+ based on N, B-doped CQDs, which may be ascribed to the formation of an N, B-CQDs/Cd2+ complex, resulting in reduced internal charge transfer (ICT) and improved photo-generated electron transfer. Based on a similar co-doping system introduced for detection of Cr6+ by Luo’s group [150]. Meaningfully, the prepared QDs showed triple emission of purple, blue, and green fluorescence under the different excitation of 300, 330, and 490 nm, respectively. Yang’s group [151] developed sensitive fluorescent probes relying on synthetic N/Cl co-doped CDs for rapid testing of Cr6+ based on the combined action of DQE and IFE. Based on a similar co-doping system, Wang et al. [152] synthesized S,N-CQDs as sensors for the detection of Cr3+. The quenching caused by Cr3+ could be due to the strong interaction between Cr3+ and certain groups, such as -COOH and -OH groups, in the S, N-CQDs’ structures. This sensor obtained a low LOD of 6 µM. Mohamed et al. [153] reported MPA-capped CdS QDs as sensors for sensitive detection and quantification of Co2+ based on static fluorescence quenching due to the coordination between QDs and Co2+ via -COOH on the surfaces of QDs caused the resonance energy transfer from QDs to Co2+. Zikalala and coworkers [154] synthesized TGA capped ZnInS QDs used in the detection of Co2+, and this probe exhibited a linear range from 0.1 to 100 µM with an LOD of 99 nM. In this case, the fluorescence quenching could be caused by aggregation of the QDs resulting from the formation of surface ligand–Co2+ complex. Song et al. [155] investigated N, S co-doped CQDs as fluorescent sensors for the detection of Fe2+ and Al3+. In this system, the fluorescence intensity of QDs increased with increasing Fe2+ due chelation, while Al3+ could contribute to aggregation through combination with the N/S-CDs and lead to an aggregation-induced enhancement of emissions. Interestingly, the sensor revealed a dual-mode detection for Fe2+ via visual recognition and fluorescence turn-on.
In recent years, several fluorescent sensors have been applied for the determination of Co2+, Zn2+, Cd2+, Cr6+, Cr3+, Ag+, and Al3+ based on QDs such as SCQDs, CQDs, and GQDs. A summary of fluorescent sensors for metal ions, including the type of QDs, sensitivity, linearity range, LOD, and detection mechanism, is provided in Table 5.

4.2. QDs as Fluorescence Sensor for Pesticide Residues

Pesticides, such as insecticides, herbicides, and fungicides, can effectively control diseases and pests, kill weeds, and improve crop production and quality. However, the excessive use of pesticides can cause serious harm to human health and the ecological environment. Pesticides not only can directly remain on the surfaces of fruits and vegetables, causing various diseases after consumption but also can migrate to water or remain in soil, damaging the ecological environment. Hence, accurate, convenient, and efficient monitoring of pesticide residues in food and the environment is of paramount importance. Recently, the detection of different pesticide residues has been a research hotspot in the development of fluorescent QDs [36,163,164,165]. Based on a mechanism similar to that of heavy metal ion detection, the fluorescence of QDs can be changed by pesticide residues due to the ACE, FRET, SQE, DQE, IFE, and PET processes. In addition, the selectivity and sensitivity of QD-based sensor can be greatly improved by surface modification or design optimization of QDs.
Table 6 displays various QD-based sensors for pesticide residue detection. To date, various pesticides can be accurately analyzed by observing the effects of signal quenching, enhancement, or ratiometric responses via fluorescence.

4.2.1. Single-Irradiation Fluorescence Sensor

Generally, the sensing of QDs is based on the physical or chemical reactions between QDs and target analytes, which can result in photoluminescence enhancement or quenching. At present, quantum dots as fluorescent probes are widely used in analytical chemistry for simple, easy, and in-situ pesticides detection, which is significantly superior to traditional analytical methods.
Cheng et al. [170] proposed CdSe QDs as “turn-off” fluorescence sensor for detection of malathion based on ligand exchange between the surface capping molecules of CdSe QDs and malathion in organic solutions. Malathion could replace the surface ligands of CdSe QDs, resulting in the disruption of previous surface passivation and forming more surface defects. As a result, the fluorescence of QDs was quenched by malathion. The specificity of the sensor was investigated in the presence of several organic solvents and pesticide emulsifier, indicating that these substances do not interfere in the detection of malathion. Malathion has been quantifiable in concentrations ranging from 0.908 to 303 nM, with an LOD of 0.303 nM. Moreover, the applicability of the sensor has been explored for malathion detection in food samples.
Yang et al. [171] synthesized CdSe/ZnS core-shell QDs for phosphorothiolate detection based on an electron transfer effect. Thiophosphorus hydrolysate can be formed via the alkaline hydrolysis of phosphorothiolates. In this study, the fluorescence of the proposed sensor could be strongly quenched based on the direct coordinative interactions between the alkaline hydrolyzates of organophosphorus pesticides and the surface metal of the QDs. Moreover, the detection limit of the proposed sensor was found to be 0.967 ng/mL, whereas the linearity range was found to be 0.0001 to 160 µg/mL. The interference of the proposed sensor has been investigated using a mixture of ethion, including glucose, glycine, bovine serum albumin, vitamin C, vitamin E, cholesterol, K+, and Mg2+, demonstrating that these compounds do not significantly interfere in the detecting of phosphorothionates. The sensor has been employed for the detection of ethion residues in fortified tomatoes, suggesting that it has broad application potential in food analysis.
Nair et al. [172] reported S-doped GQDs as fluorescence probes, utilizing specific recognition and binding characteristics of aptamers for the ultrasensitive determination of omethoate based on the fluorescent turn-on mode. In this sensing strategy, the present of aptamers could include the ACQ effect of S-GQD due to the formation of the S-GQD-aptamer complex, while the incorporation could cause the disaggregation of the S-GQD-aptamer complex, resulting in the fluorescence recovery of S-GQD. Furthermore, based on a turn-off fluorescent detection mode, this research group designed an ultrasensitive sensor for carbamate pesticides in real samples with ppb level sensitivity [173]. Intriguingly, a flexible solid-state fluorescence sensing platform was prepared by incorporating S-GQD into polyvinyl alcohol matrix.

4.2.2. Ratiometric Fluorescence Sensors

Although single-irradiation fluorescence sensors have been widely reported and used in trace analyte detection, the analytical results are easily affected by the concentration of sensing materials and incident light drift. Compared with the single-irradiation fluorescence sensors based on the change in fluorescence signals, the ratiometric fluorescence probes rely on the ratio of two fluorescence signals, with a stronger self-regulation function. In addition, according to the change in fluorescence intensity and color of the fluorescent substance after adding the detection analytes, the visual detection of pollutants can also be realized.
Yang et al. [166] first designed the N, P-CQDs and Au NPs sensing system for carbendazim determination based on FRET. For instance, Au NPs could easily quench the fluorescence of N, P-CQDs resulting from FRET, while the fluorescence resumed exhibition after the incorporation of carbendazim as a result of the reduction in the fluorescence quenching of N, P-CQDs caused by the interaction between carbendazim and Au NPs. The sensor showed a linear response range within 0.005–0.16 µM and obtained a low LOD of 2 nM. The selectivity of the proposed sensor was investigated in the presence of various analytes, including common metal ions, thiamethoxam, acetamiprid, fluazinam, thriadimefon, dipterex, dursban, imidacloprid, and dinotefuran, suggesting that the sensor is resistant to interference substances. Furthermore, the proposed sensor has been successfully applied to detect carbendazim in fruit samples, demonstrating latent application for carbendazim monitoring.
Similarly, Wang’s group [167,168] developed a ratiometric fluorescence sensing system for the sensitive determination of carbendazim and dinotefuran based on CQDs. An N, CQDs/AuNC [167] ratiometric fluorescence probe was constructed for the detection of carbendazim by recording the ratio change between the fluorescence signal of N-CQDs and the second-order Rayleigh scattering (SRS) signal of N-CQDs/AuNCs probes. The selectivity of the proposed sensor was examined in the presence of various interfering cations and pesticides, suggesting that the sensor has higher selectivity than other pesticides because the strong affinities among AuNCs, aromatic configurations, and nitrogen atoms are conductive to the adsorption of carbendazim on the surfaces of AuNCs. Moreover, the practical application of the sensor for detecting carbendazim was evaluated in strawberries and apples, and the accuracy of the sensor was comparable to traditional HPLC analyses, with the extra advantage of being highly sensitive, simple, and rapid. As a result, it has enormous promise for carbendazim determinations in fruit samples. An S-CQD/CuNC [168] ratiometric fluorescence probe was also designed for the detection of dinotefuran, and a linear range was found between 10 µM and 500 µM with an LOD of 7.04 µM. Different from carbendazim sensing systems, this sensor was mainly based on IFEs between S-CQDs and CuNCs. The presence of dinotefuran could induce the aggregation of CuNCs, resulting in the IFEs, leading to the fluorescence recovery of S-CQDs and decayed CuNC fluorescence. Bera and Mohapatra [169] developed the fabrication of a CdTe-CQD integrated probe, which was used in glyphosate detection relying on PET between CdTe and CQDs. Compared with the sensing-based QDs and metal nanoparticles, the sensor is simple with high sensitivity, and it is suitable for real sample detection. Due to the high selectivity of glyphosate, this sensing system could achieve a low LOD of 2 pM within a broad linear response range of 0–1000 nM.
Recently, Chen and colleagues [174] constructed a ratiometric fluorescent sensing system for the detection of thiram and paraquat, and the LODs obtained were 7.49 nM and 3.03 nM, respectively. The applicability of the proposed sensor has been evaluated in several plant products including semen nelumbinis, coix lacryma-jobi, and ginseng. Furthermore, the distinct visual analysis for thiram and paraquat solutions with higher concentrations has been performed, suggesting that the sensor can be applied in the direct identification of pesticide varieties. In conclusion, the proposed sensor provides effective implementation for on-site monitoring of pesticide residues in food.

4.3. Colorimetric Sensor

The colorimetric sensor is a simple analytical device that depends on changes in color to measure analytes. The method is cost-effective and visualizes color changes, which allows for rapid visual evaluation without the need for precision instruments. Recently, due to the advantages of simple preparation, portability, fast reading, and convenience, the paper-based test strip was used as a promising sensing platform for easy and simple fluorescence detection.
Liu et al. [64] developed a ZnO QD-based test strip sensor for the determination of Cu2+. The Cu2+ test strips were prepared through a dipping–drying process. Under ultraviolet irradiation, the test papers emit yellow at different intensities due to the difference concentrations of Cu2+. Here, the test paper can qualitatively judge the Cu2+ concentration within the range of 40–1000 μM by the naked eye. Furthermore, quantitative detection is achieved by analyzing the gray level of transformed RGB images of test strips based on an ordinary camera and computer. Subsequently, the proposed sensor provided linearity ranges from 10 to 1000 µM (R2 = 0.9763).
Hao et al. [96] proposed a test strip sensor for Hg2+ based on CQDs. In this study, the visual LOD of test strip sensors for Hg2+ can be determined to be 0.1 μM. Moreover, the test strip sensor can be recovered by EDTA and reused at least three times. Similarly, Sebastian et al. [110] developed paper strips for the rapid assessing of Pb2+ based on DTPAN-fn-GQD. Consequently, the combination of fluorescent materials with paper substrates is expected to be used to develop a portable and convenient platform for the qualitative and quantitative detection of metal ions or other analytes.

