Skip to content
BY 4.0 license Open Access Published by De Gruyter November 5, 2020

Recent advances in fluorescence probes based on carbon dots for sensing and speciation of heavy metals

  • Pingjing Li EMAIL logo and Sam F. Y. Li ORCID logo EMAIL logo
From the journal Nanophotonics

Abstract

Heavy metal (HM) pollution is a major global concern. Carbon dots (CDs) have demonstrated unique properties as sensing platforms for HMs detection. This review summarizes the progress made in recent years in fluorescence methods to determine HMs and their species using CDs. First, the strategies to synthesize and purify CDs are reviewed. The photoluminescence principles of CDs and their sensing mechanisms as HMs sensors are then summarized. The binding strategies between CDs and HMs are proposed to provide salient principles to design desirable CD-based HMs sensors. The preparation and merits of “turn-on” and ratiometric CDs for HMs detection with higher accuracy are discussed compared with commonly used “turn-off” sensors. Subsequently, the progress on detecting single HM ions, multi-HMs, and different metal species in solution, and the development of gel/solid-state sensor platforms such as paper-based devices, sensor arrays, hydrogels, polymer films, and ion-imprinted polymers are critically accessed. Furthermore, the advances in the cell, bacterial, plant, and animal bioimaging of HMs with CDs as promising bioimaging reagents are presented. Finally, the challenges and prospects of CDs as HMs sensors in future investigations are discussed.

1 Heavy metals and their species

HMs are elements with density higher than 5 g/cm3 including lead (Pb), cadmium (Cd), mercury (Hg), arsenic (As), chromium (Cr), copper (Cu), selenium (Se), nickel (Ni), silver (Ag), and zinc (Zn). Other less metallic elements include irons (Fe), aluminum (Al), cesium (Cs), cobalt (Co), manganese (Mn), molybdenum (Mo), strontium (Sr), and uranium (U) [1]. They can exist as different species (electronic or oxidation state, complex or molecular structure, inorganic, and organic molecular structure) depending upon their binding with various compounds or the existing environmental conditions. They can be presented as inorganic ions in the form of cations (e.g., Hg(II), Cd(II), Pb(II), and Cr(III)), anions (e.g., As(III), As(V), and Cr(VI)), alkylated forms (e.g., monomethylmercury, monomethylarsonic acid, and dimethylarsinic acid), and even more complex structures. HMs are classified into two groups, i.e., essential and nonessential HMs, based on their toxicity. According to the World Health Organization, Cd, Pb, Hg, and As are classified as extremely toxic, carcinogenic, and mutagenic agents [2], which are nonessential elements for plants and animals. They cannot be metabolized and biodegraded in human or animal bodies, leading to gradual accumulation. Thus, even at low concentrations, severe environmental and health risks could be induced. Nonessential HMs such as Cu, Zn, and Cr are given lower toxic ratings. They are required as micronutrients at low concentrations, while they become toxic at higher concentrations [3]. Moreover, the toxicity of certain HMs is more relevant to its specific species. For instance, cationic species of Cr(III) is an essential element, but anionic species of Cr(VI) is genotoxic and carcinogenic; arsenites [As(III)] are more toxic than arsenates [As(V)], and methylmercury (MeHg(I)) have much higher toxicity than inorganic Hg(II). HM pollution happens through natural processes and anthropogenic activities and exists in complex environmental, industrial, and biological sample matrices. Therefore, the determination of the concentration levels of HMs and their species, monitoring the pollution index, and assessing their health risks are of great importance. Table 1 gives some information on the most concerned toxic HMs.

Table 1:

Toxicity of heavy metals of environmental and health concern.

HM Toxicity information
Hg 1. Hg has extremely toxicity, non-biodegradability, and facile accumulation and may cause brain damage, kidney failure, and various cognitive and motion disorders [141].

2. The main Hg species of environmental concerned are inorganic Hg(II) and monomethylmercury (MeHg(I)), and MeHg(I) is much toxic than Hg(II);

3. The maximum permitted level of Hg(II) in drinking water is 2 μg/L as regulated by the United States Environmental Protection Agency (USEPA), 6 μg/L by the World Health Organization (WHO), and 2 μg/L by the Chinese National Standard [76], [ 142].
Cr 1. Cr has two main stable species in a solution of Cr(III) and Cr(VI). Cr(III) is essential with the recommended human intake of 50–200 μg/day [143], but excessive uptake of Cr(III) will cause harm to human health [144]; Cr(VI) is approximately 100 times more toxic than Cr(III) with carcinogenic and mutagenic effects [57], [ 143].

2. Environmental Health Hazards Assessment of California regulates the maximum permitted level for total Cr in water is 100 μg/L and gives a guide of 0.02 μg/L for Cr(VI) as a protection level; EC Directive 98/83/EC regulates the maximum amounts of total Cr and Cr(VI) in drinking water of 50 and 20 μg/L, respectively [143].
Pb 1. Pb is hazardous to the nervous system, heart, kidneys, and gastrointestinal tract and can cause cognitive impairments for children.

2. USEPA sets the maximum permissible limit of 15 μg/L for Pb (II) in drinking water [96].
As 1. The most environmental concerned As species are arsenite (As (III)) and arsenate (As(V)), and As(III) is 10 times higher As(V), 70 times higher than organic species of monomethylarsonic acid (MMA), dimethylarsinic acid (DMA) [144].

2. The pKa,1 for As(III) and As(V) is 9.22 and 2.20, respectively, which means As(III) exists as uncharged As(OH)3 while As(V) is negatively charged in a neutral medium.

3. WHO regulated a permissible limit of 10 μg/L total As.
Cd 1. Continuously exposure of even a minute amount of Cd may cause lung damage, bone fractures, liver, and kidney dysfunction and reproductive defects, and even certain cancers [95].

2. The maximum permitted level of Cd(II) in drinking water is 5 μg/L as regulated by USEPA.
  1. HM, heavy metal.

2 CDs as HMs sensor

For the detection of HMs, the development of selective, sensitive, feasible, practical, and high-throughput analysis methods is desirable. Traditional elemental specific-based approaches such as atomic absorption spectrometry, atomic emission spectrometry, inductively coupled plasma mass spectrometry have high sensitivity and accuracies [4] but require bulky instrumentation, extensive sample pretreatment processes, and sophisticated operators, which limit their application, especially in in situ analyses. A variety of new sensing platforms based on relatively low cost and simple operational routes are alternatives such as electrochemical methods, surface plasmon resonance detections, quartz crystal microbalance, chemiluminescence (CL), fluorescence methods, etc. [5], [6], [7]. Among them, fluorescence methods have the advantages of high sensitivity, high accuracy, and relative simplicity. The fluorescent sensor’s performance, such as sensitivity, reliability, dynamic range, and reproducibility, relies on selecting the appropriate fluorophore. The key factors that will affect the performance of fluorophore for HMs detection include fluorescence lifetime, fluorescence quantum yield (QY), Stokes shift, the width and shape of its absorption and emission bands, transmission and emission band maxima, molar absorption coefficients, etc. An effective fluorophore must be bright with a high molar absorption coefficient and high fluorescence QY under the optimal excitation wavelength. It should also be stable with sufficient functional groups, which could efficiently interact with the target HMs. Different fluorophores including small organic dyes, metal–ligand complexes, metal–organic frameworks (MOFs) [8], [ 9], semiconductors (NCs) and metal nanoclusters [10], silica nanoparticles (NPs) [11], and zero-dimensional carbon dots (CDs) [12] could be designed separately or combined as HMs sensing platforms. Among them, small molecular organic dyes such as rhodamine, fluorescein, cyanine, and coumarin are the most chosen in traditional fluorophore and broadly used. However, their disadvantages include small Stokes shifts (typically less than 25 nm), the excitation or emission wavelengths easily affected by environmental variations, and the self-quenching effect which limits their applications. In comparison with other fluorophores, CDs are alternative and superior to HMs fluorescent sensors.

CDs, which were first discovered from arc-discharge soot in 2004 [13], are defined as nanosized carbon-based materials with characteristic sizes of <10 nm and fluorescence as their instinct properties [14]. They could be divided into three categories according to their elementary structures. The CDs possessing graphene layers with chemical groups on the edges are usually called graphene quantum dots (GQDs), whereas the spherical carbon NPs with amorphous or nanocrystalline cores are named as carbon nanodots (CNs) and carbon quantum dots (CQDs), respectively, and CDs with polymer cross-linking or grafting are defined as polymer dots (PDs). CDs are outstanding electron acceptor and electron donator, which are commonly used in sensing applications either with fluorescence or CL detection. In the CL system, no photoexcitation is needed, and the luminescence is related to the electromagnetic radiation formed in the chemical reaction process [15]. CDs could be used as oxidants, emitting species, energy acceptors of chemical reaction energy, or even catalysis. The applications and reviews of CDs as sensors based on the CL principle could be found elsewhere and will not be discussed in this review [16], [17], [18], [19]. With fluorescence detection, photoexcitation is necessary to generate photoluminescence (PL) emissions, and CDs are thus good signal reporters.

In terms of CDs as HMs sensor, the general merits may be characterized as follows. First, CDs possess fluorescent optical properties with high photostability, low photobleaching, and hence are suitable as fluorophores for fluorescence detection of HMs. Secondly most of the CDs are hydrophilic with good solubilities in aqueous media, enabling their promising usage as HMs sensor in the environmental, biological, and medical fields [20]. With their carbon core and their easily tailored biocompatible functional groups on the surface, most of the CDs exhibit low toxicity, endowing them with great potential applications as fluorescent probes for HMs imaging in cells, bacterium, plants, and animals. In recent decades, there have been significant advances on CDs as fluorescence sensors for HM ions and their species and novel applications in environmental and food safety testing [21], cell and animal imaging [22], [ 23], chemical logic gates [24], anticounterfeiting ink [25], [ 26], fingerprint security, and document counterspy [27]. This review summarizes the recent advances on CDs, including CNs, CQDs, C-dots, carbon nitride dots (CNDs), GQDs, and PDs as fluorescence sensing platforms for HMs. The strategies to synthesize and purify CDs, their PL principles, the sensing mechanisms for HMs, and the binding strategies between CDs and HMs are concisely introduced. The signal detection mode and the progress on detection of single HM ions, multi-HMs, and different metal species are critically accessed. The development of solid-state sensor platforms and the applications in the cell, bacterial, plant, and animal bioimaging of HMs with CDs are presented. For consistency, the abbreviation “CDs” will be used in this review except for instances where more specific names need to be used.

3 Preparation CDs as HMs sensor

3.1 Synthesis of CDs

For the synthesis of CDs, there are a variety of synthetic routes. The desired routes are eco-friendly, size-controllable, convenient, cost-effective, environment friendly, and easy for large-scale production. In general, most of these routes can be categorized into “Top-down” and “Bottom-up” methods. In the “Top-down” approaches, CDs are prepared by breaking down bulk carbon materials such as graphite and carbon nanotubes, graphene, and suspended C-powders by electrochemical oxidation, acidic oxidation, laser ablation, plasma treatment, arc discharge, and so on. In the “Bottom-up” approaches, organic molecules such as glycerol, citric acid, ethylenediamine, amino acids and natural materials such as coffee grounds, soy milk, soybeans [28], grass, eggs, gelatin, pomelo peel, honey, etc., are candidates as carbon sources and precursor to generate CDs by hydrothermal, solvothermal, thermal decomposition, carbonization, ultrasonic and microwave experimental procedures [28]. For both “Top-down” and “Bottom-up” methods, reaction conditions such as selected reaction precursors, solvent, temperature, and reaction time are critical factors that will affect the CD’s chemical structure, size, and QYs [15], [18], [29], [30]. Until now, there is still a great challenge for controllable synthesis of CDs with precise chemical structure, desirable size, and high emission efficiency with high QYs due to the multistep reactions involved and complicate formation mechanisms during nucleation and growth processes. More reviews on the preparation strategies of CDs could be found in publications [12], [31], [32], [33], [34], [35], [36].

Concerning CD-based HM fluorescence sensors, the preparation of CDs with effective and sufficient functional groups is essential to provide binding or reaction sites with HMs. Heteroatom doping and surface modification are feasible methods. For heteroatom doping, carbon materials containing heteroatoms, such as nitrogen (N), sulfur (S), boron (B), silicon (Si), phosphorus (P) are promising candidates as the carbon source or precursor [23], [37], [38], [39], [40], and N- and S-doping of CDs are the most adopted strategies. By adopting different synthetic methods and synthesis conditions, the ratio of heteroatoms with the main element of carbon and the surface chemistry can be manipulated. Sometimes the ratio of the heteroatoms in doped CDs can vary significantly in raw and purified CDs, which shows a significant improvement of the optical properties of CDs. Compared with nondoped CDs, the excitation band of CDs may shift from typical 260–320 nm to longer wavelengths, and the emission band may shift to green, yellow, and red regions. As a result, the ability of antimatrix interference of CDs as HMs sensor was increased, as the natural fluorescence of the sample matrix at low wavelengths could be avoided. The QYs could also be improved with heteroatoms doping, which is beneficial to enhance the sensitivity of CD-based HM sensors. Many functional groups that can bind with HMs could be generated during the doping reaction, although the rational structure design is difficult. If the functional groups on the surface of heteroatoms doped CDs are lack of affinity and selectivity to target HMs, they can be further conjugated with other organic molecular, oligomeric polymers, and biomolecules through elimination, substitution, condensation, dehydration, carbonization, or passivation reactions to obtain specific binding sites for HMs.

3.2 Purification of CDs

For the successful application of CDs as HMs sensors, the purification is an essential and crucial preliminary step because the product of synthesized CDs is usually in a complex mixture containing unreacted starting materials and by-products alongside CDs. The unsuccessful purification will affect the QY, the fluorescent properties, and the ratio of heteroatoms in CDs, leading to significant challenges in understanding the structure and optical properties, generating the reproducible batch-to-batch results of reaction CDs, and further scaling-up application. However, because of their small size (< 10 nm) and the presence of plenty of polar functional groups on the surface, CDs show high colloidal stability in aqueous media and cause difficulty in obtaining high purify CDs. Newly reported purification methods include filtration, column chromatography, centrifugation, precipitation, pH control cloud point extraction [41], solid-phase extraction (SPE) [42], and dialysis. So far, dialysis is the predominant technique used to separate CDs from colloidal solutions using a semipermeable membrane. By choosing suitable molecular weight cutoff of dialysis bags, the small-sized target CDs penetrate through the membrane pores, while the bigger-sized nontarget compounds are kept in the dialysis bag. The main drawbacks of this method include the dilution of the dialyzed CDs solution, and the time-consuming process of dialysis. Centrifugation is another method reported for purification of CDs. Adjustment of centrifuge speeds can provide several fractions of NPs of different masses [43]. An organic solvent such as acetone could be added to separate the fractions in small mass ranges. Size-exclusion chromatography has also been applied to separate synthesized CDs products into several fractions, but the operation is tedious, and the elution conditions need to be carefully optimized. Koutsioukis et al. [42] reported an effective purification method by SPE after organic solvents extraction. CDs were first extracted into an organic solvent such as acetone and ethanol, further adsorbed on the surface of alumina as the solid phase, then eluted from alumina using water. Beiraghi and Najibi-Gehraz [41] applied pH-controlled cloud-point extraction (CPE) to separate two fractions using Triton X-114 as the enrichment phase and water as the back-extraction phase. Additional advances on purification techniques for CDs had been summarized by Kokorina et al. [44].