5. Conclusions and Outlook

Recently, various QD-based sensors have been developed for the analysis of heavy metal ions and pesticide residues in water and food samples, and significant progress has been made in the field of nanomaterials. Fluorescent, ratiometric, and colorimetric methods are commonly used. Among these methods, fluorescence sensors have high sensitivity and selectivity, rapid responses, and simplicity. The specific selectivity of QDs for the detection of heavy metal ions and pesticide residues can be designed based on the surface functionalization of QDs. Ratiometric fluorescence sensors can eliminate the interference of environmental elements and improve detection accuracy by self-tuning the two fluorescence signals. The development of ratiometric fluorescence sensors provides feasibility for multiple detection. Colorimetric sensors can realize visual detection. Combining paper test strips with a smartphone platform enables on-site rapid testing. However, compared with fluorescence and ratiometric fluorescence sensors, the accuracy of colorimetric sensors should be improved. With the development of nanomaterials and technology, sensing performance, including the LOD, sensitivity, and selectivity of the sensor, will be effectively improved. We believe that the QD-based sensors have great potential application in environmental monitoring, food safety, agriculture, and biomedicine.

Author Contributions

Writing—original draft and investigation: Z.W.; investigation, B.Y.; investigation, Y.X.; investigation, X.T.; supervision, writing—reviewing and editing, Y.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Nos. 41876055 and 61761047), the Yunnan Provincial Department of Science and Technology through the Key Project for Science and Technology (Grant No. 2017FA025), and the Project of the Department of Education of Yunnan Province (2023Y0263).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (Nos. 41876055 and 61761047), the Yunnan Provincial Department of Science and Technology through the Key Project for Science and Technology (Grant No. 2017FA025), and the Project of the Department of Education of Yunnan Province (2023Y0263).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sonali, J.M.I.; Subhashree, S.; Kumar, P.S.; Gayathri, K.V. New analytical strategies amplified with carbon-based nanomaterial for sensing food pollutants. Chemosphere 2022, 295, 133847. [Google Scholar]
  2. Lan, L.Y.; Yao, Y.; Ping, J.F.; Ying, Y.B. Recent progress in nanomaterial-based optical aptamer assay for the detection of food chemical contaminants. ACS Appl. Mater. Interfaces 2017, 9, 23287–23301. [Google Scholar] [CrossRef] [PubMed]
  3. Zhang, K.; Li, H.Y.; Wang, W.J.; Cao, J.X.; Gan, N.; Han, H.Y. Application of multiplexed aptasensors in food contaminants detection. ACS Sens. 2021, 5, 3721–3738. [Google Scholar] [CrossRef] [PubMed]
  4. Xu, M.M.; Wang, X.Y.; Liu, X.P. Detection of heavy metal ions by ratiometric photoelectric sensor. J. Agric. Food Chem. 2022, 70, 11468–11480. [Google Scholar] [CrossRef] [PubMed]
  5. Huang, Y.F.; Peng, J.H.; Huang, X.J. Allylthiourea functionalized magnetic adsorbent for the extraction of cadmium, copper and lead ions prior to their determination by atomic absorption spectrometry. Microchim. Acta 2019, 186, 5. [Google Scholar] [CrossRef]
  6. Mir, S.A.; Dar, B.N.; Mir, M.; Sofi, S.A.; Shah, M.A.; Sidiq, T.; Sunooj, K.V.; Hamdani, A.M.; Khaneghah, A.M. Current strategies for the reduction of pesticide residues in food products. J. Food Compos. Anal. 2022, 106, 104274. [Google Scholar] [CrossRef]
  7. Panda, S.; Jadav, A.; Panda, N.; Mohapatra, S. A novel carbon quantum dot-based fluorescent nanosensor for selective detection of flumioxazin in real samples. New J. Chem. 2018, 42, 2074–2080. [Google Scholar] [CrossRef]
  8. Xu, L.Y.; Abd El-Aty, A.M.; Eun, J.B.; Shim, J.H.; Zhao, J.; Lei, X.M.; Gao, S.; She, Y.X.; Jin, F.; Wang, J.; et al. Recent advances in rapid detection techniques for pesticide residue:A review. J. Agric. Food Chem. 2022, 70, 13093–13117. [Google Scholar] [CrossRef]
  9. May, B.M.; Bambo, M.F.; Hosseini, S.S.; Sidwaba, U.; Nxumalo, E.N.; Mishra, A.K. A review on I-III-VI ternary quantum dots for fluorescence detection of heavy metals ions in water: Optical properties, synthesis and application. RSC Adv. 2022, 12, 11216–11232. [Google Scholar] [CrossRef]
  10. Wu, Q.H.; Li, Y.P.; Wang, C.; Liu, Z.M.; Zang, X.H.; Zhou, X.; Wang, Z. Dispersive liquid–liquid microextraction combined with high performance liquid chromatography–fluorescence detection for the determination of carbendazim and thiabendazole in environmental samples. Anal. Chim. Acta 2009, 638, 139–145. [Google Scholar] [CrossRef]
  11. Chen, X.; Lin, M.S.; Sun, L.; Xu, T.T.; Lai, K.Q.; Huang, M.Z.; Lin, H.T. Detection and quantification of carbendazim in Oolong tea by surface-enhanced Raman spectroscopy and gold nanoparticle substrates. Food Chem. 2019, 293, 271–277. [Google Scholar] [CrossRef] [PubMed]
  12. Gao, X.Y.; Gao, Y.; Bian, C.C.; Ma, H.Y.; Liu, H.L. Electroactive nanoporous gold driven electrochemical sensor for the simultaneous detection of carbendazim and methyl parathion. Electrochim. Acta 2019, 310, 78–85. [Google Scholar] [CrossRef]
  13. Mahdavi, V.; Ghorbani-Paji, F.; Ramezani, M.K.; Ghassempour, A.; Aboul-Enein, H.Y. Dissipation of carbendazim and its metabolites in cucumber using liquid chromatography tandem mass spectrometry. Int. J. Environ. Anal. Chem. 2019, 99, 968–976. [Google Scholar] [CrossRef]
  14. Zhao, G.; Wang, X.; Liu, G.; Thuy, N.T.D. A disposable and flexible electrochemical sensor for the sensitive detection of heavy metals based on a one-step laser-induced surface modification: A new strategy for the batch fabrication of sensors. Sens. Actuator B Chem. 2022, 350, 130834. [Google Scholar] [CrossRef]
  15. Wu, P.; Zhao, T.; Wang, S.L.; Hou, X.D. Semicondutor quantum dots-based metal ion probes. Nanoscale 2014, 6, 43–64. [Google Scholar] [CrossRef]
  16. Kazemifard, N.; Ensafi, A.A.; Dehkordi, Z.S. A review of the incorporation of QDs and imprinting technology in optical sensors-imprinting methods and sensing responses. New J. Chem. 2021, 45, 10170–10198. [Google Scholar] [CrossRef]
  17. Sharma, G.; Kumar, A.; Naushad, M.; Kumar, A.; Al-Muhtaseb, A.H.; Dhiman, P.; Ghfar, A.A.; Stadler, F.J.; Khan, M.R. Photoremediation of toxic dye from aqueous environment using monometallic and bimetallic quantum dots-based nanocomposites. J. Clean. Prod. 2018, 172, 2919–2930. [Google Scholar] [CrossRef]
  18. Iga, A.M.; Robertson, J.H.P.; Winslet, M.C.; Seifalian, A.M. Clinical potential of quantum dots. J. Biomed. Biotechnol. 2007, 1, 76087. [Google Scholar] [CrossRef] [Green Version]
  19. Farzin, M.A.; Abdoos, H. A critical review on quantum dots: From synthesis toward applications in electrochemical biosensors for determination of disease-related biomolecules. Talanta 2021, 224, 121828. [Google Scholar] [CrossRef]
  20. Shellaiah, M.; Sun, K.W. Review on sensing applications of perovskite nanomaterials. Chemosensors 2020, 8, 55. [Google Scholar] [CrossRef]
  21. Sanjayan, C.G.; Jyothi, M.S.; Balakrishna, R.G. Stabilization of CsPbBr3 quantum dots for photocatalysis, imaging and optical sensing in water and biological medium: A review. J. Mater. Chem. C 2022, 10, 6935–6956. [Google Scholar]
  22. Ren, X.X.; Zhang, X.; Xie, H.X.; Cai, J.H.; Wang, C.H.; Chen, E.G.; Xu, S.; Ye, Y.; Sun, J.; Yan, Q.; et al. Perovskite quantumdots for emerging displays: Recent progress and perspectives. Nanomaterials 2022, 12, 2243. [Google Scholar] [CrossRef] [PubMed]
  23. Rakshit, S.; Piatkowski, P.; Mora-Sero, I.; Douhal, A. Combining perovskites and quantum dots: Synthesis, characterization, and applications in solar cells, LEDs, and photodetectors. Adv. Opt. Mater. 2022, 10, 2102566. [Google Scholar] [CrossRef]
  24. Shellaiah, M.; Awasthi, K.; Chandran, S.; Aazaad, B.; Sun, K.W.; Ohta, N.; Wu, S.P.; Lin, M.C. Methylammonium tin tribromide quantum dots for heavy metal ion detection and cellular imaging. ACS Appl. Nano Mater. 2022, 5, 2859–2874. [Google Scholar] [CrossRef]
  25. Ai, B.; Fan, Z.W.; Wong, Z.J. Plasmonic-perovskite solar cells, light emitters, and sensors. Microsyst. Nanoeng. 2022, 8, 5. [Google Scholar] [CrossRef]
  26. Yang, Y.Y.; Han, A.L.; Hao, S.J.; Li, X.; Luo, X.Y.; Fang, G.Z.; Liu, J.F.; Wang, S. Fluorescent methylammonium lead halide perovskite quantum dots as a sensing material for the detection of polar organochlorine pesticide residues. Analyst 2022, 147, 3258. [Google Scholar] [CrossRef]