3.3 Photoluminescence (PL) principles

When CDs are excited under photons, the electrons jump to excited states and then return to the equilibrium states/ground state. The excess energy was released through the radiative process, which generates light emission known as PL. PL is the most attractive and important feature of CDs, determining their optical–optical properties. For fluorescence sensors, the detection is established based on the changes in the optical properties. Thus, a well understanding of PL principles helps design effective sensing strategies of CDs for HMs detection.

CDs with PL emissions could range from deep ultraviolet to near-infrared (NIR). The illumination of PL principles can guide exploring unique and novel applications of CDs, including HMs sensors. Although some controversy exists, four PL principles could be elucidated for CDs: the quantum confinement effect, surface/edge state, molecule state, and cross-link enhanced emission (CEE) effect [37]. The quantum confinement effect is mainly determined by conjugated π domains and cluster size. The emission wavelength was found to decrease with the decrease of the size of CDs, as the energy gap between the valence shell and conduction band widens [38], while the PL emissions increased with the increase of the fused aromatic rings as the bandgap of conjugated π domain narrows. The surface/edge state is related to the synergetic hybridization of the carbon core and connected chemical groups. Some reports found that more functional groups such as C=N/C=O and C–N groups are generated on the surface of the full-colored CDs than the CDs with single emissions. The molecule state relates to organic fluorophore on the surface of the carbon core, which can exhibit PL emission directly. When CDs were prepared at low reaction temperatures, small fluorophore molecules may be formed beside the carbogenic core obtained [39]. The exist of molecular fluorophores is beneficial to the improvement of QYs. CEE effect involved the formation of a new PL center induced by aggregation or cross-link. Yang’s group [40] found that the fluorescence of polyethyleneimine (PEI)-based PDs was amplified by the CEE effect, attributed to the decreased vibration and rotation in crosslinked PEI-based PDs structures. Commonly, the fluorescence of CDs is determined by one or a combination of these principles.

3.4 Sensing mechanism

The sensing applications for HMs detection using CDs were based on the phenomenons that the interactions between HMs and CDs either quench or enhance the fluorescence. The quenching or enhancing effect involve different sensing mechanisms including static and dynamic quenching mechanism, photoinduced electron transfer (PET), energy transfer (ET), inner filter effect (IFE), aggregation-induced emission quenching (ACQ), and aggregation-induced emission enhancement (AIEE) or coordination-induced aggregation (CIA) [45]. The quenching or enhancing of fluorescence intensity may be caused by one or simultaneously two sensing mechanisms, and some characterizations could be tested to verify and confirm the sensing mechanisms.

In terms of kinetic mechanism, static quenching occurs when the molecules form a complex in the ground state, i.e., before excitation occurs, through the interaction between CDs and HMs. In contrast, dynamic quenching happens when excited states return to the ground state by collisions between CDs and HM ions due to ET or charge transfer. To investigate the kinetic mechanisms of static and dynamic quenching, UV-Vis spectra, Stern-Volmer plot, and fluorescence lifetime decay in the absence and presence of HMs by time-correlation single photon counting method are usually investigated. In static quenching mode, the fluorescence lifetimes of CDs will have no change in the absence (τ 0) and presence (τ) of various concentrations of HMs. The absorption spectra of UV-Vis usually have some changes after the addition of HMs due to the formation of a complex in the ground state. Also, the calculated binding constants between HMs and CDs will reduce with increasing the temperature in the sensing system. On the contrary, in dynamic quenching, τ is shortened after the addition of HMs compared to τ 0; no changes would be observed from the UV–Vis of CDs with the addition of HMs. Moreover, the increase in binding constant along with the increase of temperature could be calculated, as the higher frequency of effective collisions between CDs and HMs would occur at a higher temperature.

If positively charged HM ions were added into negatively charged CDs solutions, an effective PET process might occur from CDs to HM ions. In that case, the excited state electron transfer from CDs to the lowest unoccupied molecular orbital’s of the HM ions, usually leading to quenching of the fluorescent intensity of CDs.

The ET include förster resonance energy transfer (FRET) and dexter energy transfer (DET). In FRET, the photonic energy of the donor (CDs) is acquired by the acceptor (HMs). The efficiency of ET between CDs and HMs is controlled by the extent of spectral overlapping between the emission of CDs and the absorption by the metal ions, the spatial distance (generally in the range of 10–100 Å) between CDs and HMs [46], and the fluorescence QYs of CDs. The fluorescence lifetime of CDs would decrease during the FRET process. In DET, the electron transfer is caused by redox reaction, not photoinduced. As mentioned, CDs are excellent electron acceptor and electron donator, which could act as oxidants and reductants, respectively. Therefore, DET might occur in the HMs sensing process using CDs as the match between the redox potentials of CDs and HMs is satisfied.

The IFE phenomenon occurs when the absorption spectra of the HMs precisely overlap with the excitation or emission spectra of fluorophore materials (CDs). As a result, the attenuation of the excitation beam or absorption of emitted radiation leads to the apparent fluorescence quenching of CDs. In IFE, no new substance is formed, and the average fluorescence lifetimes of CDs measured do not exhibit significant changes in the presence of HMs. As Cr (VI) has a broad absorption band which can effectively overlap with the emission or excitation spectra of CDs with blue light, the IFE strategy has often been applied to detect Cr(VI) [47], [ 48].

AIEE and CIA are related to the aggregation of CDs caused by temperature, solvent, additives like HMs [49], [50], [51]. When aggregation happens, the intramolecular motion is restricted, which blocks the nonradiative path and activated radiative decay. Characterization methods like transmission electron microscopy (TEM), dynamic light scattering, and small-angle X-ray scattering can be used to prove ACQ or AIEE processes directly. The variations of Zeta-potential and UV-VIS spectra could be measured to know the formation of HM ion–CD complexes indirectly.

4 CDs-HMs binding strategies

Fluorescent CD-based HMs sensors typically consist of a fluorescent component of CDs as a signal generator and probe elements for specific binding. Thus, the functionalized surface of CDs with different moieties is very crucial. The strong and selective binding between the moieties and target HMs are essential to selective and sensitive detection of HMs. The binding strategies could be coordination, chelation, electrostatic effects, oxidation–reduction reactions, and interactions with deoxyribonucleic acid (DNA) aptamers.

4.1 Coordination and chelation

Coordination and chelation are the most common and straightforward strategies to design CD-based HMs sensors. The empirical hard and soft acid and bases (HSABs) principle could be an essential consideration for designing suitable ligands to a specific HM. According to the HSAB principle, soft HMs like Ag(I), Hg(II), Cd(II), and MeHg(I) prefer to form covalent bonds with soft ligands such as –SH, –S–S, and –S–R; transitional metals including Fe(II), Co(II), Ni(II), Cu(II), Zn(II), and Pb(II) trend toward binding borderline ligands like –CO–NH–, >N=H, –C6H4NH2; and hard HMs of Mg(II), Sr(II), Mn(II), Al(III), Cr(III), Fe(III), U(II) have priority to bind with coordinating groups of –OH, –F, –Cl, –PO4 3−, –OOCR, –OR, –NH2, –SO3H, >PO4, and –NH2R [52]. Of course, it still needs to notice that the rule is not absolute, especially for transitional HMs. The interaction with hard oxygen/amino groups and soft thiol groups is observed in many complexes. Different coordination structure modes formed like linear or tetrahedral coordination, size-controlled macromolecule ligands involved, and other factors all can affect the association constant. Furthermore, the hardness of different species of the same HM also can vary a lot and needs to adopt different coordination ligands. For instance, As(III) is a typical soft metal, while As(V) is an analog of PO4 3− belonging to the hard metal group. Also, considering an abundance of interference ions existed in real sample matrices and most of the synthesized CDs with many oxygen-containing functional groups, the interference caused by nontarget ions needs to be systematically investigated, rendering the sufficient selectivity.

To render the surface of CDs with moieties to bind strongly with HMs, precursors with suitable ligands according to the HSAB principle could be chosen. Mainly heteroatoms like S, N, O, and P could be functionalized as ligand atoms in the form of chemical groups like –SH, –S–S, –NH2, =NH, –OH, –OPO3H, or >C=O. These as-prepared CDs could be used as nonlabel probes to detect HMs directly. For instance, thiol groups (–SH) provide an excellent binding affinity to Hg(II), thus was commonly induced to detect Hg(II), even though some drawbacks such as undesired oxidation of sulfides during long-term storage at ambient temperature and possible ineligibility for use in sulfur-rich environments need to be considered and carefully avoided. Ding et al. [49] prepared N-, S-doped CDs using glycerol as the reaction solvent and cystine as the source for C, N, and S. The resulting CDs can specifically interact with Hg(II).

Except for the nonlabeled CDs with suitable ligands as HM sensors, prepared CDs could be further modified with functional groups to increase selectivity via surface chemistry or other interactions like covalent bonding, coordination, ππ interactions, and sol–gel technology. CDs with carboxyl groups were prepared by Ottoor et al. [53] using jeera as a carbon source, then capped with cystamine by carbodiimide chemistry, where 1-(3-dimethylaminopropyl)-3-ethyl carbodiimide (EDC) is used as an effective crosslinking agent along with N-hydroxysuccinimide (NHS). The cystamine functionalized CDs show much selective quenching response to Cr(VI) compared to unconjugated CDs toward various interference metal ions. The enhancement of selectivity was attributed to the formation of Cr(VI)-cystamine complex, facilitating the charge transfer between cystamine-CDs and Cr(VI). In Kaur et al.’s work [54], surface modification of their as-synthesized CDs with calix[4]arene derivative was done by stirring the reaction mixture of 1,3-disubstituted calix[4]arene and CDs in DMSO at room temperature, FT-IR was used to monitor the course of the substitution at the surface of the CDs. These modified CDs have high selectivity to detect Zn(II) against the other interference metal ions such as Cd(II) and Hg(II) with similar nature. The high selectivity benefits from the structure of calix[4]arene with plentiful rigid cavities of unalterable sizes, which can embrace Zn(II) with exact size.

4.2 Electrostatic effect

If the binding sites on the surface of CDs involve ionizable functional groups at certain pH ranges, electrostatic effects may exist between CDs and HMs. For hydroxyl, carboxyl, thiol, sulfonate, and phosphonate groups, they will be deprotonated and negatively charged when the pH value of the medium exceeds their pKa values; For amine, imine, amide, and imidazole groups, they will be positively charged when protonated at pH lower than their pKa. In acidic medium, anionic metal species like Cr(VI) tend to bind better with positively charged groups like amine; cationic metal species reach their maximum binding in a neutral medium when most of the acidic binding groups like carboxyl are deprotonated, and the cationic metals are not precipitated. Therefore, in this strategy, pH conditions will affect sensor performance. Li et al. [40] fabricated Tb(III)-2, 6-pyridinedicarboxylic acid-N-doped CDs (Tb-DPA-CDs) sensor system for detecting Hg(II) in seafood. The quenching efficiency of N-CDs was found to increase in the pH range of 2.0–7.0. That is because the deprotonation of –COOH on the surface of CDs increased with the increase of pH, leading to the negatively charged –COO, which could maximally capture the oppositely charged Hg(II).

4.3 Oxidation-reduction reaction

Photoexcited CDs are reported as excellent electron donors and can reduce metal ions in solution [55], [ 56]. Therefore, an oxidation-reduction reaction could be employed to design CDs based HM sensor. The oxidation–reduction reaction can occur between HMs and CDs, which generate electron transfer and led to nonradiative electron/hole recombination, resulting in the PL changing. Huang et al. [57] prepared fluorescent N-doped GQDs (N-GQDs) to the selective determination of Cr(VI) based on the quenching effect. The quenching mechanism was proved that parts of Cr(VI) were reduced to Cr(III) on the surface of N-GQDs by X-ray photoelectron spectroscopy characterization. The quenching equilibrium reached in 3 min, which means the kinetic behavior of oxidation–reduction reactions between N-GQDs and Cr(VI) was fast. Wang et al. [58] also found that the fluorescence intensity of fabricated CDs was significantly and selectively quenched in the presence of Cr(VI) due to the oxidation–reduction reaction between Cr(VI) and the oxygen-containing groups and S-related species on the surface of CDs.

4.4 DNA aptamer affinity

Aptamer oligonucleotide sequences can selectively bind with HMs to form stable metal-mediated base pairs by groove binding, electrostatic interaction, and intercalation [59], [60], [61], [62]. For instance, DNA with multiple thymine (T) sequences was well known to specifically bind to Hg2+ to form a T–Hg2+–T structure with two T residues. Their binding undergoes proton substitution reactions to coordinate with Hg(II) via Hg–N bonds. The binding strength of the T–Hg(II)–T complex is even higher than that of the base T-A pair [63]. Thus, T-rich sequences could be applied as recognition molecules to selectively detect Hg(II). DNA-based sensors are rapid, low cost, and suitable for real-time detection due to the stability of DNA. Now, a lot of methods based on DNA detection of HMs have been developed, such as colorimetric detection, electrochemistry, and surface plasmas resonance, fluorescence, and others [64], [ 65]. Fluorescence methods usually involved the chemical conjugating of DNA with a fluorophore, e.g., dyes [64], metal NPs, and quantum dots [66] as a molecular recognition probe. Until now, few works have been reported using CD-based sensors employing aptamer affinity to selectively detect HMs [67].

Carboxyfluorescein (FAM)-DNA macromolecules were strong assembled on the surface of the N-CDs through ππ stacking by Ni et al. [68]. With the aid of 6-mercaptopurine (6-MP), a fluorescence peak at 525 nm of NCDs–ssDNA-6-MP conjugation system was observed. With the further addition of Hg(II), the large system of NCDs–ssDNA-6-MP was broken up due to the formation of stronger T–Hg(II)–T complex, resulting in the quenching of the fluorescence at 525 nm. Thus, the Hg(II) was detected by this quenching effect. Salimi and Hamd-Ghadareh [69] designed a fluorescence aptamer sensor system for quantitation of Hg(II) based on the hybridization of T-rich ssDNA immobilized on Rhodamine B(RB)–CDs (RB–CDs–ssDNA) and gold nanoparticles (AuNPs) modified with complementary aptamer (AuNPs-cDNA). By excitation at 450 nm, RB–CDs–ssDNA nanohybrid showed the dual well-resolved emissions at 580 and 668 nm, respectively. With the addition of AuNPs-cDNA, the affinity binding of DNA strands decreases the distance between RB–CDs–ssDNA and AuNPs-cDNA and produced enhanced FRET pairs, resulting in a decrease of fluorescence. In the presence of Hg(II) as the analyte, the hybridized DNA strand will dehybridized due to the high affinity of ssDNA toward Hg(II), and AuNPs-cDNA was released from the assemblies, resulting in the increase of fluorescent intensity at 580 nm and decrease at 668 nm. The Hg(II) was thus high selectively detected by calculation of the I 580/I 668 intensity ratio.

The fluorescence of N/Ce-doped CDs prepared by Liang et al. [70] was found to be quenched by As(III)-specified aptamer via the formation of a large CD–aptamer conjugate due to ππ electron stacking, hydrogen bond, hydrophobic force, and intermolecular force. When As(III) was added, As(III) can specifically react with aptamer to form a stable aptamer–As(III) complex, resulting in the restoration of fluorescence intensity in this As(III) sensing system.