  27. Kaur, A.; Pandey, K.; Kaur, R.; Vashishat, N.; Kaur, M. Nanocomposites of carbon quantum dots and graphene quantum dots: Environmental applications as sensors. Chemosensors 2022, 10, 367. [Google Scholar] [CrossRef]
  28. Reshma, V.G.; Mohanan, P.V. Quantum dots: Applications and safety consequences. J. Lumin. 2018, 205, 287–298. [Google Scholar] [CrossRef]
  29. Resch-Genger, U.; Grabolle, M.; Cavaliere-Jaricot, S.; Nitschke, R.; Nann, T. Quantum dots versus organic dyes as fluorescent labels. Nat. Methods 2008, 5, 763–775. [Google Scholar] [CrossRef]
  30. Wang, J.R.; Yu, J.L.; Wang, X.Y.; Wang, L.Y.; Li, B.W.; Shen, D.Z.; Kang, Q.; Chen, L.X. Functional ZnS:Mn(II) quantum dot modified with L-cysteine and 6-mercaptonicotinic acid as a fluorometric probe for copper (II). Microchim. Acta 2018, 185, 420. [Google Scholar] [CrossRef]
  31. Zhang, Y.J.; Liu, B.; Liu, Z.M.; Li, J.K. Research progress in the synthesis and biological application of quantum dots. New J. Chem. 2022, 46, 20515–20539. [Google Scholar] [CrossRef]
  32. Zhu, C.Y.; Chen, Z.; Gao, S.; Goh, B.L.; Bin Samsudin, I.; Lwe, K.W.; Wu, Y.L.; Wu, C.S.; Su, X.D. Recent advances in non-toxic quantum dots and their biomedical applications. Prog. Nat. Sci. Mater. Int. 2020, 29, 628–640. [Google Scholar] [CrossRef]
  33. Zou, L.; Gu, Z.H.; Sun, M. Review of the application of quantum dots in the heavy-metal detection. Toxicol. Environ. Chem. 2015, 97, 477–490. [Google Scholar] [CrossRef]
  34. Wang, X.Y.; Kong, L.B.; Zhou, S.Q.; Ma, C.Y.; Lin, W.C.; Sun, X.Y.; Kirsanov, D.; Legin, A.; Wan, H.; Wang, P. Development of QDs-based nanosensors for heavy metal detection: A review on transducer principles and in-situ detection. Talanta 2022, 239, 122903. [Google Scholar] [CrossRef] [PubMed]
  35. Mukherjee, S.; Bhattacharyya, S.; Ghosh, K.; Pal, S.; Halder, A.; Naseri, M.; Mohammadniaei, M.; Sarkar, S.; Ghosh, A.; Sun, Y.; et al. Sensory development for heavy metal detection: A review on translation from conventional analysis to field-portable sensor. Trends Food Sci. Technol. 2021, 109, 674–689. [Google Scholar] [CrossRef]
  36. Nsibande, S.A.; Forbes, P.B.C. Fluorescence detection of pesticides using quantum dot materials—A review. Anal. Chim. Acta 2016, 945, 9–22. [Google Scholar] [CrossRef] [PubMed]
  37. Nazir, F.; Asad, M.; Fatima, L.; Bokhari, A.; Majeed, S.; Fatima, B.; Mohammed, A.A.A.; Karri, R.R. Silica quantum dots; an optical nanosensing approach for trace detection of pesticides in environmental and biological samples. Environ. Res. 2023, 231, 116147. [Google Scholar] [CrossRef] [PubMed]
  38. Anand, K.V.; Subala, A.S.; Sumathi, K.S.; Merin, S.A.L. A review on the removal of dye, pesticide and pathogens from waste water using quantum dots. Eur. J. Adv. Chem. Res. 2020, 1, 14. [Google Scholar] [CrossRef]
  39. He, K.Y.; Wang, L.; Xu, X.H. Chapter 17—Chemical sensing of pesticides in water. In Advanced Sensor Technology; Elsevier: Amsterdam, The Netherlands, 2023; pp. 647–668. [Google Scholar]
  40. Caroleo, F.; Magna, G.; Naitana, M.L.; Di Zazzo, L.; Martini, R.; Pizzoli, F.; Muduganti, M.; Lvova, L.; Mandoj, F.; Nardis, S.; et al. Advances in optical sensors for persistent organic pollutant environmental monitoring. Sensors 2022, 22, 2649. [Google Scholar] [CrossRef] [PubMed]
  41. Chen, S.Y.; Yun, S.N.; Liu, Y.J.; Yu, R.J.; Tu, Q.; Wang, J.Y.; Yuan, M.S. A highly selective and sensitive CdS fluorescent quantum dot for the simultaneous detection of multiple pesticides. Analyst 2022, 147, 3258–3265. [Google Scholar] [CrossRef]
  42. Spanhel, L.; Haase, M.; Weller, H.; Henglein, A. Photochemistry of colloidal semiconductors surface modification and stability of strong luminescing CdS particles. J. Am. Chem. Soc. 1987, 109, 5649–5655. [Google Scholar] [CrossRef]
  43. Chen, Y.F.; Rosenzweig, Z. Luminescent CdS quantum dots as selective ion probes. Anal. Chem. 2002, 74, 5132–5138. [Google Scholar] [CrossRef] [PubMed]
  44. Zhu, J.; Zhao, Z.J.; Li, J.J.; Zhao, J.W. CdTe quantum dot-based fluorescent probes for selective detection of Hg (II): The effect of particle size. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2017, 177, 140–146. [Google Scholar] [CrossRef] [PubMed]
  45. Uppa, Y.; Kulchat, S.; Ngamdee, K.; Pradublai, K.; Tuntulani, T.; Ngeontae, W. Silver ion modulated CdS quantum dots for highly selective detection of trace Hg2+. J. Lumin. 2016, 178, 437–445. [Google Scholar] [CrossRef]
  46. Han, S.Q.; Liu, J.L.; Gan, Z.G.; Liang, J.G.; Zhao, S.M. Application of luminescent BSA-capped CdS quantum dots as a fluorescence probe for the detection of Cu2+. J. Chin. Chem. Soc.-Taip. 2008, 55, 1069–1073. [Google Scholar] [CrossRef]
  47. Shellaiah, M.; Sun, K.W. Review on carbon dot-based fluorescent detection of biothiols. Biosensors 2023, 13, 335. [Google Scholar] [CrossRef]
  48. Khan, Z.G.; Patil, P.O. A comprehensive review on carbon dots and graphene quantum dots based fluorescent sensor for biothiols. Microchem. J. 2020, 157, 105011. [Google Scholar] [CrossRef]
  49. Gonzalez-Gonzalez, R.B.; Morales-Murillo, M.B.; Martinez-Prado, M.A.; Melchor-Martinez, E.M.; Ahmed, I.; Bilal, M.; Parra-Saldivar, R.; Iqbal, H.M.N. Carbon dots-based nanomaterials for fluorescent sensing of toxic elements in environmental samples: Strategies for enhanced performance. Chemosphere 2022, 300, 134515. [Google Scholar] [CrossRef]
  50. Kamat, P.V. Photochemistry on nonreactive and reactive (semiconductor) surfaces. Chem. Rev. 1993, 93, 267–300. [Google Scholar] [CrossRef]
  51. Mehta, V.N.; Desai, M.L.; Basu, H.; Singhal, R.K.; Kailasa, S.K. Recent developments on fluorescent hybrid nanomaterials for metal ions sensing and bioimaging applications: A review. J. Mol. Liq. 2021, 333, 115950. [Google Scholar] [CrossRef]
  52. Ghali, M. Static quenching of bovine serum albumin conjugated with small size CdS nanocrystalline quantum dots. J. Lumin. 2010, 130, 1254–1257. [Google Scholar] [CrossRef]
  53. Xu, Z.Y.; Wang, Y.Z.; Zhang, J.R.; Shi, C.; Yang, X.T. A highly sensitive and selective fluorescent probe using MPA-InP/ZnS QDs for detection of trace amounts of Cu2+ in water. Foods 2021, 10, 2777. [Google Scholar] [CrossRef] [PubMed]
  54. Lee, S.; Kang, S.H. Wavelength-dependent metal-enhanced fluorescence biosensors via resonance energy transfer modulation. Biosensors 2023, 13, 376. [Google Scholar] [CrossRef]
  55. Wang, K.; Dong, E.F.; Fang, M.; Zhu, W.J.; Li, C. Construction of hybrid fluorescent sensor for Cu2+ detection using fluorescein- functionalized CdS quantum dots via FRET. J. Fluoresc. 2022, 32, 1099–1107. [Google Scholar] [CrossRef] [PubMed]
  56. Huang, D.W.; Niu, C.G.; Ruan, M.; Wang, X.Y.; Zeng, G.M.; Deng, C.H. Highly sensitive strategy for Hg2+ detection in environmental water samples using long lifetime fluorescence quantum dots and gold nanoparticles. Environ. Sci. Technol. 2013, 47, 4392–4398. [Google Scholar] [CrossRef]
  57. Wang, J.; Song, F.J.; Ai, Y.L.; Hu, S.W.; Huang, Z.Z.; Zhong, W.Y. A simple FRET system using two-color CdTe quantum dots assisted by cetyltrimethylammonium bromide and its application to Hg(II) detection. Luminescence 2019, 34, 205–211. [Google Scholar] [CrossRef]
  58. Li, P.; Luo, C.; Chen, X.X.; Huang, C.B. An off–on fluorescence aptasensor for trace thrombin detection based on FRET between CdS QDs and AuNPs. RSC Adv. 2022, 12, 35763–35769. [Google Scholar] [CrossRef]
  59. Zhang, K.; Zhang, J.M. A fluorescent probe for the detection of Hg2+ based on rhodamine derivative and modified CdTe quantum dots. Res. Chem. Intermed. 2020, 46, 987–997. [Google Scholar] [CrossRef]
  60. Liu, B.Y.; Fang, Z.; Wu, G.F.; Wu, S.Z. Nanoparticles as scaffolds for FRET-based ratiometric detection of mercury ions in water with QDs as donors. Analyst 2012, 137, 3717. [Google Scholar] [CrossRef]
  61. Zhao, R.J.; Wang, Z.Z.; Tian, X.; Shu, H.; Yang, Y.; Xiao, X.C.; Wang, Y.D. Excellent fluorescence detection of Cu2+ in water system using N-acetyl-L-cysteines modified CdS quantum dots as fluorescence probe. Nanotechnology 2021, 32, 405707. [Google Scholar] [CrossRef]
  62. Wang, Z.Z.; Xiao, X.C.; Zou, T.; Yang, Y.; Xing, X.X.; Zhao, R.Z.; Wang, Z.D.; Wang, Y.D. Citric acid capped CdS quantum dots for fluorescence detection of copper ions (II) in aqueous solution. Nanomaterials 2019, 9, 32. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Zou, T.; Xing, X.X.; Yang, Y.; Wang, Z.Z.; Wang, Z.D.; Zhao, R.Z.; Zhang, X.; Wang, Y.D. Water-soluble ZnO quantum dots modified by (3-aminopropyl) triethoxysilane: The promising fluorescent probe for the selective detection of Cu2+ ion in drinking water. J. Alloys Compd. 2020, 825, 153904. [Google Scholar] [CrossRef]
  64. Liu, X.; Yang, Y.; Xing, X.X.; Wang, Y.D. Grey level replaces fluorescent intensity: Fluorescent paper sensor based on ZnO nanoparticles for quantitative detection of Cu2+ without photoluminescence spectrometer. Sens. Actuators B Chem. 2018, 255, 2356–2366. [Google Scholar] [CrossRef]
  65. Liu, J.M.; Lin, L.P.; Wang, X.X.; Lin, S.Q.; Cai, W.L.; Zhang, L.H.; Zheng, Z.Y. Highly selective and sensitive detection of Cu2+ with lysine enhancing bovine serum albumin modified-carbon dots fluorescent probe. Analyst 2012, 137, 2637. [Google Scholar] [CrossRef]
  66. Dong, Y.Q.; Wang, R.X.; Li, G.; Chen, C.Q.; Chi, Y.W.; Chen, G.N. Polyamine-functionalized carbon quantum dots as fluorescent probes for selective and sensitive detection of copper ions. Anal. Chem. 2012, 84, 6220–6224. [Google Scholar] [CrossRef]
  67. Wang, Z.M.; Zhou, C.; Wu, S.W.; Sun, C.Y. Ion-imprinted polymer modified with carbon quantum dots as a highly sensitive copper (II) ion probe. Polymers 2021, 13, 1376. [Google Scholar] [CrossRef]
  68. Jiang, X.M.; Kou, Y.X.; Lu, J.J.; Xue, Y.Y.; Wang, M.J.; Tian, B.W.; Tan, L. Fluorescence “on-off-on” assay of copper ions and EDTA using amino-functionalized graphene quantum dots. J. Fluoresc. 2020, 30, 301–308. [Google Scholar] [CrossRef]
  69. Liu, Y.F.; Zhu, T.; Deng, M.; Tang, X.S.; Han, S.; Liu, A.P.; Bai, Y.Z.; Qu, D.R.; Huang, X.B.; Qiu, F. Selective and sensitive detection of copper (II) based on fluorescent zinc-doped AgInS2 quantum dots. J. Lumin. 2018, 201, 182–188. [Google Scholar] [CrossRef]
  70. Jiang, P.; Li, S.L.; Ha, M.L.; Liu, Y.; Chen, Z.L. Biocompatible Ag2S quantum dots for highly sensitive detection of copper ions. Analyst 2019, 144, 2604–2610. [Google Scholar] [CrossRef]
  71. Zhou, Q.X.; Li, Z.F.; Wang, Q.F.; Peng, L.; Luo, S.H.; Gu, F.L. Polymer-capped CdSe/ZnS quantum dots for the sensitive detection of Cu2+ and Hg2+ and the quenching mechanism. Anal. Methods 2021, 13, 2305–2312. [Google Scholar] [CrossRef]
  72. Xie, Y.F.; Jiang, Y.J.; Zou, H.Y.; Wang, J.; Huang, C.Z. Discrimination of copper and silver ions based on the label-free quantum dots. Talanta 2020, 220, 121430. [Google Scholar] [CrossRef] [PubMed]
  73. Liu, Y.J.; Zhu, Y.; Liu, X.Y.; Dong, L.M.; Zheng, Q.L.; Kang, S.; He, Y.H.; Wang, J.; Abd El-Aty, A.M. CdSe/ZnS QDs embedded polyethersulfone fluorescence composite membrane for sensitive detection of copper ions in various drinks. J. Environ. Sci. Health Part B 2023, 58, 120–123. [Google Scholar] [CrossRef] [PubMed]
  74. Passos, S.G.B.; Kunst, T.H.; Freitas, D.V.; Navarro, M. Paired electrosynthesis of ZnSe/ZnS quantum dots and Cu2+ detection by fluorescence quenching. J. Lumin. 2020, 228, 117611. [Google Scholar] [CrossRef]
  75. Omer, K.M. Highly passivated phosphorous and nitrogen co-doped carbon quantum dots and fluorometric assay for detection of copper ions. Anal. Bioanal. Chem. 2018, 410, 6331–6336. [Google Scholar] [CrossRef] [PubMed]
  76. Liao, S.; Huang, X.Q.; Yang, H.; Chen, X.Q. Nitrogen-doped carbon quantum dots as a fluorescent probe to detect copper ions, glutathione, and intracellular pH. Anal. Bioanal. Chem. 2018, 410, 7701–7710. [Google Scholar] [CrossRef]
  77. El-Malla, S.F.; Elshenawy, E.A.; Hammad, S.F.; Mansour, F.R. Rapid microwave synthesis of N, S-doped carbon quantum dots as a novel turn off-on sensor for label-free determination of copper and etidronate disodium. Anal. Chim. Acta 2022, 1197, 339491. [Google Scholar] [CrossRef]
  78. Preethi, M.; Viswanathan, C.; Ponpandian, N. Fluorescence quenching mechanism of P-doped carbon quantum dots as fluorescent sensor for Cu2+ ions. Colloids Surf. A Physicochem. Eng. Asp. 2022, 653, 129942. [Google Scholar] [CrossRef]
  79. Gao, B.; Chen, D.; Gu, B.L.; Wang, T.; Wang, Z.H.; Xie, F.; Yang, Y.S.; Guo, Q.L.; Wang, G. Facile and highly effective synthesis of nitrogen-doped graphene quantum dots as a fluorescent sensing probe for Cu2+ detection. Curr. Appl. Phys. 2020, 20, 538–544. [Google Scholar] [CrossRef]
  80. Wang, F.X.; Gu, Z.Y.; Lei, W.; Wang, W.J.; Xia, X.F.; Hao, Q.L. Graphene quantum dots as a fluorescent sensing platform for highly efficient detection of copper(II) ions. Sens. Actuators B Chem. 2014, 190, 516–522. [Google Scholar] [CrossRef]
  81. Liu, X.; Han, J.; Hou, X.D.; Altincicek, F.; Oncel, N.; Pierce, D.; Wu, X.; Zhao, J.X. One-pot synthesis of graphene quantum dots using humic acid and its application for copper (II) ion detection. J. Mater. Sci. 2021, 56, 4991–5005. [Google Scholar] [CrossRef]
  82. Yao, J.; Wang, L. Graphene quantum dots as nanosensor for rapid and label-free dual detection of Cu2+ and tiopronin by means of fluorescence “on-off-on” switching: Mechanism and molecular logic gate. New J. Chem. 2021, 45, 20649–20659. [Google Scholar] [CrossRef]
  83. Li, W.T.; Jiang, N.J.; Zhang, L.M.; Chen, Y.Q.; Gao, J.; Zhang, J.H.; Yang, B.S.; He, J.X. Facile synthesis of aminated graphene quantum dots for promising and selective detection of cobalt and copper ions in aqueous media. Molecules 2022, 27, 7844. [Google Scholar] [CrossRef] [PubMed]
  84. Tam, T.V.; Choi, W.M. One-pot synthesis of highly fluorescent amino-functionalized graphene quantum dots for effective detection of copper ions. Curr. Appl. Phys. 2018, 18, 1255–1260. [Google Scholar] [CrossRef]
  85. Choudhary, Y.S.; Nageswaran, G. Branched mercapto acid capped CdTe quantum dots as fluorescence probes for Hg2+ detection. Sens. Bio-Sens. Res. 2019, 23, 100278. [Google Scholar] [CrossRef]
  86. Kaur, J.; Komal; Renu; Kumar, V.; Tikoo, K.B.; Bansal, S.; Kaushik, A.; Singhal, S. Glutathione modified fluorescent CdS QDs synthesized using environmentally benign pathway for detection of mercury ions in aqueous phase. J. Fluoresc. 2020, 30, 773–785. [Google Scholar] [CrossRef] [PubMed]
  87. You, J.Q.; Yan, D.; He, Y.; Zhou, J.G.; Ge, Y.L.; Song, G.W. Polyethyleneimine-protected Ag2S quantum dots for near-infrared fluorescence-enhanced detection of trace-level Hg2+ in water. J. Water Chem. Technol. 2020, 42, 36–44. [Google Scholar] [CrossRef]
  88. Granados-Oliveros, G.; Pineros, B.S.G.; Calderon, F.G.O. CdSe/ZnS quantum dots capped with oleic acid and L-glutathione: Structural properties and application in detection of Hg2+. J. Mol. Struct. 2022, 1254, 132293. [Google Scholar] [CrossRef]
  89. Yang, Y.; Xiao, X.C.; Xing, X.X.; Wang, Z.Z.; Zou, T.; Wang, Z.D.; Wang, Y.D. One-pot synthesis of N-doped graphene quantum dots as highly sensitive fluorescent sensor for detection of mercury ions water solutions. Mater. Res. Express 2019, 6, 095615. [Google Scholar] [CrossRef]
  90. Yang, Y.; Zou, T.; Wang, Z.Z.; Xing, X.X.; Peng, S.J.; Zhao, R.J.; Zhang, X.; Wang, Y.D. The fluorescent quenching mechanism of N and S co-doped graphene quantum dots with Fe3+ and Hg2+ ions and their application as a novel fluorescent sensor. Nanomaterials 2019, 9, 738. [Google Scholar] [CrossRef] [Green Version]
  91. Yang, Y.; Xiao, X.C.; Xing, X.X.; Wang, Z.Z.; Zou, T.; Wang, Z.D.; Zhao, R.J.; Wang, Y.D. Rhodamine B assisted graphene quantum dots flourescent sensor system for sensitive recognition of mercury ions. J. Lumin. 2019, 207, 273–281. [Google Scholar] [CrossRef]
  92. Khose, R.V.; Chakraborty, G.; Bondarde, M.P.; Wadekar, P.H.; Ray, A.K.; Some, S. Red-fluorescent graphene quantum dots from guava leaf as a turn-off probe for sensing aqueous Hg(II). New J. Chem. 2021, 45, 4617–4625. [Google Scholar] [CrossRef]
  93. Llaver, M.; Barrionuevo, S.D.; Prieto, E.; Wuilloud, R.G.; Ibanez, F.J. Functionalized graphene quantum dots obtained from graphene foams used for highly selective detection of Hg2+ in real samples. Anal. Chim. Acta 2022, 1232, 340422. [Google Scholar] [CrossRef] [PubMed]
  94. Abdolmohammad-Zadeh, H.; Azari, Z.; Pourbasheer, E. Fluorescence resonance energy transfer between carbon quantum dots and silver nanoparticles: Application to mercuric ion sensing. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2021, 245, 118924. [Google Scholar] [CrossRef] [PubMed]
  95. Dhandapani, E.; Maadeswaran, P.; Raj, R.M.; Raj, V.; Kandiah, K.; Duraisamy, N. A potential forecast of carbon quantum dots (CQDs) as an ultrasensitive and selective fluorescence probe for Hg (II) ions sensing. Mater. Sci. Eng. B 2023, 287, 116098. [Google Scholar] [CrossRef]
  96. Hao, Y.Q.; Yu, L.Y.; Li, T.T.; Chen, L.H.; Han, X.; Chai, F. The synthesis of carbon dots by folic acid and utilized as sustainable probe and paper sensor for Hg2+ sensing and cellular imaging. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2023, 285, 121865. [Google Scholar] [CrossRef]