Yang et al. [65] labeled the synthesized CDs with DNA as fluorophore and use graphene oxide (GO) as a quencher to quench the fluorescence through ππ stacking interaction. The detection of Hg(II) was realized by measurement of fluorescence recovery in the presence of Hg(II) due to the release of DNA-CDs from GO through the formation of T–Hg(II)–T duplex. Similarly, Srinivasan et al. [71] adopted the MoS2 nanosheet as a quencher to quench the fluorescence of DNA functionalized CDs and detect the Hg(II) by the recovery of the fluorescence signals. Wei et al. [72] reported an amplified assay for Hg(II) determination based on an ssDNA-labeled FAM-CQDs hybrid. Firstly, the quenched fluorescence of ssDNA-labeled FAM by QCDs was restored in the presence of Hg(II) like other works above. With the further addition of DNase I, a 110% signal increase was obtained over that without DNase I. The reason is that the nuclease of DNase I can cleave the ssDNA–Hg(II) complex and release the Hg(II). The released Hg(II) then binds another ssDNA, and the cycle starts, which leads to amplification of the signal. With this amplification strategy, a 20-fold lower limit of detection (LOD) than that of traditional unamplified homogeneous assays was achieved.

Alternatively, Song et al. [73] directly used DNA as a carbon source to synthesize CDs by hydrothermal method. The characterizations showed residual thymine (T) and cytosine (C) groups are still existing on the CDs, which could strongly bind with Hg(II) and Ag(I), respectively, resulting in a quenched fluorescence.

More aptamer-based CD sensors for HM detection are summarized in Table 2.

Table 2:

Aptamer-based CDs fluorescence sensors for HMs detection.

HM Signal response DNA sequence LOD Sample Reference
Hg(II) RF 5′-FAM-CTT GTT GTC CTG TTG TTC-3′ 1.26 nM Serum, water [68]
Hg(II) ON–OFF–ON 5′-NH2-(CH2)6-TTCTTTCTTCGCGTTGTTTGTT-3′-CDs 2.6 nM Citrus leaf [65]
Hg(II) On–OFF–ON 5′-TTCTTTCTTCCCCTTGTTTGTT-FAM-3′-CDs 0.5 nM Water [72]
Hg(II) On–OFF–ON 5′-NH2-(CH2)6-GTTTCTTCTTTGGTTTGATT-3′-CDs 1.02 nM Water [71]
Hg(II) On–OFF 5′-NH 2 C6TTCTTTCTTCCCTTGTTTGTT-3′-GQDs 0.25 nM Cell imaging [145]
Hg(II) OFF–ON 5′-TTC TTT CTT CCC CTT GTT TGTT-FAM-3′ 0.5 nM NA [72]
Hg(II)

Ag(I)
On–OFF sigma number 31149-10G-F 48 nM;

0.31 μM
Water [73]
As(III) On–OFF–ON ATG CAA ACC CTT AAG AAA GTG GTC GTC CAA AAA ACC ATT G 0.43 μg/L Water [70]
Pb(II) On–OFF–ON 5′-NH2-(CH2)6-GGGTGGGTGGGTGGGT-3′ 0.6 nM NA [67]
Pb(II) On–OFF–ON 5′-3′GGT TGG TGT GGT TG 0.00025 μM Water [146]
Hg(II) On–OFF–ON 5′-TTCTTTGTTCCCCTTCTTTGTT-NH2-3′ 5  nM Water, serum, cell imaging [69]
  1. DNA, deoxyribonucleic acid; RF, ratiometric fluorescent; HM, heavy metal.

5 Signal response

5.1 Turn-on

To date, most of the fluorescent probes in detecting HMs are based on a fluorescence quenching (“turn-off”). For fluorescent probes with turn-off response, interference ions or compounds in the matrix may also quench the fluorescence, lowering the selectivity and limit the practical applications [49]. In comparison, fluorescent probes with turn-on (switch-on or OFF–ON) responses could reduce false-positive signals and the dark background, resulting in increased selectivity. Therefore, it is highly desirable to establish fluorescent ‘turn-on’ probes to detect HMs. AIEE and chelation-enhanced fluorescence (CHEF) effect is most employed to design turn-on probes. The AIEE-based fluorophores show non- or very weak emission in its dissolved state at first. When target HMs were added to form aggregation with the intramolecular motion restriction, the fluorescence turn to show strong emission. The CHEF effect of fluorophores is attributed to the enhancement of fluorophore when some types of complexation with nonquenching HMs are formed. Different strategies such as doping, conjugating, oxidation–reduction reaction, and OFF–On response could be adopted to make turn-on fluorescent sensors. Up to now, turn-on sensors mainly focus on organic dyes, while rarely on fluorescent CDs.

Ding et al. [49] found that the prepared N-, S-doped CDs have a turn-on fluorescence response to Hg(II) based on AIEE, and thus used as selective Hg(II) probe. The aggregation is attributed to the formation of a new complex between CDs and Hg(II) by S–Hg bonds, and the larger percentage of sulfur content of CDs contained, the higher enhancement of the fluorescent intensity caused by AIEE was observed.

In Zhang et al.’s method [50], amine-functionalized CDs were firstly prepared and purified, then conjugated with reduced glutathione (GSH) by the carbodiimide-activated coupling using EDC/NHS. Benefit with free carboxyl and amino groups of GSH to provide binding sites for metal cations, the fluorescence enhancement of resultant GSH-CDs in the presence of Fe(III) ions was observed, and the “turn-on” probe for Fe(III) detection was thus constructed.

Guo et al. [51] designed a fluorescence “turn-on” sensing for Pb(II) detection via CDs immobilized in spherical polyelectrolyte brushes (SPBs) with “OFF–ON” response. The fluorescence of CDs was firstly ”turned off” via electrostatic interaction induced quenching by SPB, then specifically turned on by the addition of Pb(II) via the aggregation of the SPB particles. Zhang et al. [74] also developed a “turn on” fluorescence CDs sensor for selective Hg(II) determination based on bis(dithiocarbamato)copper(II) (CuDTC2)-conjugated CDs (CuDTC2-CDs). They found that the CuDTC2 complex at the CDs’ surface could strongly quench the fluorescence of the CDs by a combination of electron transfer and ET mechanism. With the addition of Hg(II), the quenched fluorescence of the CuDTC2-CDs was switched on because Hg(II) displaced the Cu(II) in the CuDTC2 complex, and thus shutting down the ET pathway.

Chen et al. [75] constructed a turn-on fluorescence sensor with green luminescent emission to detect Ag(I) based on the oxidation–reduction reaction that occurred on the surface of CDs. Because the reduction potential of Ag(I)/Ag is relatively high (0.7991 V), Ag(I) has precedence over other metal ions to be reduced to Ag0, which could enhance the radiative emission of CDs.

In Table 3, a list of turn-on CD-based fluorescence sensors for HM detection is given.

Table 3:

The applications of CDs as HMs sensor with turn-on signal response.

HM Sensor system Strategy Mechanism LODs Sample Reference
Hg(II) N,S-doped CDs Doping AIEE 0.5 μM Water [49]
Fe(III) GSH-conjugated CDs Conjugating AIEE 0.1 μM Cell imaging [50]
Hg(II) CuDTC2 conjugated-CDs OFF–ON FRET 20 nM Water [74]
Pb(II) SPB-CDS OFF–On AIEE 22.8 μM Water [51]
Cd(II) N,P-CDs Doping AIEE 0.16 μM Serum, urine [147]
Ag(I) CDs-AgNCs Redox reaction ND 320 nM Water [75]
Al(III),

Zn(II)
CDs Doping CHEF 4 nM

100 nM
NA [148]
Zn(II) CDs-sCdTe QD-EDTA OFF–ON ND 0.33 μM NA [149]
  1. ND, not discussed; NA, not analyzed; AIEE, aggregation-induced emission enhancement; FRET, förster resonance energy transfer; CHEF, chelation-enhanced fluorescence; GSH, glutathione; SPBs, spherical polyelectrolyte brushes; HM, heavy metal.

5.2 Ratiometric measurement

Even though most of the fluorescent sensors were designed using a single emission wavelength for quantification by either off or on effect, the drawbacks are obvious. Many analyte-independent factors could affect the fluorescence signals including photobleaching of the fluorophore, light scattering caused by the sample matrix, fluctuations in the excitation source, and the changes of the solution environment. Ratiometric fluorescent (RF) sensors, which detect the analytes by measuring the change of the ratios of fluorescence intensities at two wavelengths, could provide a built-in self-calibration as an internal standard to correct analyte-independent parameters and eliminate the impact of false signals caused by the background fluorescence mentioned above. Therefore, RF measurement is expected to be more reliable and accurate. RF sensors can be designed by reference mode and dynamic mode. In reference mode, a fluorescence channel without a signal change in the presence of target analytes acts as a reference and another channel with signal increase/decrease/shift upon analyte’s presence as a signal reporter. In dynamic mode, two dynamic channels with signal changes in opposite or same directions with the exist of analytes were applied [76]. Fluorescent organic dyes [77], lanthanide elements [78], and quantum dots [79], [ 80] could be employed to design RF-based sensors. Among them, organic fluorophores were most applied but usually needed elaborate molecular design and complex synthesis. RF-based CDs sensors are relatively less reported but are very promising. Different strategies could be utilized to design RF-based CD sensors for HMs detection.

5.2.1 CDs with multichannel emissions

The construction of a ratiometric probe based on dual-emission fluorescent CDs is the simplest effective way. Zhang et al. [81] obtained a CD product solution exhibiting two strong fluorescence peaks at 470 and 678 nm under a single excitation of 406 nm by solvothermal treating of corn bract, and the obtained CDs product was used as an RF-based sensor for Hg(II) detection. The NIR emission of 678 nm was a signal peak quenching by Hg(II), while an emission peak of 470 nm was constant to Hg(II) used as a reference. Dong et al. [82] constructed a ratiometric fluorescence assay with N-CDs to detect Ag(I) based on the ratio (I 618/I 532) of rising peak emerges at 532 nm and the decreasing peak at 618 nm with the introduction of Ag(I). Ding et al. [83] prepared multicolor CDs with tunable fluorescence emissions from violet to red without obvious compromising the fluorescence intensities. For RF measurement of target Fe(III), the CDs were excited under two wavelengths of 300 and 540 nm, resulting in emission at 368 and 605 nm, respectively. Both emission peaks under different excitation wavelength were gradually decreased with the increase of Fe(III) concentration. The fluorescence intensity ratio at emissions of 368 and 605 nm was thus calculated to determine the Fe(III) concentration. In this approach, twice records of emission spectra for each test are needed and tedious.

5.2.2 Mixture of different kinds of CDs

The RF-CDs sensor also can be realized by choosing two suitable CDs to mix as a sensor system. Two kinds of CDs with different emission ranges of red and blue were synthesized separately by Yu et al. [84] and mixed to provide reference fluorophore and signal response fluorophore, respectively. The CDs with blue emission (B-CDs) was prepared with gelatin and 3-aminobenzeneboronic acid as carbon sources by hydrothermal method, while the CDs with red emission (R-CDs) was synthesized by phenylenediamine solvothermal method. Upon addition of Hg(II), the fluorescence intensity of B-CDs continuously decreases along with the increasing concentration of Hg(II), while the fluorescence intensity of the R-CDs remains almost unchanged. Zhao et al. [85] reported a similar approach to design RF-CDs as Cu(II) sensor by mixing two color CDs (Figure 1). The mechanism is interesting: Cu(II) ions can efficiently bind to the red CDs (R-CDs) to produce a strong visible absorption, which overlaps the emission of blue CDs (B-CDs). The adsorption of small B-CDs onto the surface of larger R-CDs with existing of Cu(II) in solution was observed by TEM due to the dual-coordinating interactions of Cu(II) with the surface ligands of both R-CDs and B-CDs. As a result, FRET could occur between B-CDs and Cu(II)-R-CDs, leading to quenching of the fluorescence of B-CDs, and unchanging of the fluorescence of R-CDs. Gogoi et al. [86] reported a carbon-based nanohybrid system comprising GSH-functionalized reduced CDs (GSH-f-rCDs) with emission in the violet region as signal reporters and green-emitting GQDs as reference for As(III) detection with a lower detection limit of 0.5 ppb.

Figure 1: 
(A) Transmission electron microscopy (TEM) image of small carbon dots with blue emission (b-CDs) adsorbed onto lager Red-CDs after the addition of Cu(II) ions; (B) schematic illustration of the conjugation of blue-CDs and red-CDs through a Cu(II) ions bridge. (C) Schematic of the ratiometric fluorescent (RF) sensor designed to determine Cu(II) by mixing red-CDs and blue-CDs; Reprinted from Zhao et al.’s work with kind permission of American Chemical Society [85]. Copyright (2017) American Chemical Society.
Figure 1:

(A) Transmission electron microscopy (TEM) image of small carbon dots with blue emission (b-CDs) adsorbed onto lager Red-CDs after the addition of Cu(II) ions; (B) schematic illustration of the conjugation of blue-CDs and red-CDs through a Cu(II) ions bridge. (C) Schematic of the ratiometric fluorescent (RF) sensor designed to determine Cu(II) by mixing red-CDs and blue-CDs; Reprinted from Zhao et al.’s work with kind permission of American Chemical Society [85]. Copyright (2017) American Chemical Society.

5.2.3 Combination of CDs with other fluorophores

The design of the RF-CD–based sensing system for HMs detection can also combine CDs with other fluorophores such as organic dyes, metal clusters, quantum dots, and MOFs, as long as they can be excited at the same time. The combination can be the direct mixture of CDs and other fluorophores, postmodification of CDs with other fluorophores, or covalent bonding between CDs and fluorophores. CDs can be applied as either reference fluorophore or signal response fluorophore.

A dual-signal RF sensor of CDs/AuNCs nanosystem was developed by He et al. [87] by simply mixing CDs with Au nanoclusters (AuNCs) to detect Ag(I). CDs and AuNCs have no interaction with each other. CDs served as reference fluorophore, and Au NCs served as a recognition unit with turn-on response. Wu et al. [88] developed a specific RF probe by adding Rhodamine B (RhB) as a reference to CDs as a sensing signal for sensitive detection of Hg(II). He et al. [89] synthesized nitrogen and cobalt(II) codoped CDs encapsulated in europium MOFs (CDs@Eu-MOFs) as a RF sensor for the detection of Cr(VI). The fluorescence intensity of the CDs decreased and the intensity of the Eu-MOFs remained in the increase of Cr(VI) concentration. Lu et al. [90] embedded CDs and AuNCs into classical MOFs of ZIF-8 to form CDs/AuNCs@ZIF-8 nanocomposites for RF detection of Cu(II) and CDs served as reference. Mahmoudi et al. [91] combined blue emission CDs (B-CDs) and thioglycolic acid (TGA)-capped yellow emissive cadmium telluride quantum dots (TGA–CdTe–QDs) to consist of ratiometric fluorescence probe for Hg(II) detection. With the exposure of Hg(II) ions to the probe, the fluorescence of TGA–CdTe–QDs was selectively quenched and red-shifted, while the blue emission of CDs remained constant as an internal standard.

Wang et al. [92] conjugated FTIC dye with PEI-modified CDs to fabricate CDs-FITC composites as Cu(II) RF sensor and observed that the spectral peaks both at 412 nm (for CDs) and 518 nm (for FITC) gradually decreased with the increase of Cu(II). Tan et al. [93] conjugated a europium complex onto the surface of CDs to form dual-fluorophore NPs of CDs–BHHCT–Eu(III) (Figure 2) and applied them in the RF detection of Cu(II) with emission at 615 nm as signal and 410 nm as reference.