  97. Dewangan, L.; Chawre, Y.; Korram, J.; Karbhal, I.; Nagwanshi, R.; Jain, V.; Satnami, M.L. N-doped, silver, and cerium co-doped carbon quantum dots based sensor for detection of Hg2+ and captopril. Microchem. J. 2022, 182, 107867. [Google Scholar] [CrossRef]
  98. Qi, H.Y.; Sun, X.N.; Jing, T.; Li, J.L.; Li, J. Integration detection of mercury(II) and GSH with a fluorescent “on-off-on” switch sensor based on nitrogen, sulfur co-doped carbon dots. RSC Adv. 2022, 12, 1989–1997. [Google Scholar] [CrossRef]
  99. Xing, X.X.; Yang, Y.; Zou, T.; Wang, Z.Z.; Wang, Z.D.; Zhao, R.J.; Zhang, X.; Wang, Y.D. Thioglycolic acid-capped ZnSe quantum dots as nanoprobe for cobalt(II) and iron(III) via measurement of grey level, UV-vis spectra and dynamic light scattering. Microchim. Acta 2019, 186, 444. [Google Scholar] [CrossRef]
  100. Ma, J.; Yu, H.H.; Jiang, X.; Luo, Z.Z.; Zheng, Y. High sensitivity label-free detection of Fe3+ ion in aqueous solution using fluorescent MoS2 quantum dots. Sens. Actuators B Chem. 2019, 281, 989–997. [Google Scholar] [CrossRef]
  101. Huang, C.; Wang, H.W.; Xu, Y.J.; Ma, S.J.; Gong, B.L.; Ou, J.J. Carbon dot-functionalized macroporous adsorption resin for bifunctional ultra-sensitive detection and fast removal of iron(III) ions. Anal. Methods 2022, 14, 3727. [Google Scholar] [CrossRef]
  102. Zhou, M.J.; Guo, J.J.; Yang, C.X. Ratiometric fluorescence sensor for Fe3+ ions detection based on quantum dot-doped hydrogel optical fiber. Sens. Actuators B Chem. 2018, 264, 52–58. [Google Scholar] [CrossRef]
  103. Kuang, Y.S.; Wang, X.; Tian, X.K.; Yang, C.; Li, Y.; Nie, Y.L. Silica-embedded CdTe quantum dots functionalized with rhodamine derivative for instant visual detection of ferric ions in aqueous media. J. Photochem. Photobiol. A Chem. 2019, 372, 140–146. [Google Scholar] [CrossRef]
  104. Srivastava, R.R.; Singh, V.K.; Srivastava, A. Facile synthesis of highly fluorescent water-soluble SnS2 QDs for effective detection of Fe3+ and unveiling its fluorescence quenching mechanism. Opt. Mater. 2020, 109, 110337. [Google Scholar] [CrossRef]
  105. Wu, D.M.; Qu, C.J.; Wang, J.Z.; Yang, R.; Qu, L.B. Highly sensitive and selective fluorescence sensing and imaging of Fe3+ based on a novel nitrogen-doped graphene quantum dots. Luminescence 2021, 36, 1592–1599. [Google Scholar] [CrossRef]
  106. Wang, Z.H.; Chen, D.; Gu, B.L.; Gao, B.; Liu, Z.D.; Yang, Y.S.; Guo, Q.L.; Zheng, X.H.; Wang, G. Yellow emissive nitrogen-doped graphene quantum dots as a label-free fluorescent probe for Fe3+ sensing and bioimaging. Diam. Relat. Mater. 2020, 104, 107749. [Google Scholar] [CrossRef]
  107. Lou, Y.; Ji, J.Y.; Qin, A.M.; Liao, L.; Li, Z.Y.; Chen, S.P.; Zhang, K.Y.; Ou, J. Cane molasses graphene quantum dots passivated by PEG functionalization for detection of metal ions. ACS Omega 2020, 5, 6763–6772. [Google Scholar] [CrossRef] [Green Version]
  108. Zhou, H.; Ou, M.G.; Sun, D.H.; Yang, C.L. Facile preparation of highly fluorescent nitrogen-doped graphene quantum dots for sensitive Fe3+ detection. Opt. Laser Technol. 2022, 156, 108542. [Google Scholar] [CrossRef]
  109. Wu, S.W.; Zhou, C.; Ma, C.M.; Yin, Y.Z.; Sun, C.Y. Carbon quantum dots-based fluorescent hydrogel hybrid platform for sensitive detection of iron ions. J. Chem. 2022, 2022, 3737646. [Google Scholar] [CrossRef]
  110. Huang, M.Q.; Tong, C.L. A dual-emission ratiometric fluorescence probe for highly selective and simultaneous detection of tetracycline and ferric ions in environmental water samples based on a boron-doped carbon quantum dot/CdTe-Eu3+ composite. Environ. Sci. Nano 2022, 9, 1712–1723. [Google Scholar] [CrossRef]
  111. Sohal, N.; Maity, B.; Basu, S. Transformation of bulk MnO2 to fluorescent quantum dots for selective and sensitive detection of ferric ions and ascorbic acid by turn-off-on strategy. J. Photochem. Photobiol. A Chem. 2023, 434, 114280. [Google Scholar] [CrossRef]
  112. Lv, S.H.; Zhang, S.S.; Zuo, J.J.; Liang, S.; Liu, L.P.; Wei, D.Q. The efficient detection of Fe3+ by sulfonamidated lignin composite carbon quantum dots. Polym. Eng. Sci. 2023, 2023, 26295. [Google Scholar] [CrossRef]
  113. Lai, C.M.; Lin, S.M.; Xiong, L.; Wu, Y.X.; Liu, C.; Jin, Y.Q. High quantum yield nitrogen-doped carbon quantum dots: Green synthesis and application as “on-off” fluorescent sensors for specific Fe3+ ions detection and cell imaging. Diam. Relat. Mater. 2023, 133, 109702. [Google Scholar] [CrossRef]
  114. Singh, S.; Kansal, S.K. Dual Fluorometric detection of Fe3+ and Hg2+ ions in an aqueous medium using carbon quantum dots as a “turn-off” fluorescence sensor. J. Fluoresc. 2022, 32, 1143–1154. [Google Scholar] [CrossRef] [PubMed]
  115. Kang, K.M.; Liu, B.Y.; Yue, G.; Ren, H.W.; Zheng, K.Y.; Wang, L.M.; Wang, Z.Q. Preparation of carbon quantum dots from ionic liquid modified biomass for the detection of Fe3+ and Pd2+ in environmental water. Ecotoxicol. Environ. Saf. 2023, 255, 114795. [Google Scholar] [CrossRef] [PubMed]
  116. Kang, S.; Han, H.; Lee, K.; Kim, K.M. Ultrasensitive detection of Fe3+ ions using functionalized graphene quantum dots fabricated by a one-step pulsed laser ablation process. ACS Omega 2022, 7, 2074–2081. [Google Scholar] [CrossRef] [PubMed]
  117. Singh, H.; Bamrah, A.; Bhardwaj, S.K.; Deep, A.; Khatri, M.; Kim, K.H.; Bhardwaj, N. Nanomaterial-based fluorescent sensors for the detection of lead ions. J. Hazard. Mater. 2021, 407, 124379. [Google Scholar] [CrossRef]
  118. Shi, H.L.; Zhao, Q.; Zhou, C.H.; Jia, N.Q. Nitrogen and sulfur co-doped carbon quantum dots as fluorescence sensor for detection of lead ion. Chin. J. Anal. Chem. 2022, 50, 63–68. [Google Scholar] [CrossRef]
  119. Hadidi, M.; Ahour, F.; Keshipour, S. Electrochemical determination of trace amounts of lead ions using D-penicillamine-functionalized graphene quantum dot-modified glassy carbon electrode. J. Iran. Chem. Soc. 2021, 19, 1179–1189. [Google Scholar] [CrossRef]
  120. Wu, H.M.; Liang, J.G.; Han, H.Y. A novel method for the determination of Pb2+ based on the quenching of the fluorescence of CdTe quantum dots. Microchim. Acta 2008, 161, 81–86. [Google Scholar] [CrossRef]
  121. Zhong, W.; Zhang, C.; Gao, Q.; Li, H. Highly sensitive detection of lead(II) ion using multicolor CdTe quantum dots. Microchim. Acta 2012, 176, 101–107. [Google Scholar] [CrossRef]
  122. Mabrouk, S.; Rinnert, H.; Balan, L.; Jasniewski, J.; Medjahdi, G.; Ben Chaabane, R.; Schneider, R. Aqueous synthesis of core/shell/shell ZnSeS/Cu:ZnS/ZnS quantum dots and their use as a probe for the selective photoluminescent detection of Pb2+ in water. J. Photochem. Photobiol. A 2022, 431, 114050. [Google Scholar] [CrossRef]
  123. Cheng, S.S.; Hou, D.Y.; Li, C.; Liu, S.; Zhang, C.L.; Kong, Q.Q.; Ye, M.Q.; Wu, S.; Xian, Y.Z. Dual-ligand functionalized Ag2S quantum dots for turn-on detection of lead(II) ions in mineral samples based on aggregation-induced enhanced emission. ChemistrySelect 2021, 6, 4063–4066. [Google Scholar] [CrossRef]
  124. Bamrah, A.; Singh, H.; Singh, S.; Bhardwaj, S.K.; Khatri, M.; Deep, A.; Bhardwaj, N. Surface-functionalized fluorescent carbon dots (CDs) for dual-mode detection of lead ions. Chem. Pap. 2022, 76, 6193–6203. [Google Scholar] [CrossRef]
  125. Babu, A.M.; Bijoy, G.; Keerthana, P.; Varghese, A. Fluorescent detection of Pb2+ pollutant in water samples with the help of Delonix regia leaf-derived CQDs. Synth. Met. 2023, 291, 117211. [Google Scholar] [CrossRef]
  126. Qi, Y.X.; Zhang, M.; Fu, Q.Q.; Liu, R.; Shi, G.Y. Highly sensitive and selective fluorescent detection of cerebral lead(II) based on graphene quantum dot conjugates. Chem. Commun. 2013, 49, 10599–10601. [Google Scholar] [CrossRef] [PubMed]
  127. Qian, Z.S.; Shan, X.Y.; Chai, L.J.; Chen, J.R.; Peng, H. A fluorescent nanosensor based on graphene quantum dots-aptamer probe and graphene oxide platform for detection of lead (II) ion. Biosens. Bioelectron. 2015, 68, 225–231. [Google Scholar] [CrossRef]
  128. Sebastian, D.; Kala, R.; Parvathy, K.P.N.; Savitha, D.P. A fluorescent probe based on visible light-emitting functionalized graphene quantum dots for the sensitive and selective detection of Pb (II) ions. J. Mater. Sci. 2021, 56, 18126–18146. [Google Scholar] [CrossRef]
  129. Sharma, P.; Mehata, M.S. Rapid sensing of lead metal ions in an aqueous medium by MoS2 quantum dots fluorescence turn-off. Mater. Res. Bull. 2020, 131, 110978. [Google Scholar] [CrossRef]
  130. Cai, Z.X.; Shi, B.Q.; Zhao, L.; Ma, M.H. Ultrasensitive and rapid lead sensing in water based on environmental friendly and high luminescent l-glutathione-capped-ZnSe quantum dots. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2012, 97, 909–914. [Google Scholar] [CrossRef]