Typical RF CD-based fluorescence sensors for HMs detection are listed in Table 4.

Figure 2: 
Preparation of carbon dots (CDs)-BHHCT-Eu(III) dual-fluorophore system. Reprinted from Tan et al. works with the kind permission of The Royal Society of Chemistry [93]. Copyright (2014) The Royal Society of Chemistry.
Figure 2:

Preparation of carbon dots (CDs)-BHHCT-Eu(III) dual-fluorophore system. Reprinted from Tan et al. works with the kind permission of The Royal Society of Chemistry [93]. Copyright (2014) The Royal Society of Chemistry.

Table 4:

Summary of RF CD-based sensors for HMs detection.

HMs RF mode CD sensor system LODs Sample Reference
Cr(VI) Mode 1 CDs at 430 nm(D), 510 nm(D) 0.4 μM Textile, steel, industrial wastewater [150]
Hg(II) Mode 1 CDs at 406 nm, 678 nm 9 nM Serum, water [81]
Cu(II) Mode 1 CDs at 412 nm, 553 nm 23 nM Water [151]
Ag(I) Mode 1 CDs at 532 nm, 619 nm 0.27 μM Cell imaging [82]
Pb(II) Mode 1 CDs at 477 nm, 651 nm 0.055 μM NA [152]
Fe(III) Mode 1 CDs with multicolor from violet to red 10 nM Water [83]
Cu(II) Mode 2 B-CDs, R-CDs 37 nM Water [85]
As(III) Mode 2 GSH-f-rCDs(S)+GQDs(R) 0.5 ppb NA [86]
Hg(II)

Cr(VI)
Mode 3 CDs (R)+ BSA-Au NCs(S)

BSA-Au(R) + CDs (S)
1.85 nM

5.34 nM
Water, human blood [107]
Hg(II) Mode 3 BSA-CDs(R)+ AuNCs(S) 0.73 nM Water [153]
Ag(I) Mode 3 CDs(R)+AuNCs(S) 2 nM NA [87]
Cd(II) Mode 3 CDs(R)+AuNCs(S) 32.5 nM NA [154]
Cu(II) Mode 3 CDs(R)+ Au NCs(S) 0.33 nM Serum [90]
Hg(II) Mode 3 CDs(S)+SiNCs 7.63 nM NA [155]
Zn(II) Mode 3 Fe doped CDs(S)+ Au NCs(R) 0.1 μM NA [156]
Cr(VI) Mode 3 RhB -doped silica nanoparticles(R)+CDs(S) 1.3 ng/mL Water [89]
Hg(II) Mode 3 CDs(S)+ RhB dye(R) 25 nM serum [88]
Fe(III) Mode 3 CDs(S)+RhB dye(S) 0.73 μM NA [157]
Cu(II) Mode 3 CDs(R)+ RhB dye(S) 2 ppb Water [158]
Hg(II) Mode 3 CDs(S)+ RhB dye(R) 30 nM Water [159]
Hg(II) Mode 3 NCDs(S)-RhB(S)@COF 15.9 nM NA [130]
Cr(VI) Mode 3 Eu-MOFs(R)+ Co doped CDs(S) 0.21 μM Water [160]
Hg(II) Mode3 CDs(R)+ CdTe-QDs(S) 4.6 nM Water [91]
Cu(II) Mode 3 CDs(R)+CdTe-QDs (R) 0.36 nM Water [161]
Hg(II) Mode 3 CDs(S)+ CdTe@SiO2(R) 0.47 nM Water [162]
Fe(III) Mode 3 CDs(R)+CdSe QDs(R) 5.4 μM Cell imaging [163]
Hg(III) Mode 3 CDs(R)+ CdSe @ ZnS QDs (S) 0.1 μM Water [164]
Cu(II) Mode 3 CDs(R)+Mn‐doped ZnS QDs(S) 33.1 nM NA [165]
Cu(II) Mode 3 CDs(R)@Eu-DPA MOFs(S) 26.3 nM NA [166]
Cu(II) Mode 3 CDs(R)/QDs(S)@ZIF-8 1.53 nM Water [167]
Cu(II) Mode 3 PEI-CDs+FTIC dye 7 nM Serum, yogurt [92]
Cu(II) Mode 3 CDs(R)–BHHCT–Eu(III)(S) 4 nM Water [93]
  1. Mode 1: CDs with multichannel emissions; Mode 2: Mixture of different kinds of CDs; Mode 3: Combination of CDs with other fluorophores; ND, not discussed; NA, not analyzed; MOFs, metal-organic frameworks; GQDs, graphene quantum dots; HM, heavy metal; RF, ratiometric fluorescent; PEI, polyethyleneimine.

6 Detection of HMs in solution

6.1 Single HM detection

Until now, most efforts have been made to improve the sensitivity and selectivity of the CD-based fluorescence sensors to detect single HMs either based on turn-on or turn-off effect [94], [ 95]. Devi et al. [96] reviewed many applications on the determination of single HM in water.

6.2 Multiple HMs detection

To date, CD-based fluorescence sensors designed for the detection of two or more kinds of metal ions simultaneously have seldom been reported due to the lack of separation ability of fluorescence detection. From recent publications focusing on multi-HMs detection, the following strategies have been adopted for multi-HMs sensing.

6.2.1 Indistinguishable detection

When the prepared CDs have similar fluorescence responses to multi-HMs [97], [ 98], they are referred to as indistinguishable detection sensors for HMs detection. In this situation, the detection is based on the total changes caused by multiple ions and cannot discriminate against each ion. Thus, it is hard to be used in real applications. Rajendiran et al. [98] found that the CDs obtained by a hydrothermal process using jackfruit (Artocarpus heterophyllus) as a carbon source can be quenched by both Hg(II) and Cr(VI), with limits of detection for the determination of Hg(II) and Cr(VI) of about 8 and 10 nM, respectively.

6.2.2 Discrimination by CDs’ emission bands

In real applications, multi-HMs needs to be discriminated, and the discrimination could base on the different responses of multi-HMs to CDs’ emission band. For example, if two HM ions have opposite responses to one emission band of prepared CDs, CDs can be used to differentiate and determinate these two metal ions. Nag et al. [99] found as-prepared nitrogen-doped red-emitting CDs can be used for dual sensing of In(III) and Pd(II) based on the “turn-on” fluorescence response with a red-shift to Ln(III) and the “turn-off” response to Pd(II).

Also, if the synthesized CDs have multichannel emissions and these emission peaks exhibit different response patterns for different HM ions, they could be employed to detect two or more HM ions. The tuning of fluorescence emissions can be realized by choosing suitable precursors, optimizing synthesis conditions and methods, changing the contents of heteroatoms in CDs [100]. Liu et al. [101] prepared blue and red dual-emission CDs by a solvothermal method in a water–formamide binary system using citric acid and ethylenediamine as precursors. After modification, the red channel was used to detect Pb(II), and a blue channel was applied to detect Cr(III), respectively, by quenching mechanism. Similarly, dual-emission CDs containing blue emitters (386 nm) and yellow emitters (530 nm) were obtained by You et al. [102] and applied to simultaneously detect Fe(II), Bi(III), and Al(III) based on their different responses on emitters. Fe(II) was detected by quenching effect on blue emitter at 386 nm; Bi(III) and Al(III) were detected and differentiated by the yellow emitter at 530 nm, which showed the decrease of fluorescence with the addition of Bi(III) and the enhancement of fluorescence with blue-shift upon the addition of Al(III). Chai et al. [103] synthesized dual‐emissive fluorescent GSH-CDs with emission at 460 and 683 nm, and successfully detected Zn(II), Mn(II), and Cu(II) ions based on their different response at two emission wavelengths of GSH-CDs. The addition of Zn(II) only would decrease peak intensities of both 460 and 683 nm, resulting in the fluorescent color turning from cyan to pink under UV light. With the addition of Mn(II) and Cu(II), the red emission peak at 683 nm reduced remarkably, but the blue emission peak at 460 nm has different changes towards Mn(II) and Cu(II). The addition of Mn(II) led to the slight redshifted at 460 nm but no shift with Cu(II). The change in the fluorescence spectra at 460 nm was thus utilized to distinguish Mn(II) and Cu(II) ions in real samples. Alternatively, Ma et al. [104] employed their prepared CDs to detect Cu(II) at red emission of 671 nm based on the ACQ effect and Al(III) at blue emission of 478 nm by the AIEE effect.

6.2.3 Differentiation by CDs contained multifluorophores

Multichannel emissions of CD-based sensing systems also could be constructed through assigning two or more types of fluorophores in the sensing system to detect multi-HMs. Zhang et al. [105] mixed two types of multiemission CDs to form nanohybrids for the determination of Pb(II) and Hg(II). Interestingly, these two CDs were synthesized from the same natural biomass with and without the addition of Na2CO3, respectively, during solvothermal treatment. In the presence of Pb(II), the decrease in fluorescence at 493 nm and the increase in fluorescence at 653 nm were observed, and a peak intensity ratio of I 653/I 493 was recorded to determine Pb(II) concentration. With the existence of Hg(II), fluorescence intensity at 611 nm decreased with a redshift occurring and no changes at 491 nm. Therefore, the I 611/I 491 ratios were calculated to determine Hg(II) concentrations. Shen et al. [106] mixed as-prepared three kinds of CDs from different initial materials, which had different specific responses to three metal ions (Hg(II), Fe(III), and Cu(II)), and developed a smartphone-based three-channel ratiometric fluorescence device for simultaneous determination of these three target metal ions. Su et al. [107] designed a dual emissive multicolorful fluorescent nanoprobe of Au NCs-CDs (gold nanocluster-CDs) to simultaneously detect two ions of Hg(II) and Cr(VI) (Figure 3). When Hg(II) was present without Cr(VI), the signal peak of Au NCs at 625 nm decreased, and the reference peak of CDs at 444 nm remain consistent, leading to the fluorescence color changing from pink to blue under UV light. In the presence of Cr(VI) without Hg(II), the fluorescence emission of CDs was quenched, while the emission of Au NCs was stable and was used as an internal reference. When both ions were exposed to the Au NCs-CDs, the dual-emissions of the Au NCs-CDs were all decreased.

Figure 3: 
The schematic illustration of the preparation of gold nanoclusters (Au NCs)-carbon dots (CDs) and detection for metal ions of Hg(II) and Cr(VI). Reprinted from Su et al.’s work with the kind permission of Elsevier [107]. Copyright (2018) Elsevier.
Figure 3:

The schematic illustration of the preparation of gold nanoclusters (Au NCs)-carbon dots (CDs) and detection for metal ions of Hg(II) and Cr(VI). Reprinted from Su et al.’s work with the kind permission of Elsevier [107]. Copyright (2018) Elsevier.

6.2.4 Discrimination by restoration reagent

If the fluorescence of CDs is quenched by multiple metal ions simultaneously, the restoration reagent could be added to discriminate each ion further. Gogoi et al. [108] successfully developed a CD-based sensor to detect Hg(II) and Cu(II) in parallel by an on–off–on process with the addition of vitamin C and trisodium citrate, respectively (Figure 4). With the addition of trisodium citrate, the quenched fluorescence from the Cu(II)-CDs mixture was recovered, while the fluorescence from Hg(II)-CDs did not recover, which was attributable to the fact that the Cu(II)–SC adduct was more stable than the Cu(II)–CDs adduct. Similarly, the fluorescence of CDs quenched by Hg(II) ions could be recovered with vitamin C, while the quenched fluorescence of Cu(II)–CDs remained unchanged. The reason was that ascorbic acid (AA) was capable of reducing Hg(II) to Hg(I) but not Cu(II), leading to the recovery of fluorescence of Hg(II)–CDs mixture by freeing the CDs, while Cu(II)–CDs mixture remained in the quenched state.

Figure 4: 
Simultaneously sensing of Cu(II) and Hg(II) by fluorescence ON–OFF–ON process with the addition of restoration reagents. Reprinted from Zhao et al.’s work with the kind permission of The Royal Society of Chemistry [108]. Copyright (2019) The Royal Society of Chemistry.
Figure 4:

Simultaneously sensing of Cu(II) and Hg(II) by fluorescence ON–OFF–ON process with the addition of restoration reagents. Reprinted from Zhao et al.’s work with the kind permission of The Royal Society of Chemistry [108]. Copyright (2019) The Royal Society of Chemistry.

6.2.5 Detection of multi-HMs under different conditions

Also, multiple HMs ions detection could be realized under different solution conditions such as pH values and buffer components. The CDs prepared by Dong et al. [109] displayed both blue (∼410 nm) and yellow (∼565 nm) emission peaks. By changing the pH of the sensing solution, CDs were successfully applied for detecting Hg(II) at pH 2.5 and detecting Cr(VI) at pH 3, 7, and 12, respectively. Similarly, Liu et al. [110] synthesized dual emission CDs for determination of Fe(III) ions in acidic environments (pH 5.4) based on the obvious fluorescence quenching around 655 nm, and for determination of Zn(II) ions in alkaline environments (pH 9.4) based on the fluorescence quenching around 470 nm. Wang et al. [111] found that many HM ions including Al(III), Hg(II), Pb(II), Ni(II), Cu(II), Cr(III), Co(II), and Fe(II) could quench the fluorescence of prepared CDs in HNO3 solution at pH 5.0. However, the high selectivity to detect Fe(III) could be switched at the same pH in the NaAc–HAc buffer, to detect Pb(II) in the Britton–Robinson buffer, and to detect Hg(II) in PBS solution. The reason for the undistinguished quenching in HNO3 solution was attributed to the nonspecific interactions between the carboxylate groups on the surface of CDs and the HM ions, while high selectivity of CDs to Fe(III), Pb(II), and Hg(II) at each specific buffer solution is ascribed to masking effects of the buffer solutions.

6.2.6 Combination with other techniques

To discriminate multi-MHs in CD-based sensors, other detection techniques like colorimetric changes, UV–VIS absorbance spectrophotometry also could be combined with fluorescence detection [112]. Naccache et al. [113] reported the use of a dual absorbing/fluorescing CD system to detect four HM ions of Co(II), Fe(III), Hg(II), and Pb(II). In fluorescence assays, Co(II), Fe(III), Hg(II), and Pb(II) could be detected by ratiometric analysis based on the difference in the response of the blue and red fluorescence bands. By analyzing changes of the absorbance spectra of the CDs after the addition of metal ions, discrimination of four target metal ions in solution could be achieved as the absorbance spectra of the CDs changed in a specific way for each of the HM ions.

The applications on multi-HMs detection of CD-based fluorescence sensors are listed in Table 5.

Table 5:

Summary of CD-based sensors for multi-HMs detection.