  131. Zhou, J.R.; Li, B.W.; Qi, A.J.; Shi, Y.J.; Qi, J.; Xu, H.Z.; Chen, L.X. ZnSe quantum dot based ion imprinting technology for fluorescence detecting cadmium and lead ions on a three-dimensional rotary paper-based microfluidic chip. Sens. Actuators B Chem. 2020, 305, 127462. [Google Scholar] [CrossRef]
  132. Sefid-Sefidehkhan, Y.; Jouyban, A.; Rahimpour, E. Fluorescence turn-off sensing of lead and gentamicin based on phosphorus and chlorine co-doped carbon dots. Microchem. J. 2022, 181, 107753. [Google Scholar] [CrossRef]
  133. Kumar, A.; Chowdhuri, A.R.; Laha, D.; Mahto, T.K.; Karmakar, P.; Sahu, S.K. Green synthesis of carbon dots from Ocimum sanctum for effective fluorescent sensing of Pb2+ ions and live cell imaging. Sens. Actuators B Chem. 2017, 242, 679–686. [Google Scholar] [CrossRef]
  134. Tabaraki, R.; Abdi, O. Fluorescent sensing of Pb2+ by microwave-assisted synthesized N-doped carbon dots: Application of response surface methodology and Doehlert design. J. Iran. Chem. Soc. 2020, 17, 839–846. [Google Scholar] [CrossRef]
  135. Bandi, R.; Dadigala, R.; Gangapuram, B.R.; Guttena, V. Green synthesis of highly fluorescent nitrogen-Doped carbon dots from Lantana camara berries for effective detection of lead(II) and bioimaging. J. Photochem. Photobiol. B Biol. 2018, 178, 330–338. [Google Scholar] [CrossRef]
  136. Sebastian, D.; Soman, S.; Kala, R. A nanohybrid system based on covalently functionalized graphene quantum dots with dithienopyrrole derivative for the sensitive and selective fluorometric detection of Pb2+ ions. Luminescence 2021, 36, 1743–1750. [Google Scholar] [CrossRef]
  137. Niu, X.F.; Zhong, Y.B.; Chen, R.; Wang, F.; Liu, Y.J.; Luo, D. A “turn-on” fluorescence sensor for Pb2+ detection based on graphene quantum dots and gold nanoparticles. Sens. Actuators B Chem. 2018, 255, 1577–1581. [Google Scholar] [CrossRef]
  138. Bian, S.Y.; Shen, C.; Hua, H.; Zhou, L.; Zhu, H.L.; Xi, F.N.; Liu, J.Y.; Dong, X.P. One-pot synthesis of sulfur-doped graphene quantum dots as a novel fluorescent probe for highly selective and sensitive detection of lead(II). RSC Adv. 2016, 6, 69977–69983. [Google Scholar] [CrossRef]
  139. Wang, Z.Z.; Yang, Y.; Zou, T.; Xing, X.X.; Zhao, R.J.; Wang, Y.D. Novel method for the qualitative identification of chromium ions (III) using L-aspartic acid stabilized CdS quantum dots. J. Phys. Chem. Solids 2020, 136, 109160. [Google Scholar] [CrossRef]
  140. Liu, X.; Yang, Y.; Li, Q.; Wang, Z.Z.; Xing, X.X.; Wang, Y.D. Portably colorimetric paper sensor based on ZnS quantum dots for semi-quantitative detection of Co2+ through the measurement of grey level. Sens. Actuators B Chem. 2018, 260, 1068–1075. [Google Scholar] [CrossRef]
  141. Wang, Z.Z.; Xing, X.X.; Yang, Y.; Zhao, R.J.; Zou, T.; Wang, Z.D.; Wang, Y.D. One-step hydrothermal synthesis of thioglycolic acid capped CdS quantum dots as fuorescence determination of cobalt ion. Sci. Rep. 2018, 8, 8953. [Google Scholar] [CrossRef] [Green Version]
  142. Tong, Z.; Xing, X.X.; Yang, Y.; Hong, P.; Wang, Z.D.; Zhao, R.J.; Zhang, X.; Peng, S.J.; Wang, Y.D. Fluorescent ZnO quantum dots synthesized with urea for the selective detection of Cr6+ ion in water with a wide range of concentrations. Methods Appl. Fluoresc. 2019, 7, 035007. [Google Scholar] [CrossRef] [PubMed]
  143. Yang, Y.; Zou, T.; Zhao, R.J.; Kong, Y.L.; Su, L.F.; Ma, D.; Xiao, X.C.; Wang, Y.D. Fluorescence ‘turn-on’ probe for Al3+ detection in water based on ZnS/ZnO quantum dots with excellent selectivity and stability. Nanotechnology 2021, 32, 375001. [Google Scholar] [CrossRef] [PubMed]
  144. Wang, Z.Z.; Xiao, X.C.; Yang, Y.; Zou, T.; Xing, X.X.; Zhao, R.J.; Wang, Z.D.; Wang, Y.D. L-aspartic acid capped CdS quantum dots as a high performance fluorescence assay for sliver ions (I) detection. Nanomaterials 2019, 9, 1165. [Google Scholar] [CrossRef] [Green Version]
  145. Xia, Y.S.; Cao, C.; Zhu, C.Q. Two distinct photoluminescence responses of CdTe quantum dots to Ag (I). J. Lumin. 2008, 128, 166–172. [Google Scholar] [CrossRef]
  146. Chen, B.; Liu, J.J.; Yang, T.; Chen, L.; Hou, J.; Feng, C.H.; Huang, C.Z. Development of a portable device for Ag+ sensing using CdTe QDs as fluorescence probe via an electron transfer process. Talanta 2018, 191, 357–363. [Google Scholar] [CrossRef]
  147. Liu, Y.F.; Gao, Z.J.; Shao, X.W.; Yang, J.P.; Tang, X.S.; Wang, J.; Chen, W.W.; Lin, H.; Deng, M.; Zhu, T. A turn-on ratiometric fluorescent nanoprobe based on AgInZnS and nitrogen-doped graphene quantum dots for Cd2+ detection in lake water. J. Mater. Sci. 2022, 57, 17336–17346. [Google Scholar] [CrossRef]
  148. Xu, H.; Miao, R.; Fang, Z.; Zhong, X.H. Quantum dot-based “turn-on” fluorescent probe for detection of zinc and cadmium ions in aqueous media. Anal. Chim. Acta 2011, 687, 82–88. [Google Scholar] [CrossRef]
  149. Yan, Z.H.; Yao, W.; Mai, K.; Huang, J.Q.; Wan, Y.T.; Huang, L.; Cai, B.; Liu, Y. A highly selective and sensitive “on–off” fluorescent probe for detecting cadmium ions and L-cysteine based on nitrogen and boron co-doped carbon quantum dots. RSC Adv. 2022, 12, 8202–8210. [Google Scholar] [CrossRef]
  150. Zhu, L.L.; Shen, D.K.; Luo, K.H. Triple-emission nitrogen and boron co-doped carbon quantum dots from lignin: Highly fluorescent sensing platform for detection of hexavalent chromium ions. J. Colloid Interface Sci. 2022, 617, 557–567. [Google Scholar] [CrossRef]
  151. Hu, Q.; Li, T.; Gao, L.; Gong, X.J.; Rao, S.Q.; Fang, W.M.; Gu, R.X.; Yang, Z.Q. Ultrafast and energy-saving synthesis of nitrogen and chlorine co-doped carbon nanodots via neutralization heat for selective detection of Cr(VI) in aqueous phase. Sensors 2018, 18, 3416. [Google Scholar] [CrossRef] [Green Version]
  152. Wang, C.L.; Xu, J.; Li, H.Z.; Zhao, W.L. Tunable multicolour S/N co-doped carbon quantum dots synthesized from waste foam and application to detection of Cr3+ ions. Luminscence 2020, 35, 1373–1383. [Google Scholar] [CrossRef]
  153. Mohamed, N.B.H.; Ben Brahim, N.; Mrad, R.; Haouari, M.; Ben Chaabane, R.; Negrerie, M. Use of MPA-capped CdS quantum dots for sensitive detection and quantification of Co2+ ions in aqueous solution. Anal. Chim. Acta 2018, 1028, 50–58. [Google Scholar] [CrossRef] [PubMed]
  154. Zikalala, N.; Parani, S.; Oluwafemi, O.S. Facile aqueous synthesis of ZnInS quantum dots and its application for selective detection of Co2+ ions. Nanotechnology 2021, 32, 295503. [Google Scholar] [CrossRef] [PubMed]
  155. Song, Y.X.; Xia, X.F.; Xiao, Z.W.; Zhao, Y.; Yan, M.J.; Li, J.Y.; Li, H.Q.; Liu, X.T. Synthesis of N, S co-doped carbon dots for fluorescence turn-on detection of Fe2+ and Al3+ in a wide pH range. J. Mol. Liq. 2022, 368, 120663. [Google Scholar] [CrossRef]
  156. Peng, C.F.; Zhang, Y.Y.; Qian, Z.J.; Xie, Z.J. Fluorescence sensor based on glutathione capped CdTe QDs for detection of Cr3+ ions in vitamins. Food Sci. Hum. Wellness 2018, 7, 71–76. [Google Scholar] [CrossRef]
  157. Oliveira, J.J.P.; Carneiro, S.V.; Cruz, A.A.C.; Fechine, L.M.U.D.; Michea, S.; Antunes, R.A.; Neto, M.L.A.; Moura, T.A.; Cesar, C.L.; Carvalho, H.F. Determination of Co2+ ions in blood samples: A multi-way sensing based on NH2-rich carbon quantum dots. Dye Pigment. 2023, 215, 11253. [Google Scholar] [CrossRef]
  158. Jing, N.; Tian, M.; Wang, Y.T.; Zhang, Y. Nitrogen-doped carbon dots synthesized from acrylic acid and ethylenediamine for simple and selective determination of cobalt ions in aqueous media. J. Lumin. 2019, 206, 169–175. [Google Scholar] [CrossRef]
  159. Wang, C.J.; Sun, Q.; Li, C.X.; Tang, D.B.; Shi, H.X.; Liu, E.Z.; Guo, P.Q.; Xue, W.M.; Fan, J. Biocompatible double emission boron nitrogen co-doped carbon quantum dots for selective and sensitive detection of Al3+ and Fe2+. Mater. Res. Bull. 2022, 155, 111970. [Google Scholar] [CrossRef]
  160. Wu, Q.A.; Zhou, M.; Shi, J.; Li, Q.J.; Yang, M.Y.; Zhang, Z.X. Synthesis of water-soluble Ag2S quantum dots with fluorescence in the second near-infrared window for turn-on detection of Zn(II) and Cd(II). Anal. Chem. 2017, 89, 6616–6623. [Google Scholar] [CrossRef]
  161. Dastidar, D.G.; Mukherjee, P.; Ghosh, D.; Banerjee, D. Carbon quantum dots prepared from onion extract as fluorescence turn-on probes for selective estimation of Zn2+ in blood plasma. Colloids Surf. A Physicochem. Eng. Asp. 2021, 611, 125781. [Google Scholar] [CrossRef]
  162. Zhou, Z.Q.; Yang, L.Y.; Liao, Y.P.; Wu, H.Y.; Zhou, X.H.; Huang, S.; Liu, Y.; Xiao, Q. A Mn:ZnSe quantum dot-based turn-on fluorescent sensor for the highly selective and sensitive detection of Cd2+. Anal. Methods 2020, 12, 552–556. [Google Scholar] [CrossRef]