HM Precursors Detection strategies LODs Sample Reference
Cr(VI)

Hg(II)
Jackfruit Strategy 1 10 nM

8 nM
NA [98]
Fe(III)

Hg(II)
Candle soot Strategy 1 10 nM

50 nM
NA [97]
Zn(II),

Mn(II), Cu(II)
Glutathione, formamide Strategy 2 9.64 nM Water [103]
Fe(III), Cu(II) Phenylenediamine ethanol, ammonia Strategy 2 0.19 μM

0.12 μM
NA [100]
Fe(III),

Bi(III),

Al(III)
AA, alcohol Strategy 2 200 nM,

150 nM,

100 nM
Drugs [102]
Cu(II),

Al(III)
Red tea Strategy 2 0.1 μM

0.5 μM
NA [104]
In(III),

Pd(II)
Amino benzoic acid, p-Phenylenediamine Strategy 2 300 nM

60 nM
Cell imaging [108]
Pb(II)

Hg(II)
Bamboo leaves Strategy 3 0.14 nM

0.22 nM
NA [105]
Hg(II),

Fe(III),

Cu(II)
Ammonium citrate, citric acid, 1,10-phenanthroline, EDTA + thiourea Strategy 3 3 nM,

0.5 nM,

30 nM
Water [106]
Hg(II)

Cu(II)
Urea, ethylenediaminetetraacetic (EDTA) Strategy 4 6.2 nM

2.3 nM
NA [108]
Hg(II)

Cr(VI)
Ortho-phenylenediamine, ethyl alcohol Strategy 5 60 nM

260 nM
Water [109]
Fe(III),

Pb(II),

Hg(II)
Citric acid Strategy 5 2.8 μM,

7.2 μM,

5.7 μM
Water [111]
Fe(III)

Zn(II)
Glutathione Strategy 5 0.8 μM

1.2 μM
Serum [110]
Co(II),

Fe(III),

Hg(II),

Pb(II)
Glutathione, formamide Strategy 6 96.8 nM

61.7 nM

39.5 nM,

37.1 nM
NA [113]
  1. Strategy 1: Indistinguishable detection; Strategy 2: Discrimination by CDs emission band; Strategy 3: Differentiation by CDs contained multifluorophores system; Strategy 4: Discrimination by restoration reagent; Strategy 5: Detection multi-HMs under different conditions; Strategy 6: Combination with other techniques; ND, not discussed; NA, not analyzed; HM, heavy metal.

6.3 Speciation of HMs

Speciation of HMs is often more important than the determination of total HMs considering their toxicity and bioavailability. However, although there has been tremendous progress in developing fluorescent probes to detect the total concentration of HMs, few publications have reported on the discrimination of different species of a particular HM. Most of these previous publications reported on the speciation of Cr(III) and Cr(VI) by CD-based fluorescence methods, probably because the chemical behavior of Cr(VI) and Cr(III) exhibits relatively large differences in solution [57]. Li et al. [114] developed a simple approach based on calcination treatment of diethylenetriaminepentaacetic acid to obtain CDs with a high QY of approximately 53.7%. The fluorescence of CDs could be quickly and efficiently quenched by Cr(VI) rather than by Cr(III) based on an IFE. Su et al. [115] prepared CDs by using graphene oxide powder through hydrothermal methods and found that Cr(III) could quench the CDs because of electrostatic adsorption of Cr(III) on graphene oxide derivatives (GQDs) and the strong chelation between Cr(III) and the –COOH and –OH groups on the surface of GQDs. As AA and acid phosphatase could reduce Cr(VI) to Cr(III), the indirect detection of AA and acid phosphatase was proposed with the addition of Cr(VI). In Yang et al.’s approach [116], Cr(VI) was detected based on the selective quenching effect, and concentrations of Cr(III) were calculated by subtraction of Cr(VI) concentration from the total Cr concentration, which was detected after the addition of KMnO4 to oxidize all possible Cr(III) in the sample solution to Cr(VI). More publications on the speciation of Cr are summarized in Table 6.

Table 6:

Typical applications of CD-based fluorescence sensor for speciation of Cr(VI) and Cr(III).

Species Precursors Synthesis LODs Sample References
Cr(VI) Citric acid, ammonia Hydrothermal 40 nM Water [57]
Cr(VI) Diethylenetriaminepentaacetic acid Calcination 0.15 μM Water, serum [114]
Cr(VI) Citric acid, glutamic acid Pyrolysis 0.3 μM Water, serum [168]
Cr(VI) 1-(2-Pyridylazo)-2-naphthol (PAN), cobalt chloride Solvothermal 1.17 μM Water and seafood [169]
Cr(VI) Citric acid, thiourea Hydrothermal 0.2 μM River water [170]
Cr(VI) Cellulose Hydrothermal 0.72 mg/L Water, soil [171]
Cr(VI) Wool and pig hair Hydrothermal 29.6 nM Water [58]
Cr(VI) Melamine Calcination 9.6 nM NA [172]
Cr(VI) Citric acid, thiosemicarbazide Hydrothermal 0.33 nM NA [173]
Cr(VI) Citric acid, glycine Calcination 4.16 μM Water [48]
Cr(VI) Citric acid, reduced glutathione Pyrolysis 0.03 μM Water, soil [174]
Cr(VI) Glucosamine, ethylenediamine Hydrothermal 0.08 μM Urine [175]
Cr(VI) Shrimp shells Carbonization 0.1 µM Water [176]
Cr(III) Graphene oxide powder Hydrothermal 8.9 µM serum [115]
Cr(VI) PVP K90 Hydrothermal 91 nM Water [177]
Cr(VI) Luffa sponge Hemical oxidation NA NA [178]
Cr(VI) Glucose, phosphotungstic acid Hydrothermal 0.16 μM NA [179]
Cr(VI) Apple peels Hydrothermal 0.73 μM Water [180]
Cr(VI) Citric acid, ethylenediamine Hydrothermal 0.14 μM Industrial effluents [181]
Cr(VI) l-ascorbic acid, ethylene diamine Hydrothermal 2.598 nM NA [182]
Cr(VI) 1,2-Ethylenediamine, concentrated H3PO4 Stirring 23 nM Fruit juices [183]
Cr(VI) Kelp Hydrothermal 0.52 μM Water [116]
Cr(VI) Ammonium citrate Hydrothermal 0.01 μmoM NA [184]
  1. ND, not discussed; NA, not analyzed.

Beside Cr, there have been very few reports on the speciation of other HMs. Li et al. [117] used lanthanide coordination polymers doped CDs (CDs@AMP-Tb) to detect Fe(II) and Fe(III) species based on the experimental phenomenon that CDs could be quenched explicitly by Fe(II), and Fe(III) could be detected after being reduced to Fe(II) with the help of a reducing agent of NH2OH·HCl. Recently, Costas-Mora et al. [118] reported in situ ultrasound-assisted syntheses of CDs as optical nanoprobes to selective recognize methylmercury species (Figure 5). The mixture of d-fructose, PEG, NaOH, and EtOH was subjected to ultrasound irradiation for 60 s to obtain the target CDs. The selectivity of the assay was attributed to the different abilities of Hg(II) and MeHg(I) species to interact with CDs encapsulated in PEG. CH3Hg(I) with hydrophobic property could easily cross the PEG coating, come into contact with CDs and lead to fluorescence quenching, while Hg(II), a hydrophilic Hg species, could not interact in this way with CDs, and no fluorescence quenching was observed.

Figure 5: 
Schematic mechanism of selective speciation of MeHg(I) in presence of Hg(II). Reprinted from Bendicho et al.’s work with the kind permission of the American Chemical Society [118]. Copyright (2014) American Chemical Society.
Figure 5:

Schematic mechanism of selective speciation of MeHg(I) in presence of Hg(II). Reprinted from Bendicho et al.’s work with the kind permission of the American Chemical Society [118]. Copyright (2014) American Chemical Society.

7 Gel/solid-state sensor platforms

Although most of the CD-based fluorescence sensors detect HMs in solution, some researchers have made significant advances in developing gel-based or solid-state sensor platforms.

7.1 Paper-based devices

Inspired by the classical success of pH test paper, the fluorescent test papers have been explored by assembling or printing the fluorescent probes onto a piece of paper-based substrates in the visual assays to detect HMs. The filter paper, anion exchange filter paper, cation exchange filter paper, silica gel filter paper, and cellulose nanofibrils paper are potential candidates to be tested with optimal paper surface. The observation of the variations of fluorescence brightness and color with test samples could be achieved by the naked eyes with or without the aid of a simple ultraviolet (UV) lamp. The colorimetric signals could be further collected using a smartphone, which is a portable technology found everywhere today. Like pH test paper, fluorescent test papers also have advantages such as low cost, convenience, easy operation, fast response, and portability, which is especially suitable for on-site detection applications in environmental, medical, and food testing. Meanwhile, some limitations also exist: the construction of paper-based devices is often complicated; the stability and sensitivity of such solid-phase fluorescence detection devices are usually lower than that of liquid-phase detection; most of the reports can only successfully observation of the variations of fluorescence brightness [119] or color changes in the narrow range [84]; usually only semiquantitative of target analytes could be achieved by naked eyes and hence it is not easy to tell the slight changes of brightness and color variations caused by a small dosage of analytes. The quantitative analysis needs the aid of a camera to collect images and software such as Image J to extract signal intensities from the images for further calculation. In this review, some typical applications of CD-based paper devices for HMs detection and efforts to overcome some limitations are discussed.

Lu et al. [120] printed N-doped CDs to fabricate portable paper strips as sensitive fluorescent probes for Cr(VI). Yan et al. [121] hybridized CDs and Au NCs for printing on cellulose acetate circular filter paper to detect Hg(II), which produced a distinct fluorescence color evolution from pink to blue under UV lamp excitation. The cyan CDs and red GSH/DTT-QDs were mixed and printed on filter paper by Zhou et al. [122] as As(III) detection sensors, and the ratio of 1:5 of emission intensity of cyan to red was confirmed to achieve the highest dosage-sensitive ability for visual quantification of As(III) (Figure 6). The detection was based on the aggregation-induced quenching mechanism. The red fluorescence of GSH/DTT-QDs was effectively quenched with As(III), whereas the cyan fluorescence was not, resulting in the change of ratiometric fluorescence and a wide range of color variations in response to the dosage of As(III), thus exhibiting a similar performance like pH test paper.

Figure 6: 
Schematic illustration of fluorescence paper strips for the determination of As(III) with a similar performance like pH test paper. (A) Visualization mechanism of As(III) using glutathione (GSH)/DTT-QDs as fluorescent sensory probe and carbon dots (CDs) as an internal standard probe; (B, C) transmission electron microscopy (TEM) images of GSH/DTT-QDs before and after the addition of As(III); (D) Fluorescent spectra of mixing glutathione (GSH)/DTT-QDs/CDs (20/10 μL in 1.5 mL of Tris HCl buffer, pH 7.4) with the addition of As(III). Reprinted from Zhang et al.’s work with the kind permission of the American Chemical Society [122]. Copyright (2016) American Chemical Society.
Figure 6:

Schematic illustration of fluorescence paper strips for the determination of As(III) with a similar performance like pH test paper. (A) Visualization mechanism of As(III) using glutathione (GSH)/DTT-QDs as fluorescent sensory probe and carbon dots (CDs) as an internal standard probe; (B, C) transmission electron microscopy (TEM) images of GSH/DTT-QDs before and after the addition of As(III); (D) Fluorescent spectra of mixing glutathione (GSH)/DTT-QDs/CDs (20/10 μL in 1.5 mL of Tris HCl buffer, pH 7.4) with the addition of As(III). Reprinted from Zhang et al.’s work with the kind permission of the American Chemical Society [122]. Copyright (2016) American Chemical Society.

Gogoi and Patir [108] developed a filter paper–based microfluidic device loaded with nitrogen-doped CDs (NCDs), vitamin C(Vit C), and trisodium citrate (SC) to simultaneously detect multiple metal ions of Hg(II) and Cu(II) ions. As shown in Figure 7, three channels (a, b, c) were dispersed with NCDs, NCDs plus Vit C, and NCDs plus SC, respectively. After drying at room temperature, the filter paper–based microfluidic device was used as a sensing platform under 365 nm UV light. When Hg(II) only was added in the reaction zone, the regions a and c showed no fluorescence, while region b showed fluorescence restoration. When Cu(II) only was added, the regions a and b showed no fluorescence, while region c showed fluorescence restoration. If Hg(II) and Cu(II) both existed, just the fluorescence of region a was quenched, while the fluorescence signals of both regions b and c were recovered. The concentrations of Hg(II) and Cu(II) ions could be detected by calculating the fluorescence intensity in the images captured by a digital camera using the ImageJ software.

Figure 7: 
Schematic illustration of filter paper–based microfluidic device to parallel detect multiple metal ions of Hg(II) and Cu(II) ions. Reprinted from Zhao et al.’s work with the kind permission of The Royal Society of Chemistry [108]. Copyright (2019) The Royal Society of Chemistry.
Figure 7:

Schematic illustration of filter paper–based microfluidic device to parallel detect multiple metal ions of Hg(II) and Cu(II) ions. Reprinted from Zhao et al.’s work with the kind permission of The Royal Society of Chemistry [108]. Copyright (2019) The Royal Society of Chemistry.

Xu et al. [123] integrated N-CD paper chip, smartphone detection setup, and software application of the Android system together and developed a portable platform for selective detection of Hg(II). In this platform, nitrogen-doped CDs (N-CDs) were immobilized on the filter paper chip for sensing Hg(II). The smartphone detection setup was 3D printed and composed of three components: a smartphone case to fix the smartphone, a black box with two ultraviolet LED light sources (350–370 nm), a rechargeable battery, a power switch, and a smartphone camera. A self-developed software application was applied for the analysis of the fluorescence of N-CDs decorated paper chip. As a sensing platform, the detection limit of 10.7 nM for Hg(II) detection was achieved based on the quenching effect. A paper-based fluorescence sensor with instrument-free signal quantitation was developed for selective detection of Hg(II) by Henry et al. [124]. The quantitation results were obtained by calculating the quenching distance on the filter paper after the addition of Hg(II) by simple measurement with a ruler. With heating preconcentration, the detection limit for this distance-based paper strip with naked-eye detection of Hg(II) was lowered to 5 ug/L, which was sufficient for monitoring Hg(II) in drinking water.

Xu et al. [125] developed a novel microfluidic paper analytical devices (μPADs) by in situ CDs and AuNPs sequential patterning techniques (Figure 8). Combining the two reading regions of CDs and AuNPs, this μPADs provide reliable measurement for serum Fe(III) and serum ferritin simultaneously in whole blood. Fe(III) was detected in the CDs reading region by quenching effect, and ferritin was measured in AuNPs by ELISA reaction. This technique provides a possibility of an iron health test in real clinic applications for blood analysis.

Figure 8: 
Schematic illustration of microfluidic paper analytical devices (μPADs) for the measurement of serum iron and serum ferritin in whole blood. Reprinted from Xu et al.’s work with the kind permission of the American Chemical Society [125]. Copyright (2016) American Chemical Society.
Figure 8:

Schematic illustration of microfluidic paper analytical devices (μPADs) for the measurement of serum iron and serum ferritin in whole blood. Reprinted from Xu et al.’s work with the kind permission of the American Chemical Society [125]. Copyright (2016) American Chemical Society.

7.2 Sensor arrays

A fluorescent sensor array is a differential sensing platform for analytes (usually qualitative detection) in a complex environment. Fluorescence array–based sensing systems are fast, portable, and inexpensive, promising in chemical and biological sensing research. Unlike conventional ”lock and key” type-specific sensor systems, sensor arrays usually contain a series of nonselective sensors with different chemical properties, and target analytes can be differentiated and recognized by analysis of their nonspecific responses to each sensor element. These different responses generate fingerprint patterns, which are later analyzed by statistical methods to recognize the patterns, classify the data, and detect unknown samples. Input data can be fluorescence intensities or the color indices of R, G, and B. To construct a sensor array, a set of sensors should be firstly designed and prepared. Compared to other fluorescence sensors such as metal oxides, semiconductor nanocrystals, and fluorescent dyes, CDs are more attractive materials to be used in sensor arrays because of their simple synthesis procedures. However, as a set of sensors needs to be synthesized, the whole preparation of the sensor array is relatively time-consuming. To date, few publications have reported CD-based sensor arrays for differentiating metal ions.

Liu et al. [126] fabricated seven kinds of CDs (CD1, CD2, CD3, CD4, CD5, CD6, and CD7) for the differentiation of six kinds of HMs (Ag(I), Cd(II), Cr(II), Fe(III), Hg(II), and Pb(II)). Using the principal component analysis method, the hexasensor array could be simplified into a binary sensor array with CD3 and CD5, which already could differentiate six HM ions clearly with established fingerprints. This binary sensor array was successfully applied to identify unknown metal ions in fetal calf serum solutions and local tap water solutions. Also, Chen et al. [127] synthesized three kinds of N doped CDs with various surface states of different amino acids and employed as-prepared CDs in sensor arrays to distinguish six metal ions based on their diverse quenching responses to six metal ions. The discrimination accuracy of unknown samples was found to be 100% at concentrations as low as 10 μM. Furthermore, this sensor array could well recognize the relative complex binary and ternary mixtures of three metal ions.

7.3 Hydrogels or polymers films

Hydrogels or polymer films possess 3-dimensional matrices with porous nature, which allow small molecules and ions to penetrate the hydrogels via diffusion. CDs with sizes comparable or larger than the pore sizes of the hydrogels can be entrapped in the polymer networks. The use of hydrogels incorporated with functionalized fluorescent CDs as sensing systems could overcome the disadvantages of solid printing platforms such as particle aggregation, uneven analyte adsorption, and the possible problem of liquid platforms of shorter shelf life. Therefore, it is considered as potentially a good sensor platform.

Cheng et al. [128] embedded synthesized PEI-modified CDs into the microcrystalline cellulose (MCC) matrix to prepare a fluorescent smart hydrogel. PEI-CDs was found to have strong hydrogen bonding with MCC with rich oxygen- or nitrogen functional groups. In the prepared hydrogels, PEI-CDs acted as the fluorescent sensor to selectively detect Fe(III), and MCC as the gel matrix could eliminate the self-quenching of the fluorescent sensor. Truskewycz et al. [129] prepared a novel hydrogel composite of CDs/PVP/ZnO as an effective Cr(VI) sensing platform with a LOD of 1.2 μM. This hydrogel composite was cheap to manufacture and could be injected into a 96 well plate for high-throughput environmental diagnostic applications. Wang et al. [130] dispersed prepared N-doped CDs and Rhodamine B covalent organic frameworks (NCDs-RhB@COF) nanocomposites into agarose gels to make hydrogel sensor films that could be used for semiquantitative visual detection of Hg(II) under a UV lamp.

Wang et al. [131] attempted to impregnate the prepared yellow-emitting CDs into polymer films as y-CDs/polymer composite Cu(II) sensors to overcome the aggregation-induced instabilities of CDs in sensing performance. They compared the performance of two kinds of polymer films: (a) carboxymethylcellulose/polyvinylalcohol and (b) chitosan. Film (b) was found to be more sensitive than film (a) and thus was thus more suitable as a Cu(II) sensor. The detection limit of the sensor was 10 nM (i.e. 1.3 ppm). Zhang et al. [132] grafted CDs onto a poly(vinylidene fluoride) (PVDF)-g-PAA composite membrane through amidation reaction to prepare a PVDF-g-PAA-CDs composite sensor membrane with stable fluorescence and found that Cu(II), Hg(II), and Fe(III) could quench the fluorescence.

7.4 Ion-imprinted fluorescence polymers

Ion-imprinted polymers (IIPs) are an important branch of molecularly imprinted polymers, which can specifically recognize template ions. Ion imprinted strategy is a desirable and effective method to increase the selectivity to detect analytes in complex matrices. Combining the high selectivity of IIPs with the high sensitivity of CD-based fluorescence sensing is a promising approach but have seldom been used for the determination of targets HMs. The whole procedure to fabricate CD-based IIPs includes the synthesis of CDs, imprinted polymer preparation, and template elution, which is relatively time-consuming and complicated. Usually, one kind of ion-imprinted fluorescence polymers can only detect one target ions. Multitemplate imprinting for multi-HMs is a good strategy for high-throughput detection of HMs.

Liu et al. [101] fabricated Cr(III) and Pb(II) imprinted fluorescent sensors by a simple one-pot sol-gel method using blue and red dual emission CDs for signal output. CDs were first synthesized using a one-step hydrothermal method using citric acid and ethylenediamine as precursors in formamide–water (V:V = 1:3) binary systems. The prepared CDs were then mixed with Cr(III) and Pb(II) ions together with tetraethylorthosilicate and cetyltrimethylammonium bromide (CTAB) in alkaline condition and reacted at 70 °C for 3 h to afford white solid polymer particles with fluorescence. To remove template ions and CTAB micelles, EDTA and ethanol were used as the eluting solvent. Compared with nonimprinted polymers, the quenching efficiency of imprinted polymers was much higher, and the imprinting factors, defined as the ratio of the quenching constants, were calculated to be 6.7 and 4.5 for Cr(III) and Pb(II) detection, respectively. In Lu et al.’s work [133], CD-based IIPs with mesoporous structured dual-templates and dual-reference ratio peaks by one-pot synthesis method for simultaneous fluorescence detection of Fe(III) and Cu(II) were developed (Figure 9). The prepared IIPs had fluorophores of CDs at 420 nm and fluorophores of CdTe QDs at 620 nm. In the presence of Fe(II), CDs were quenched while CdTe QDs were used as a reference. For the detection of Cu(II), CdTe QDs were used for the response signal, while CDs were used as a reference. As a result, the dual-reference IIPs probe could simultaneously detect Cu(II) and Fe(III), while maintaining sensitivity and selectivity with enhanced detection efficiency. Similarly, the same group developed dual-template IIPs for simultaneous detection of Ag(I) and Pb(II) using CDs and AuNCs as fluorescence signal output [134]. Table 7 lists a summary of solid-state CDs-based fluorescence sensing platforms.

Figure 9: 
The fluorescence spectra of the ion-imprinted (A) and nonimprinted (C) fluorescence polymers under exposure to different concentrations of Cu(II). (B) and (D) are the plots of (I
420)0/I
620 as a function of the Cu(II) concentration. (E) shows the photos of IIP detection solutions under 365 nm UV light, while (F) shows the photos of NIP detection solutions under 365 nm UV light. (G) Schematic illustration of the preparation of dual template ion-imprinted dual reference ratiometric fluorescent (RF) probes to detect Fe(III) and Cu(II). Reprinted from Xu et al.’s work with the kind permission of The Royal Society of Chemistry [133]. Copyright (2019) The Royal Society of Chemistry.
Figure 9:

The fluorescence spectra of the ion-imprinted (A) and nonimprinted (C) fluorescence polymers under exposure to different concentrations of Cu(II). (B) and (D) are the plots of (I 420)0/I 620 as a function of the Cu(II) concentration. (E) shows the photos of IIP detection solutions under 365 nm UV light, while (F) shows the photos of NIP detection solutions under 365 nm UV light. (G) Schematic illustration of the preparation of dual template ion-imprinted dual reference ratiometric fluorescent (RF) probes to detect Fe(III) and Cu(II). Reprinted from Xu et al.’s work with the kind permission of The Royal Society of Chemistry [133]. Copyright (2019) The Royal Society of Chemistry.

8 Cell and other bioimaging

Due to the advantages of small size (less than 10-nm), good biocompatibility, high stability under salt stress, negligible photobleaching properties, good hydrophilicity, and long-term homogeneous aqueous phase stability, CDs are good candidates for cell imaging with excellent cell permeability and negligible toxicity in living cells. They can easily enter into different cell types such as cell lines, bacterias, fungi, plants and animals, and further be localized in different intracellular compartments including cell membrane, cytoplasm, mitochondria, endosomes, and lysosomes; the modified or conjugated functional groups on the surface of CDs could be specifically designed for sensing or imaging purpose. Until now, CDs have been wildly explored as fluorescent imaging agents to capture images of cancer cells, bacterial, fungal cells, plants, and animals, including HMs sensing images. In these applications, the CDs with blue or green emissions may face the challenges of interference from autofluorescence at the short wavelengths of biological matrices and the potential photodamage of ultraviolet excitation light to biological tissues. CDs with long-wavelength emissions, especially in far‐red to NIR regions are more attractive with advantages of little damage to the biological matrix, strong tissue penetration ability, and weak light scattering effects [18]. To explore the potential applications of CDs as HMs imaging sensors in living things, generally, MTT assay was first investigated to assess the cytotoxicity of the as-prepared CDs. The CDs with proved low cytotoxicity were further used as imaging sensing reagents in cell lines, plants, or animals.

Song et al. [135] prepared a sulfur and nitrogen dual-doped CDs (S, N-CDs) by an acid–base neutralization and exothermic carbonization method by mixing of glucose, concentrated H2SO4, and 1,2-ethylenediamine (EDA). The MTT assay showed their naturally low cytotoxicity to SMMC 7721 cell lines with the possibility of multicolor biolabeling reagent under photoexcitation at 543, 488, 405, and 515 nm. When SMMC 7721 cells were incubated with the addition of Cr(VI), the fluorescence intensity inside cells decreased with increasing concentration of Cr(VI) and could be further restored by AA. Asthana et al. [136] reported the use of microalgal biomass-derived CDs as a fluorescent turn-off sensor for endogenous imaging of Hg (II) and Cr (VI) contaminated live cells. Ottoor et al. [53] proved that cystamine capped CDs could be used as imaging reagent to image MCF-7 breast cancer cells, as well as Staphylococcus aureus and Pseudomonas aeruginosa bacterial cells but did not test the imaging capacity to target Cr(VI).

Table 7:

Summary of solid-state CDs-based fluorescence sensing platforms.

HM Sensor platform LODs Sample Reference
Fe(III) Hydrogel 65 nM Water [128]
Hg(II) Smartphone-assisted hydrogels 1.41 mg/L Water [185]
Cu(II) Polymer film 10 nM NA [131]
Cu(II),

Fe(III),

Hg(II)
PVDF-g-PAA film 1 nM

1 μM

1 nM
NA [132]
Hg(II) Agarose hydrogel film 15.9 nM NA [130]
Cu(II) Filter paper 8.82 nM Water [85]
Cr(VI) Paper strips with novel origami designs 0.14 mM Water [120]
Cu(II)

Hg(II)
Filter paper–based microfluidic device 6.2 nM, 2.304 nM NA [108]
As(III) Filter paper 5 μg/L Water [122]
Hg(II) Filter paper 0.14 nM Water [84]
Ag(I) Filter papers 2 nM NA [87]
Fe(III) Filter paper IS Blood [125]
Fe(III) Cellulose nanofibrils 10 μM NA [119]
Cu(II) Filter paper 7.31 nM Water [186]
Cu(II) Filter paper 2 ppb Water [158]
Hg(II) Filter paper 7.63 nM NA [155]
As(III) Filter paper 0.5 ppb Water [86]
Hg(II) Distance-based paper device 5 μg/L Water [124]
Hg(II) Cellulose acetate circular filter paper 28 nM Water [121]
Hg(II) Smartphone-assisted paper-based analytical device 10.7 nM Water [123]
Hg(II) Distance-based paper strips 6 μg/L Water [124]
Cr(III)

Pb(II)
Ion imprinted fluorescence polymers 27 nM,

34 nM
Water [101]
Fe(III)

Cu(II)
Ion imprinted fluorescence polymers 340 nM

130 nM
Water [133]
Ag(I), Cd(II), Cr(II), Fe(III), Hg(II), and Pb(II) Sensor array Qualitative detection Serum

Water
[126]
Cr(III), Cu(II), Eu(III), Fe(III), Hg(II), Pb(II) Sensor array Qualitative detection NA [127]
  1. ND, not discussed; NA, not analyzed; PVDF, poly(vinylidene fluoride); HM, heavy metal.

Huang et al. [137] used mung bean seeds to grow sprouts for the multicolor bioimaging of Cr(VI) by CDs. The sprouts could grow healthily in fluorescence matrices containing CDs, indicating their low toxicity and excellent biocompatibility. The bean sprout displayed cyan, green and red emissions under the excitation of UV (405 nm), blue (488 nm), and green (546 nm) light, respectively. When Cr(VI) entered the sprouts, the cyan, green, and red emissions were significantly weakened, and the emission could be recovered using AA. Zhang et al. [138] fabricated full-color CDs (from blue to red) for the in vivo bioimaging of zebrafish. The zebrafish images suggest strong fluorescence (red, green, and blue) in the entire zebrafish body, including skeletal structures such as vertebrae and cleithrum. In these in vivo staining articles, the imaging of HMs in the animal was not presented, indicating there may be a long way to go to realize CDs as HM sensors in vivo.

Zhang et al. [139] synthesized La-doped CDs and investigated their multifunctional applications as Hg(II) sensor, cell imaging, and animal imaging reagents, and antibacterial reagents (Figure 10). The synthesized La-CDs were proved to be an effective fluorescent probe for the quenchometric determination of Hg(II) in both aqueous solution and A549 cancer cells. Moreover, the use of La-CQDs as luminescent agents for in vivo animal imaging using female nude mice as a model was successfully demonstrated. In vivo fluorescence images after tail vein injection exhibited bright emissions along with the distribution of La-CQDs in mice, which demonstrated that La-CQDs could serve as a new platform of fluorescent bioimaging and showed promise for drugs delivery. Sun et al. [140] also reported similar work that N, S-codoped red-emitting CDs (R-CDs) could be used as a Fe(III) sensor through an ET quenching process by cell imaging, and showed the possibility as imaging reagent in vivo by subcutaneous injection of CDs into nude mice.

Figure 10: 
Schematic illustration of carbon dots (CDs) used as a bioimaging reagent. Reprinted from Zhou et al.’s work with the kind permission of Elsevier [139]. Copyright (2019) Elsevier.
Figure 10:

Schematic illustration of carbon dots (CDs) used as a bioimaging reagent. Reprinted from Zhou et al.’s work with the kind permission of Elsevier [139]. Copyright (2019) Elsevier.

For completeness, additional applications of CD-based sensors in bioimaging are listed in Table 8.

Table 8:

Typical applications of CDs in HMs bioimaging.