  163. Zhang, X.; Liao, X.F.; Hou, Y.J.; Jia, B.Y.; Fu, L.Z.; Jia, M.X.; Zhou, L.D.; Lu, J.H.; Kong, W.J. Recent advances in synthesis and modification of carbon dots for optical sensing of pesticides. J. Hazard. Mater. 2022, 422, 126881. [Google Scholar] [CrossRef] [PubMed]
  164. Mahajan, M.R.; Patil, P.O. Design of zero-dimensional graphene quantum dots based nanostructures for the detection of organophosphorus pesticides in food and water: A review. Inorg. Chem. Commun. 2022, 144, 109883. [Google Scholar] [CrossRef]
  165. Maluleke, R.; Oluwafemi, O.S. Synthetic approaches, modification strategies and the application of quantum dots in the sensing of priority pollutants. Appl. Sci. 2021, 11, 11580. [Google Scholar] [CrossRef]
  166. Yang, Y.X.; Huo, D.Q.; Wu, H.X.; Wang, X.F.; Yang, J.S.; Bian, M.H.; Ma, Y.; Hou, C.J. N, P-doped carbon quantum dots as a fluorescent sensing platform for carbendazim detection based on fluorescence resonance energy transfer. Sens. Actuators B Chem. 2018, 274, 296–303. [Google Scholar] [CrossRef]
  167. Yang, Y.; Xing, X.X.; Zou, T.; Wang, Z.D.; Zhao, R.J.; Hong, P.; Peng, S.J.; Zhang, X.; Wang, Y.D. A novel and sensitive ratiometric fluorescence assay for carbendazim based on N-doped carbon quantum dots and gold nanocluster nanohybrid. J. Hazard. Mater. 2020, 386, 121958. [Google Scholar] [CrossRef]
  168. Yang, Y.; Wei, Q.Y.; Zou, T.; Kong, Y.L.; Su, L.F.; Ma, D.; Wang, Y.D. Dual-emission ratiometric fluorescent detection of dinotefuran based on sulfur-doped carbon quantum dots and copper nanocluster hybrid. Sens. Actuators B Chem. 2020, 321, 128534. [Google Scholar] [CrossRef]
  169. Bera, M.K.; Mohapatra, S. Ultrasensitive detection of glyphosate through effective photoelectron transfer between CdTe and chitosan derived carbon dot. Colloids Surf. A Physicochem. Eng. Asp. 2020, 596, 124710. [Google Scholar] [CrossRef]
  170. Cheng, C.H.; Peng, F.; Hou, W.L.; Wang, X.P. Selective and sensitive detection of malathion pesticide with CdSe quantum dots as ligand-exchange probes. Environ. Eng. Sci. 2021, 39, 453–459. [Google Scholar] [CrossRef]
  171. Yang, L.M.; Zhang, X.H.; Jiang, L. Determination of organophosphorus pesticides in fortified tomatoes by fluorescence quenching of cadmium selenium-zinc sulfide quantum dots. Anal. Lett. 2019, 52, 729–744. [Google Scholar] [CrossRef]
  172. Nair, R.V.; Chandran, P.R.; Mohamed, A.P.; Pillai, S. Sulphur-doped graphene quantum dot based fluorescent turn-on aptasensor for selective and ultrasensitive detection of omethoate. Anal. Chim. Acta 2021, 1181, 338893. [Google Scholar] [CrossRef] [PubMed]
  173. Nair, R.V.; Thomas, R.T.; Mohamed, A.P.; Pillai, S. Fluorescent turn-off sensor based on sulphur-doped graphene quantum dots in colloidal and film forms for the ultrasensitive detection of carbamate pesticides. Microchem. J. 2020, 157, 104971. [Google Scholar] [CrossRef]
  174. Chen, L.; Lu, J.; Luo, M.; Yu, H.; Chen, X.J.; Deng, J.G.; Hou, X.T.; Hao, E.W.; Wei, J.C.; Li, P. A ratiometric fluorescent sensing system for the selective and ultrasensitive detection of pesticide residues via the synergetic effects of copper nanoclusters and carbon quantum dots. Food Chem. 2022, 379, 132139. [Google Scholar] [CrossRef]
  175. Shahdost-fard, F.; Fahimi-Kashani, N.; Hormozi-nezhad, M.R. A ratiometric fluorescence nanoprobe using CdTe QDs for fast detection of carbaryl insecticide in apple. Talanta 2020, 221, 121467. [Google Scholar] [CrossRef] [PubMed]
  176. Li, H.Q.; Deng, R.; Tavakoli, H.M.; Li, X.C.; Li, X.J. Ultrasensitive detection of acephate based on carbon quantum dot-mediated fluorescence inner filter effects. Analyst 2022, 147, 5462–5469. [Google Scholar] [CrossRef]
  177. Chandra, S.; Bano, D.; Sahoo, K.; Kumar, D.; Kumar, V.; Yadav, P.K.; Hasan, S.H. Synthesis of fluorescent carbon quantum dots from Jatropha fruits and their application in fluorometric sensor for the detection of chlorpyrifos. Microchem. J. 2022, 172, 106953. [Google Scholar] [CrossRef]
  178. Lv, S.C.; Liu, D.H.; Liu, X.K.; Zhou, Z.Q.; Wang, P. Aminophenol functionalized carbon quantum dots as fluorescent sensor for nitroalkenes. Microchem. J. 2023, 189, 108569. [Google Scholar] [CrossRef]
  179. Luo, J.M.; Li, S.H.; Pang, C.H.; Wang, M.Y.; Ma, X.H.; Zhang, C.H. Highly selective fluorescence probe for imidacloprid measurement based on fluorescence resonance energy transfer. Microchem. J. 2022, 175, 107172. [Google Scholar] [CrossRef]
Figure 1. (a) Fluorescence spectra of NALC-CdS QDs with different concentrations of Cu2+ in aqueous solution, (b) the relationship between the relative fluorescence intensity (F0/F) and the concentration of Cu2+, and (c) schematic of the quenching mechanism of NALC-CdS QDs by Cu2+. Reprinted with permission from ref. [61]; Copyright IOP Publishing. (d) Fluorescence spectra of the NH2-ZnO QDs with different concentrations of Cu2+ in aqueous solution; (e,f) the linear relation curves of F0/F and the concentration of Cu2+: (e) 2–20 nM; (f) 1–100 μM. (g) Schematic of quenching mechanism of NH2-ZnO QDs by Cu2+. Reprinted with permission from ref. [63]; Copyright 2020 Elsevier.
Figure 1. (a) Fluorescence spectra of NALC-CdS QDs with different concentrations of Cu2+ in aqueous solution, (b) the relationship between the relative fluorescence intensity (F0/F) and the concentration of Cu2+, and (c) schematic of the quenching mechanism of NALC-CdS QDs by Cu2+. Reprinted with permission from ref. [61]; Copyright IOP Publishing. (d) Fluorescence spectra of the NH2-ZnO QDs with different concentrations of Cu2+ in aqueous solution; (e,f) the linear relation curves of F0/F and the concentration of Cu2+: (e) 2–20 nM; (f) 1–100 μM. (g) Schematic of quenching mechanism of NH2-ZnO QDs by Cu2+. Reprinted with permission from ref. [63]; Copyright 2020 Elsevier.
Chemosensors 11 00405 g001
Figure 2. (A) UV-vis spectra (a, b) and FL spectra (excitation spectra, c, d; emission spectra, e, f) of BPEI-CQD solution in the absence (a, c, e) and presence (b, d, f) of 150 μM Cu2+ ions. The inset shows photos of BPEI-CQD solutions in the absence (left) and presence (right) of Cu2+ under UV light of 365 nm. (B) FL response of BPEI-CQDs upon addition of various concentrations of copper ions. Inset: Linear relationship between F0/F of BPEI-CQDs and the concentration of Cu2+. (C) Schematic diagram of the fluorescence quenching of the BPEI-CQDs by Cu2+. Reprinted with permission from ref. [66]; Copyright IOP Publishing.
Figure 2. (A) UV-vis spectra (a, b) and FL spectra (excitation spectra, c, d; emission spectra, e, f) of BPEI-CQD solution in the absence (a, c, e) and presence (b, d, f) of 150 μM Cu2+ ions. The inset shows photos of BPEI-CQD solutions in the absence (left) and presence (right) of Cu2+ under UV light of 365 nm. (B) FL response of BPEI-CQDs upon addition of various concentrations of copper ions. Inset: Linear relationship between F0/F of BPEI-CQDs and the concentration of Cu2+. (C) Schematic diagram of the fluorescence quenching of the BPEI-CQDs by Cu2+. Reprinted with permission from ref. [66]; Copyright IOP Publishing.
Chemosensors 11 00405 g002
Figure 3. The top views of density functional theory (DFT) simulation mode for (a) N, S-GQDs and (b) M cations (M = Fe3+, Hg2+) adsorbed on the surface of N, S-GQDs. (c) Schematic diagram of N, S-GQDs as a fluorescence probe. Reprinted with permission from ref. [90].
Figure 3. The top views of density functional theory (DFT) simulation mode for (a) N, S-GQDs and (b) M cations (M = Fe3+, Hg2+) adsorbed on the surface of N, S-GQDs. (c) Schematic diagram of N, S-GQDs as a fluorescence probe. Reprinted with permission from ref. [90].
Chemosensors 11 00405 g003
Figure 4. (a) PL intensity of QDs in the presence of Fe3+ ions. (b) The linear relationship between the fluorescence intensity and Fe3+. (c) Schematic diagram for static and dynamic quenching of f-SnS2 QDs by Fe3+ ions. Reprinted with permission from ref. [104]. Copyright 2020 Elsevier.
Figure 4. (a) PL intensity of QDs in the presence of Fe3+ ions. (b) The linear relationship between the fluorescence intensity and Fe3+. (c) Schematic diagram for static and dynamic quenching of f-SnS2 QDs by Fe3+ ions. Reprinted with permission from ref. [104]. Copyright 2020 Elsevier.
Chemosensors 11 00405 g004
Figure 5. (a) Interference test of QD solution for various metal ions. (b) The fluorescence spectra of QDs upon different concentrations of Pb2+. (c) Linear relationship between the concentration of Pb2+ and FL intensity. (d) Mechanism of florescence quenching. Reprinted with permission from ref. [125]. Copyright 2022 Elsevier.