Heavy metal Precursors Synthesis LODs Model cell Reference
Cr(VI) Glucose, H2SO4, 1,2-ethylenediamine Carbonization 76 nM SMMC 7721 [135]
Cr(VI) Aspartic acid, diethylenetriamine, phosphoric acid Hydrothermal 0.48 μM HEK-293 [187]
Cr(VI)

Hg(II)
Algal biomass Hydrothermal 0.018 μM HEK 293  [136]
Cr(VI) Ground nut Hydrothermal o.1 ppm MCF-7 cells [188]
Ag(I) Neutral red (NR), triethylamine Hydrothermal 0.27 μM SMMC7721 cells [82]
Cr(VI) Jeera Hydrothermal 1.57 μM MCF-7 cells staphylococcus aureus, Pseudomonas aeruginosa [53]
Cr(VI) Citric acid, N-acetyl-l-cysteine Microwave 20 nM SiHa cells [189]
Cr(VI) Tetrakis(hydroxymethyl) phosphonium chloride, p-aminobenzenesulfonic acid Hydrothermal 0.23 μM HeLa cells

MCF-7 cells
[190]
Cr(VI) 2-Hydroxyphenylboronic acid, ethylenediamine Hydrothermal 0.5 μM HeLa cells [25]
Ag(I) 3-Aminobenzeneboronic acid, 2,5-diaminobenzenesulfonic acid Hydrothermal 0.35 μM HeLa cells, MCF-7 cells [191]
Hg(II) Citric acid, acrylamide, formamide Solvothermal 0.19 μM Fungi [192]
Cu(II) Leaf extract Microwave 50 nM E. coli, P. vulgaris [37]
Cr(VI) Urea, Maleic anhydride Microwave 0.12 μM Mung bean sprouts [137]
Hg(II) Adenosine disodium triphosphate, LaCl3·2H2O Hydrothermal 0.1 µM HepG2, A549 nude mice [139]
Fe(III) 2,5-diaminobenzenesulfonic acid Hydrothermal 0.27 μM Hep G2 cell nude mice [140]
Co(II) o-phenylenediamine, urea Hydrothermal 52 nM Zebrafish [193]
Fe(III) 1,5-diaminonaphthalene,

1,8-diaminonaphthalene
Solvothermal NA Zebrafish [138]
  1. ND, not discussed; NA, not analyzed; HM, heavy metal.

9 Conclusion and perspectives

CD-based fluorescence sensors are excellent tools for sensing or imaging HMs of interest because of their key favorable features over fluorescent dyes or noble metal/metal oxide nanomaterials. Many CD-based fluorescence sensors have been reported showing excellent performance to detect HMs especially based on “turn-on” and ratiometric measurement. Increasing efforts are made in areas in multi-HMs detection to enhance the detection throughput, speciations of HMs for concerning high toxic species, fabrication of gel/solid sensor platforms such as paper-based devices, sensor arrays, hydrogel or polymer films, and IIPs for cost-effective detection, smart visual detection, and on-site analysis, and in vitro or in vivo HMs bioimaging.

Overall, a great deal of progress has already been made on CD-based sensing systems to detect HMs. Nevertheless, several challenges remain to be overcome in the future, and emerging trends may be dictated accordingly:

  1. The problem of low product yield with difficulty in purifications is still not well solved, limiting the large-scale production of CDs as HMs sensors. More efforts should be made to increase product yield and explore fast, reproducible, and easy-to-operate purification procedures.

  2. The controllable preparation of CDs with uniformity and desirable size is still deficient, resulting in the complexity of their composition and unclear origin of their PL emission. This may significantly influence the optical feature and activity of CDs as HM sensors and hinder the final commercialization and application in the market. Also, the existing technology and means cannot identify the accurate molecular structure of CDs after polymerization and carbonization and clearly define the HMs-CDs coordinated complexes. More works are required to sufficiently characterize the prepared CDs and further improve the preparation methods to obtain size-controlled CDs.

  3. The overwhelming reports show great performance in the detection of single HM ions. However, the research progress on CDs for sensing of multi-HMs and different species of a particular HM is still in its infancy, leaving a vast open space for future developments.

  4. CD-based test paper strips or arrays as practical kits for HMs sensing purposes have excellent potential for route quality control and monitor HMs, especially in environmental applications. Advances in the fabrication of simple, stable, and dosage-sensitive paper strips like pH test paper with high performance for semiquantitative estimation of the concentrations of HMs are expected in the future. Accordingly, more attention could be paid to developing portable, inexpensive devices for on-site detection.

  5. Until now, most of the fluorescent CD-based sensors for HMs are applied to real sample analysis with a relatively simple matrix-like water. For most environmental and biological samples, the matrices are much more complicated with various ions and organic compounds, which can cause damage or performance deterioration to fluorescent sensors. The design of stable and reliable CD-based sensors to detect HMs in complex matrices is still a big challenge.

  6. Although some reports demonstrated the successful applications of CDs in cell sensing and imaging of HMs, there is still a long way in real biomedical applications. To improve tissue penetration and to reduce autofluorescence, the synthesis of CD with fluorescence in the NIR-I (650–1000 nm) and NIR-II (1000–1700 nm) ranges will be a promising but challenging research direction. To increase the selectivity for HMs sensing in vivo, CDs modified with small oligonucleotide ligands (aptamer) and enzymes with specific and strong affinities to target HMs will be promising approaches.

In summary, although a long-term effort is still needed to overcome the challenges mentioned above, CDs have demonstrated great potential to detect HMs and their species in sensing and bioimaging applications.


Corresponding authors: Pingjing Li, Analytical Instrumentation Center, Institute of Deep-sea Science and Engineering, Chinese Academy of Sciences, Sanya 572000, China; and Instrumentation Center in Deep-sea Science and Engineering, Science & Technology Tower of Sanya Yazhou Bay, Sanya 572000, China, E-mail: ; and Sam F. Y. Li, Department of Chemistry, 3 Science Drive 3, National University of Singapore, Singapore 117543, Singapore; NUS Environmental Research Institute, National University of Singapore, T-Lab Building, 5A Engineering Drive 1, Singapore 117411, Singapore, E-mail:

Funding source: Institute of Deep-sea Science and Engineering, Chinese Academy of Sciences

Award Identifier / Grant number: R-143-000-B48-114

Acknowledgments

P.J. Li would like to acknowledge the support of the One Hundred Person Project of the Chinese Academy of Sciences, China. S. F. Y. Li would like to acknowledge the support of the Ministry of Education, Singapore (R-143-000-B48-114).

  1. Author contribution: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.

  2. Research funding: P.J. Li would like to acknowledge the support of the One Hundred Person Project of the Chinese Academy of Sciences, China. S. F. Y. Li would like to acknowledge the support of the Ministry of Education, Singapore (R-143-000-B48-114).

  3. Conflict of interest statement: The authors declare no conflicts of interest regarding this article.

References

[1] J. J. Kim, Y. S. Kim, and V. Kumar, “Heavy metal toxicity: an update of chelating therapeutic strategies,” J. Trace Elem. Med. Biol., vol. 54, pp. 226–231, 2019, https://doi.org/10.1016/j.jtemb.2019.05.003.Search in Google Scholar PubMed

[2] M. Hemmati, M. Rajabi, and A. Asghari, “Magnetic nanoparticle based solid-phase extraction of heavy metal ions: A review on recent advances,” Microchim. Acta, vol. 185, pp. 1–32, 2018, https://doi.org/10.1007/s00604-018-2670-4.Search in Google Scholar PubMed

[3] S. Clemens and J. F. Ma, “Toxic heavy metal and metalloid accumulation in crop plants and foods,” in Annual Review of Plant Biology, vol. 67, S. S. Merchant, Ed., Version, 2016, pp. 489–512.10.1146/annurev-arplant-043015-112301Search in Google Scholar PubMed

[4] L. N. Suvarapu and S. O. Baek, “Recent developments in the speciation and determination of mercury using various analytical techniques,” J. Anal. Methods Chem., vol. 2015, pp. 1–19, 2015, https://doi.org/10.1155/2015/372459.Search in Google Scholar PubMed PubMed Central

[5] Y. Shen, J. W. Hu, T. T. Liu, H. W. Gao, and Z. J. Hu, “Colorimetric and fluorogenic chemosensors for mercury ion based on nanomaterials,” Prog. Chem., vol. 31, pp. 536–549, 2019.Search in Google Scholar

[6] Y. W. Fen and W. M. M. Yunus, “Surface plasmon resonance spectroscopy as an alternative for sensing heavy metal ions: A review,” Sens. Rev., vol. 33, pp. 305–314, 2013.10.1108/SR-01-2012-604Search in Google Scholar

[7] A. Khanmohammadi, A. J. Ghazizadeh, P. Hashemi, A. Afkhami, F. Arduini, and H. Bagheri, “An overview to electrochemical biosensors and sensors for the detection of environmental contaminants,” J. Iran. Chem. Soc., pp. 1–19, 2020, https://doi.org/10.1007/s13738-020-01940-z.Search in Google Scholar

[8] P. Samanta, S. Let, W. Mandal, S. Dutta, and S. K. Ghosh, “Luminescent metal-organic frameworks (LMOFs) as potential probes for the recognition of cationic water pollutants,” Inorg. Chem. Front., vol. 7, pp. 1801–1821, 2020, https://doi.org/10.1039/d0qi00167h.Search in Google Scholar

[9] T. Rasheed, M. Bilal, F. Nabeel, H. M. N. Iqbal, C. L. Li, and Y. F. Zhou, “Fluorescent sensor based models for the detection of environmentally-related toxic heavy metals,” Sci. Total Environ., vol. 615, pp. 476–485, 2018, https://doi.org/10.1016/j.scitotenv.2017.09.126.Search in Google Scholar PubMed

[10] E. Oliveira, C. Nunez, H. M. Santos, et al.., “Revisiting the use of gold and silver functionalised nanoparticles as colorimetric and fluorometric chemosensors for metal ions,” Sensor. Actuator. B Chem., vol. 212, pp. 297–328, 2015, https://doi.org/10.1016/j.snb.2015.02.026.Search in Google Scholar

[11] P. Kumar, K. H. Kim, V. Bansal, T. Lazarides, and N. Kumar, “Progress in the sensing techniques for heavy metal ions using nanomaterials,” J. Ind. Eng. Chem., vol. 54, pp. 30–43, 2017, https://doi.org/10.1016/j.jiec.2017.06.010.Search in Google Scholar

[12] R. Ahmad, N. Tripathy, A. Khosla, et al., “Review-recent advances in nanostructured graphitic carbon nitride as a sensing material for heavy metal ions,” J. Electrochem. Soc., vol. 167, pp. 1–19, 2019.10.1149/2.0192003JESSearch in Google Scholar

[13] X. Y. Xu, R. Ray, Y. L. Gu, et al.., “Electrophoretic analysis and purification of fluorescent single-walled carbon nanotube fragments,” J. Am. Chem. Soc., vol. 126, pp. 12736–12737, 2004, https://doi.org/10.1021/ja040082h.Search in Google Scholar PubMed

[14] N. Duran, M. B. Simoes, A. C. M. de Moraes, W. J. Favaro, and A. B. Seabra, “Nanobiotechnology of carbon dots: A review,” J. Biomed. Nanotechnol., vol. 12, pp. 1323–1347, 2016, https://doi.org/10.1166/jbn.2016.2225.Search in Google Scholar PubMed

[15] J. Y. Wei, Q. Lou, J. H. Zang, et al.., “Scalable synthesis of green fluorescent carbon dot powders with unprecedented efficiency,” Adv. Opt. Mater., vol. 8, p. 1901938, 2020, https://doi.org/10.1002/adom.201901938.Search in Google Scholar

[16] L. SC, Q. L, K. LK, L. D, and X. SC, “Chemiluminescent carbon dots: synthesis, properties, and applications,” Nano Today, vol. 35, p. 100954, 2020.10.1016/j.nantod.2020.100954Search in Google Scholar

[17] C. L. Shen, G. S. Zheng, M. Y. Wu, et al.., “Chemiluminescent carbon nanodots as sensors for hydrogen peroxide and glucose,” Nanophotonics, vol. 9, pp. 3597–3604, 2020, https://doi.org/10.1515/nanoph-2020-0233.Search in Google Scholar

[18] C. L. Shen, Q. Lou, J. H. Zang, et al.., “Near-infrared chemiluminescent carbon nanodots and their application in reactive oxygen species bioimaging,” Adv. Sci., vol. 7, p. 1903525, 2020, https://doi.org/10.1002/advs.201903525.Search in Google Scholar PubMed PubMed Central

[19] C. L. Shen, Q. Lou, C. F. Lv, et al.., “Bright and multicolor chemiluminescent carbon nanodots for advanced information encryption,” Adv. Sci., vol. 6, p. 1802331, 2019, https://doi.org/10.1002/advs.201802331.Search in Google Scholar PubMed PubMed Central

[20] K. O. Boakye-Yiadom, S. Kesse, Y. Opoku-Damoah, et al.., “Carbon dots: applications in bioimaging and theranostics,” Int. J. Pharm., vol. 564, pp. 308–317, 2019, https://doi.org/10.1016/j.ijpharm.2019.04.055.Search in Google Scholar PubMed

[21] J. J. Qin, B. H. Dong, R. J. Gao, et al.., “Water-soluble silica-coated ZnS:Mn nanoparticles as fluorescent sensors for the detection of ultratrace copper(II) ions in seawater,” Anal. Methods, vol. 9, pp. 322–328, 2017, https://doi.org/10.1039/c6ay02486f.Search in Google Scholar

[22] C. G. Cheng, M. Xing, and Q. L. Wu, “A universal facile synthesis of nitrogen and sulfur co-doped carbon dots from cellulose-based biowaste for fluorescent detection of Fe3+ ions and intracellular bioimaging,” Mater. Sci. Eng. C, vol. 99, pp. 611–619, 2019, https://doi.org/10.1016/j.msec.2019.02.003.Search in Google Scholar PubMed

[23] Q. Xu, M. R. Zhang, Y. Liu, et al.., “Synthesis of multi-functional green fluorescence carbon dots and their applications as a fluorescent probe for Hg2+ detection and zebrafish imaging,” New J. Chem., vol. 42, pp. 10400–10405, 2018, https://doi.org/10.1039/c8nj01639a.Search in Google Scholar

[24] N. K. R. Bogireddy, V. Barba, and V. Agarwal, “Nitrogen-doped graphene oxide dots-based “Turn-OFF” H2O2, Au(III), and “Turn-OFF-ON” Hg(II) sensors as logic gates and molecular keypad locks,” ACS Omega, vol. 4, pp. 10702–10713, 2019, https://doi.org/10.1021/acsomega.9b00858.Search in Google Scholar PubMed PubMed Central

[25] Y. M. Guo, Y. Z. Chen, F. P. Cao, L. J. Wang, Z. Wang, and Y. M. Leng, “Hydrothermal synthesis of nitrogen and boron doped carbon quantum dots with yellow-green emission for sensing Cr(VI), anti-counterfeiting and cell imaging,” RSC Adv., vol. 7, pp. 48386–48393, 2017, https://doi.org/10.1039/c7ra09785a.Search in Google Scholar

[26] C. H. Fan, K. L. Ao, P. F. Lv, et al.., “Fluorescent nitrogen-doped carbon dots via single-step synthesis applied as fluorescent probe for the detection of Fe3+ ions and anti-counterfeiting inks,” Nano, vol. 13, pp. 1–14, 2018, https://doi.org/10.1142/s1793292018500972.Search in Google Scholar

[27] R. S. Li, J. H. Liu, T. Yang, et al.., “Carbon quantum dots-europium(III) energy transfer architecture embedded in electrospun nanofibrous membranes for fingerprint security and document counterspy,” Anal. Chem., vol. 91, pp. 11185–11191, 2019, https://doi.org/10.1021/acs.analchem.9b01936.Search in Google Scholar PubMed

[28] W. B. Zhao, K. K. Liu, S. Y. Song, R. Zhou, and C. X. Shan, “Fluorescent nano-biomass dots: ultrasonic-assisted extraction and their application as nanoprobe for Fe3+ detection,” Nanoscale Res. Lett., vol. 14, p. 130, 2019, https://doi.org/10.1186/s11671-019-2950-x.Search in Google Scholar PubMed PubMed Central