Figure 5. (a) Interference test of QD solution for various metal ions. (b) The fluorescence spectra of QDs upon different concentrations of Pb2+. (c) Linear relationship between the concentration of Pb2+ and FL intensity. (d) Mechanism of florescence quenching. Reprinted with permission from ref. [125]. Copyright 2022 Elsevier.
Chemosensors 11 00405 g005
Table 1. A summary of QD-based sensors for the detection of Cu2+.
Table 1. A summary of QD-based sensors for the detection of Cu2+.
QD-Based SensorsAnalyteLinearityLOD (nM)ApplicationsMechanismRef.
AgInS2 QDsCu2+0–340 μM27.3Drinking/natural waterFRET[69]
Ag2S QDsCu2+0.025–10 μM27.6Drinking waterPET[70]
CdSe/ZnS QDsCu2+, Hg2+0.02–0.7 μM6.94~PET[71]
MPA-InP/ZnS QDsCu2+0–1000 nM0.22River/tap/purified/mineral/drinking water, beveragesSQE[53]
TGA-CdTe QDsCu2+, Ag+40–60 nM35UrineSQE[72]
CdSe/ZnS QDs@PESMCu2+0.15–15 μM67Tap water, baijiu, orange juice, beer, milkSQE[73]
ZnSe/ZnS-MPA QDsCu2+0.049–59 μM170~SQE[74]
P, N-CQDsCu2+4–400 nM1.5WaterIFE[75]
N-CQDsCu2+, GSH, pH0.1–40 μM90Cellular imaging, natural waterSQE[76]
N, S-CQDsCu2+, EDT5–125 μM~Natural waterIFE[77]
P-CQDsCu2+0–260 μM32Tap waterIFE[78]
N–GQDsCu2+0–10 μM57~SQE[79]
GQDsCu2+1–40 μM440~ACQ[80]
GQDsCu2+0–15 μM226River waterSQE[81]
GQDsCu2+, MPG0.01–0.5 μM2.5~PET[82]
N–GQDsCu2+, Co2+0.2–1 μM~~~[83]
a-GQDs Cu2+0–100 μM5.6~PET[84]
Table 2. A summary of QD-based sensors for the detection of Hg2+.
Table 2. A summary of QD-based sensors for the detection of Hg2+.
QDs-Based SensorsAnalyte(s)LinearityLOD
(nM)
ApplicationsMechanismsRef.
CdTe@3-MIBAHg2+1.5–100 nM1.5 ± 0.5~ACQ, DQE[85]
GSH@CdS QDsHg2+10–1000 nM/
1–20 µM
0.54Tap/waterACQ, DQE[86]
CdSe/ZnS QDsHg2+, Cu2+0.1–1.4 μM20.58~ACQ[71]
PEI-Ag2SHg2+5–625 nM5Lake waterACQ[87]
CdSe/ZnS/OAHg2+0.3–6.1 µM74.8~ACQ[88]
CdSe/ZnS/GSHHg2+0.3–6.1 µM54.8~ACQ[88]
N-GQDsHg2+1–1000 nM0.45~DQE[89]
N, S-GQDsHg2+, Fe3+1–30 nM0.27Drinking waterDQE[90]
RhB-GQDsHg2+0–1000 nM0.16Drinking waterDQE[91]
GQDsHg2+0–380 µM82~DQE[92]
8-HQ-GQDsHg2+0–200 nM2.4Tap/lake/underground/dam waterDQE[93]
CQDs–Ag NPsHg2+0.5–500 nM0.1Lake water, wastewater, green/Galin/Lahijan teaFRET[94]
CQDsHg2+50–500 µM6.2~DQE[95]
FA-CDsHg2+10–1000 nM1.29Lake/tap/sea water, cellular imagingDQE[96]
N-CQDsHg2+, captopril10–100 nM1.43River/pond waterDQE[97]
N/Ag-CQDsHg2+, captopril10–100 nM0.93River/pond waterDQE[97]
N/Ce-CQDsHg2+, captopril10–100 nM1.39River/pond waterDQE[97]
NS-CDsHg2+, GSH0–100 µM50Tap waterSQE[98]
Table 3. A summary of QD-based sensors for the detection of Fe3+.
Table 3. A summary of QD-based sensors for the detection of Fe3+.
QDs-Based SensorsAnalyte(s)LinearityLOD (nM)ApplicationsMechanismRef.
NAC-CdTe QDs&
TGA-CdTe QDs
Fe3+0–3.5 µM14Tap waterIFE[102]
CdTe@SiO2 QDsFe3+0–3.25 µM26.5Tap/river/lake waterFRET[103]
f-SnS2 QDsFe3+0–68 µM840~SQE, DQE[104]
PEG-GQDsFe3+8–24 μM5770Healthy human serumFRET, ACQ[107]
N-GQDsFe3+0–100 μM740~FRET[108]
CQDsHGFe3+0–150 μM240Tap/river waterSQE[109]
N-GQDsFe3+0.02–12 μM1.43Cellular imaging, lake waterSQE[105]
N, S-GQDsFe3+, Hg2+1–90 nM/
0.1–30 μM
2.88Drinking waterIFE[90]
N-GQDsFe3+0–80 μM63Tap/river water, cellular imagingSQE[106]
B-CQDs/CdTe-Eu3+Fe3+, tetracycline0.1–15 μM53River/lake waterSQE[110]
CA-MnO2 QDsFe3+, ascorbic acid0–25 µM43WaterDQE[111]
PSA-L-CQDsFe3+0–500 µM29.5Natural water organic wastewater~[112]
N-CQDsFe3+0–110 μM177Tap water,
cellular imaging
SQE[113]
CQDsFe3+, Hg2+0–50 μM406~SQE[114]
ILB-CQDsFe3+, Pb2+0–40 μM0.001Tap/river waterPET[115]
MoS2Fe3+0–50 μM1000Tap/pond/mineral waterSQE[100]
F-GQDsFe3+0.5–50 μM500~SQE[116]
Table 4. A summary of QD-based sensors for the detection of Pb2+.
Table 4. A summary of QD-based sensors for the detection of Pb2+.
QDs-Based SensorsAnalyte(s)LinearityLOD (nM)ApplicationsMechanismRef.
TGA-CdTe QDsPb2+1.96–330 nM4.7Spinach, citrusFluorescence
quenching
[121]
CdTe QDsPb2+2–100 µM270~SQE[120]
ZnSeS/Cu:ZnS/ZnS QDsPb2+0.04–6 µM21Mineral/lake waterDQE[122]
3-MPA/GSH-Ag2S QDsPb2+0.05–20 µM15.5Tap/mineral waterAIEE[123]
MoS2Pb2+0.033–8 mM50,000~Fluorescence
quenching
[129]
GSH-ZnSe QDsPb2+10–800 nM0.71WaterFRET[130]
ZnSe QDsPb2+, Cd2+4.8–289 nM1.6Lake/sea waterFRET[131]
N, S-CQDsPb2+0.2–12 µM
40–200 µM
97/5624River waterFluorescence
quenching
[118]
CQDsPb2+10–180 µM3.3 River/lake/tap water FRET [115]
MPA@CDsPb2+0.05–6 µM0.051Tap waterAIEE[114]
P, Cl-CQDsPb2+, gentamicin0.0483–14.49 µM14.49River/tap/mineral water,Fluorescence
quenching
[132]
CQDsPb2+0.01–1 µM0.59Mineral/tap/pond water,
cellular imaging
Fluorescence
quenching
[133]
N-CQDsPb2+0.05–25 µM20River/mineral/tap waterSQE[134]
N-CQDsPb2+0–200 nM9.64Cellular imaging,
MCF cells
DQE[135]
DMA-GQDsPb2+0.01–1 nM0.009Cerebral spinal fluid of ratsFluorescence
enhancement
[126]
GQDs-GOPb2+0–400 nM0.6~PET[127]
DTPAN-fn-GQDsPb2+0.5–40 nM0.25River/tap/well water, paper stripsSQE, DQE[128]
DTP-PPD-fn-GQDPb2+3–30 nM1.02Tap/lake waterSQE[136]
GQDs-AuNPsPb2+50–4000 nM16.7~FRET[137]
S-GQDsPb2+0.1–140 µM30~FRET[138]
Table 5. A summary of QD-based sensors for the detection of Cd2+, Zn2+, Ag+, Co2+, Fe2+, Cr6+, and Cr3+.
Table 5. A summary of QD-based sensors for the detection of Cd2+, Zn2+, Ag+, Co2+, Fe2+, Cr6+, and Cr3+.
QDs-Based SensorsAnalyte(s)LinearityLOD (nM)ApplicationsMechanismRef.
TGA-CdTe QDsAg+5–200 nM5Lake waterDQE[146]
AgInZnS QDs-NGQDsCd2+0.5–100 µM28.6Lake waterCHEF[147]
L-Asp@CdS QDsAg+100–7000 nM39Drinking waterFRET[144]
CdTe QDsZn2+
Cd2+
1.6–35 µM
1.3–25 µM
1200
500
Tap/spring waterCHEF[148]
N, B-CQDsCd2+, L-cysteine2.5–22.5 µM450Human urineCHEF[149]
urea-ZnO QDsCr6+4–1000 µM19.53~ACQ, DQE[142]
N, B-CQDsCr6+1–100 µM54, 49, 77~DQE[150]
S, N-CQDsCr3+0–500 µM600DI/mineral/tap/river waterFluorescence
quenching
[152]
GSH-CdTe QDsCr3+0–2 µM3Vitamin tabletsSQE, DQE[156]
MPA-CdS QDsCo2+0.04–2 µM~~SQE[153]
ZnInS QDsCo2+0.1–100 μM99~ACQ[154]
CQDsCo2+0–250 μM3010Human bloodFluorescence
quenching
[157]
N-CQDsCo2+1–60 μM250Tap waterSQE[158]
B,N-CQDsAl3+
Fe2+
0–20 μM
0–25 μM
187
276
Tap water
Ferrous sulfate tablets
AIEE
ACQ
[159]
N,S-CQDsAl3+
Fe2+
0–20 μM
0–25 μM
6.99
5.5
~ FE
AIEE
[155]
ZnS/ZnO QDsAl3+0.001–8 μM/
8–100 μM
1Tap/lake waterFluorescence enhancement[143]
Ag2S QDsZn2+1–40 μM760Peptone/intracellular fluidFluorescence enhancement[160]
Ag2S QDsCd2+1–40 μM546Tap/lake waterFluorescence enhancement[160]
CQDsZn2+~6400Blood plasmaFluorescence enhancement[161]
CdTe QDsZn2+
Cd2+
1.6–35 μM
1.3–25 μM
1200
500
Tap/spring waterFluorescence enhancement[148]
Mn:ZnSe QDsCd2+0.02–60 μM18Tap waterFluorescence enhancement[162]
Table 6. A summary of QD-based sensors for the detection of pesticide residues.
Table 6. A summary of QD-based sensors for the detection of pesticide residues.
QDs-Based SensorsSelectivityLinearityLODApplicationsMechanismRef.
N, P-CQDs/
Au NPs
carbendazim0.005–0.16 µM2 nMCabbage, appleFRET[166]
N,CQDs/AuNCscarbendazim1–100 µM
150–1000 µM
0.83 µMStrawberry, appleFRET[167]
S-CQDs/CuNCsdinotefuran10–500 µM7.04 µMHoneyIFE[168]
CdTe-CQDsglyphosate0–1000 nM2 pMCucumber, capsicum, gingerPET[169]
CdSe QDsmalathion0.908–303 nM303 pMCabbage leavesFluorescence quenching[170]
CdSe/ZnS QDsorganophosphorus0.0001–160
μg/mL
96.7 
μg/mL
TomatoFluorescence quenching[171]
S-GQDsomethoate4.7–470 µM4.7 nM~Fluorescence enhancement[172]
S-GQDscarbofuran
thiram
0–225 ppb
0–800 ppb
0.45 ppb
1.6 ppb
ApplePET[173]
N-CQDsthiram
paraquat
10–500 nM
5–100 nM
7.49 nM
3.03 nM
Semen nelumbinis, coix lacryma-jobi, ginseng FRET [174]
CdTe QDscarbaryl0–14 μg/mL0.12 ng/mLAppleRatiometric fluorescence[175]
CQDsacephate0–100 ng/mL0.052 ppbPearIFE[176]
CQDschlorpyrifos0.02–0.18 μg/mL2.7 ng/mLRiver water,
apple juice
DQE[177]
AP-NSCDsnitroalkenes0.05–10 mg/L0.02 mg/L Rice, orange, cabbage, serum, urine, water FRET [178]
GQDs/CDsimidacloprid5–4000 nM0.823 nM Banana, apple, cabbage, cucumber FRET [179]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, Z.; Yao, B.; Xiao, Y.; Tian, X.; Wang, Y. Fluorescent Quantum Dots and Its Composites for Highly Sensitive Detection of Heavy Metal Ions and Pesticide Residues: A Review. Chemosensors 2023, 11, 405. https://doi.org/10.3390/chemosensors11070405

AMA Style

Wang Z, Yao B, Xiao Y, Tian X, Wang Y. Fluorescent Quantum Dots and Its Composites for Highly Sensitive Detection of Heavy Metal Ions and Pesticide Residues: A Review. Chemosensors. 2023; 11(7):405. https://doi.org/10.3390/chemosensors11070405

Chicago/Turabian Style

Wang, Zhezhe, Bo Yao, Yawei Xiao, Xu Tian, and Yude Wang. 2023. "Fluorescent Quantum Dots and Its Composites for Highly Sensitive Detection of Heavy Metal Ions and Pesticide Residues: A Review" Chemosensors 11, no. 7: 405. https://doi.org/10.3390/chemosensors11070405

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