[29] C. L. Shen, J. H. Zang, Q. Lou, et al.., “In-situ embedding of carbon dots in a trisodium citrate crystal matrix for tunable solid-state fluorescence,” Carbon, vol. 136, pp. 359–368, 2018, https://doi.org/10.1016/j.carbon.2018.05.015.Search in Google Scholar

[30] Q. Lou, S. N. Qu, P. T. Jing, et al.., “Water-triggered luminescent “Nano-bombs” based on supra-(carbon nanodots),” Adv. Mater., vol. 27, pp. 1389–1394, 2015, https://doi.org/10.1002/adma.201403635.Search in Google Scholar PubMed

[31] P. L. Zuo, X. H. Lu, Z. G. Sun, Y. H. Guo, and H. He, “A review on syntheses, properties, characterization and bioanalytical applications of fluorescent carbon dots,” Microchim. Acta, vol. 183, pp. 519–542, 2016, https://doi.org/10.1007/s00604-015-1705-3.Search in Google Scholar

[32] J. H. Li and S. L. Zhuang, “Research progress of carbon dots preparation,” Rare Metal Mat. Eng., vol. 48, pp. 3401–3416, 2019.Search in Google Scholar

[33] D. Gao, H. Zhao, X. Chen, and H. Fan, “Recent advance in red-emissive carbon dots and their photoluminescent mechanisms,” Mater. Today Chem., vol. 9, pp. 103–113, 2018, https://doi.org/10.1016/j.mtchem.2018.06.004.Search in Google Scholar

[34] Y. S. Liu, W. Li, P. Wu, and S. X. Liu, “Preparation and applications of carbon quantum dots prepared via hydrothermal carbonization method,” Prog. Chem., vol. 30, pp. 349–364, 2018.Search in Google Scholar

[35] J. Zuo, T. Jiang, X. J. Zhao, X. H. Xiong, S. J. Xiao, and Z. Q. Zhu, “Preparation and application of fluorescent carbon dots,” J. Nanomater., vol. 2015, pp. 1–13, 2015, https://doi.org/10.1155/2015/787862.Search in Google Scholar

[36] T. V. de Medeiros, J. Manioudakis, F. Noun, J. R. Macairan, F. Victoria, and R. Naccache, “Microwave-assisted synthesis of carbon dots and their applications,” J. Mater. Chem. C, vol. 7, pp. 7175–7195, 2019, https://doi.org/10.1039/c9tc01640f.Search in Google Scholar

[37] A. Bhati, S. R. Anand, D. Saini, P. Khare, P. Dubey, and S. K. Sonkar, “Self-doped nontoxic red-emitting Mg-N-embedded carbon dots for imaging, Cu(II) sensing and fluorescent ink,” New J. Chem., vol. 42, pp. 19548–19556, 2018, https://doi.org/10.1039/c8nj04754e.Search in Google Scholar

[38] F. Li, D. Y. Yang, and H. P. Xu, “Non-metal-heteroatom-doped carbon dots: synthesis and properties,” Chem. Eur J., vol. 25, pp. 1165–1176, 2019, https://doi.org/10.1002/chem.201802793.Search in Google Scholar PubMed

[39] F. J. Liu, Y. Jiang, C. Fan, et al.., “Fluorimetric and colorimetric analysis of total iron ions in blood or tap water using nitrogen-doped carbon dots with tunable fluorescence,” New J. Chem., vol. 42, pp. 9676–9683, 2018, https://doi.org/10.1039/c8nj00711j.Search in Google Scholar

[40] X. He, Y. Han, X. L. Luo, et al.., “Terbium (III)-referenced N-doped carbon dots for ratiometric fluorescent sensing of mercury (II) in seafood,” Food Chem., vol. 320, pp. 1–8, 2020, https://doi.org/10.1016/j.foodchem.2020.126624.Search in Google Scholar PubMed

[41] A. Beiraghi and S. A. Najibi-Gehraz, “Purification and fractionation of carbon dots using pH-controlled cloud point extraction technique,” J. Nanostruct., vol. 10, pp. 107–118, 2020.Search in Google Scholar

[42] A. Koutsioukis, A. Akouros, R. Zboril, and V. Georgakilas, “Solid phase extraction for the purification of violet, blue, green and yellow emitting carbon dots,” Nanoscale, vol. 10, pp. 11293–11296, 2018, https://doi.org/10.1039/c8nr03668c.Search in Google Scholar PubMed

[43] S. Sahu, B. Behera, T. K. Maiti, and S. Mohapatra, “Simple one-step synthesis of highly luminescent carbon dots from orange juice: application as excellent bio-imaging agents,” Chem. Commun., vol. 48, pp. 8835–8837, 2012, https://doi.org/10.1039/c2cc33796g.Search in Google Scholar PubMed

[44] A. A. Kokorina, A. V. Sapelkin, G. B. Sukhorukov, and I. Y. Goryachev, “Luminescent carbon nanoparticles separation and purification,” Adv. Colloid Interface Sci., vol. 274, pp. 1–14, 2019, https://doi.org/10.1016/j.cis.2019.102043.Search in Google Scholar PubMed

[45] Z. F. Jiang, M. J. Krysmann, A. Kelarakis, et al.., “Understanding the photoluminescence mechanism of carbon dots,” MRS Adv., vol. 2, pp. 2927–2934, 2017, https://doi.org/10.1557/adv.2017.461.Search in Google Scholar

[46] F. L. Zu, F. Y. Yan, Z. J. Bai, et al.., “The quenching of the fluorescence of carbon dots: A review on mechanisms and applications,” Microchim. Acta, vol. 184, pp. 1899–1914, 2017, https://doi.org/10.1007/s00604-017-2318-9.Search in Google Scholar

[47] M. Zheng, Z. G. Xie, D. Qu, et al.., “On off on fluorescent carbon dot nanosensor for recognition of chromium(VI) and ascorbic acid based on the inner filter effect,” ACS Appl. Mater. Interfaces, vol. 5, pp. 13242–13247, 2013, https://doi.org/10.1021/am4042355.Search in Google Scholar PubMed

[48] H. T. Wang, S. Liu, Y. S. Xie, et al.., “Facile one-step synthesis of highly luminescent N-doped carbon dots as an efficient fluorescent probe for chromium(vi) detection based on the inner filter effect,” New J. Chem., vol. 42, pp. 3729–3735, 2018, https://doi.org/10.1039/c8nj00216a.Search in Google Scholar

[49] Y. Xu, H. Y. Li, B. Wang, et al.., “Microwave-assisted synthesis of carbon dots for “turn-on” fluorometric determination of Hg(II) via aggregation-induced emission,” Microchim. Acta, vol. 185, pp. 1–7, 2018, https://doi.org/10.1007/s00604-018-2781-y.Search in Google Scholar PubMed

[50] C. X. Wang, K. L. Jiang, Z. Z. Xu, H. H. Lin, and C. Zhang, “Glutathione modified carbon-dots: from aggregation-induced emission enhancement properties to a “turn-on” sensing of temperature/Fe3+ ions in cells,” Inorg. Chem. Front., vol. 3, pp. 514–522, 2016, https://doi.org/10.1039/c5qi00273g.Search in Google Scholar

[51] Y. C. Tian, A. Kelarakis, L. Li, et al.., “Facile fluorescence “Turn on” sensing of lead ions in water via carbon nanodots immobilized in spherical polyelectrolyte brushes,” Front. Chem., vol. 6, pp. 1–13, 2018, https://doi.org/10.3389/fchem.2018.00470.Search in Google Scholar PubMed PubMed Central

[52] G. Bjorklund, G. Crisponi, V. M. Nurchi, R. Cappai, A. B. Djordjevic, and J. Aaseth, “A review on coordination properties of thiol-containing chelating agents towards mercury, cadmium, and lead,” Molecules, vol. 24, pp. 1–32, 2019.10.3390/molecules24183247Search in Google Scholar PubMed PubMed Central

[53] V. Roshni, V. Gujar, H. Pathan, et al.., “Bioimaging applications of carbon dots (C. dots) and its cystamine functionalization for the sensitive detection of Cr(VI) in aqueous samples,” J. Fluoresc., vol. 29, pp. 1381–1392, 2019, https://doi.org/10.1007/s10895-019-02448-3.Search in Google Scholar PubMed

[54] H. Kaur, P. Raj, H. Sharma, M. Verma, N. Singh, and N. Kaur, “Highly selective and sensitive fluorescence sensing of nanomolar Zn2+ ions in aqueous medium using Calix 4 arene passivated Carbon Quantum Dots based on fluorescence enhancement: real-time monitoring and intracellular investigation,” Anal. Chim. Acta, vol. 1009, pp. 1–11, 2018, https://doi.org/10.1016/j.aca.2017.12.048.Search in Google Scholar PubMed

[55] M. J. Molaei, “The optical properties and solar energy conversion applications of carbon quantum dots: A review,” Sol. Energy, vol. 196, pp. 549–566, 2020, https://doi.org/10.1016/j.solener.2019.12.036.Search in Google Scholar

[56] K. A. S. Fernando, S. Sahu, Y. M. Liu, et al.., “Carbon quantum dots and applications in photocatalytic energy conversion,” ACS Appl. Mater. Interfaces, vol. 7, pp. 8363–8376, 2015, https://doi.org/10.1021/acsami.5b00448.Search in Google Scholar PubMed

[57] F. Cai, X. D. Liu, S. Liu, H. Liu, and Y. M. Huang, “A simple one-pot synthesis of highly fluorescent nitrogen-doped graphene quantum dots for the detection of Cr(VI) in aqueous media,” RSC Adv., vol. 4, pp. 52016–52022, 2014, https://doi.org/10.1039/c4ra09320h.Search in Google Scholar

[58] S. H. Wang, H. Y. Niu, S. J. He, and Y. Q. Cai, “One-step fabrication of high quantum yield sulfur- and nitrogen-doped carbon dots for sensitive and selective detection of Cr(VI),” RSC Adv., vol. 6, pp. 107717–107722, 2016, https://doi.org/10.1039/c6ra21059g.Search in Google Scholar

[59] B. Jash and J. Muller, “A stable zinc(II)-mediated base pair in a parallel-stranded DNA duplex,” J. Inorg. Biochem., vol. 186, pp. 301–306, 2018, https://doi.org/10.1016/j.jinorgbio.2018.07.002.Search in Google Scholar PubMed

[60] Y. Miyake, H. Togashi, M. Tashiro, et al.., “Mercury(II)-mediated formation of thymine-Hg-II-thymine base pairs in DNA duplexes,” J. Am. Chem. Soc., vol. 128, pp. 2172–2173, 2006, https://doi.org/10.1021/ja056354d.Search in Google Scholar PubMed

[61] I. Sinha, C. F. Guerra, and J. Muller, “A highly stabilizing silver(I)-mediated base pair in parallel-stranded DNA,” Angew. Chem. Int. Ed., vol. 54, pp. 3603–3606, 2015, https://doi.org/10.1002/anie.201411931.Search in Google Scholar PubMed

[62] N. Zimmermann, E. Meggers, and P. G. Schultz, “A second-generation copper(II)-mediated metallo-DNA-base pair,” Bioorg. Chem., vol. 32, pp. 13–25, 2004, https://doi.org/10.1016/j.bioorg.2003.09.001.Search in Google Scholar PubMed

[63] N. Xia, F. Feng, C. Liu, et al.., “The detection of mercury ion using DNA as sensors based on fluorescence resonance energy transfer,” Talanta, vol. 192, pp. 500–507, 2019, https://doi.org/10.1016/j.talanta.2018.08.086.Search in Google Scholar PubMed

[64] Y. Wei, B. M. Li, X. Wang, and Y. X. Duan, “Magnified fluorescence detection of silver(I) ion in aqueous solutions by using nano-graphite-DNA hybrid and DNase I,” Biosens. Bioelectron., vol. 58, pp. 276–281, 2014, https://doi.org/10.1016/j.bios.2014.02.075.Search in Google Scholar PubMed

[65] X. Cui, L. Zhu, J. Wu, et al.., “A fluorescent biosensor based on carbon dots-labeled oligodeoxyribonucleotide and graphene oxide for mercury (II) detection,” Biosens. Bioelectron., vol. 63, pp. 506–512, 2015, https://doi.org/10.1016/j.bios.2014.07.085.Search in Google Scholar PubMed

[66] S. Y. Liu, W. D. Na, S. Pang, and X. G. Su, “Fluorescence detection of Pb2+ based on the DNA sequence functionalized CdS quantum dots,” Biosens. Bioelectron., vol. 58, pp. 17–21, 2014, https://doi.org/10.1016/j.bios.2014.02.013.Search in Google Scholar PubMed

[67] X. Y. Wang, X. Jiang, Z. H. Zhang, J. Li, A. H. Liang, and Z. L. Jiang, “A fluorescence and resonance Rayleigh scattering di-model probe was developed for trace K+ coupled N-doped carbon dot and aptamer,” J. Lumin., vol. 214, pp. 1–10, 2019, https://doi.org/10.1016/j.jlumin.2019.116559.Search in Google Scholar

[68] Z. Li, Y. N. Ni, and S. Kokot, “A new fluorescent nitrogen-doped carbon dot system modified by the fluorophore-labeled ssDNA for the analysis of 6-mercaptopurine and Hg (II),” Biosens. Bioelectron., vol. 74, pp. 91–97, 2015, https://doi.org/10.1016/j.bios.2015.06.014.Search in Google Scholar PubMed

[69] S. Hamd-Ghadareh and A. Salimi, “DNA-functionalized dye-loaded carbon dots: ultrabright FRET platform for ratiometric detection of Hg(II) in serum samples and cell microenvironment,” Ionics, vol. 25, pp. 4469–4479, 2019, https://doi.org/10.1007/s11581-019-02999-2.Search in Google Scholar

[70] Z. H. Zhang, J. Li, X. Y. Wang, A. H. Liang, and Z. L. Jiang, “Aptamer-mediated N/Ce-doped carbon dots as a fluorescent and resonance Rayleigh scattering dual mode probe for arsenic(III),” Microchim. Acta, vol. 186, pp. 1–9, 2019, https://doi.org/10.1007/s00604-019-3764-3.Search in Google Scholar PubMed

[71] K. Srinivasan, K. Subramanian, K. Murugan, and K. Dinakaran, “Sensitive fluorescence detection of mercury(II) in aqueous solution by the fluorescence quenching effect of MoS2 with DNA functionalized carbon dots,” Analyst, vol. 141, pp. 6344–6352, 2016, https://doi.org/10.1039/c6an00879h.Search in Google Scholar PubMed

[72] Y. Wei, B. M. Li, X. Wang, and Y. X. Duan, “A nano-graphite-DNA hybrid sensor for magnified fluorescent detection of mercury(II) ions in aqueous solution,” Analyst, vol. 139, pp. 1618–1621, 2014, https://doi.org/10.1039/c3an01482g.Search in Google Scholar PubMed

[73] T. Song, X. F. Zhu, S. H. Zhou, G. Yang, W. Gan, and Q. H. Yuan, “DNA derived fluorescent bio-dots for sensitive detection of mercury and silver ions in aqueous solution,” Appl. Surf. Sci., vol. 347, pp. 505–513, 2015, https://doi.org/10.1016/j.apsusc.2015.04.143.Search in Google Scholar

[74] C. Yuan, B. H. Liu, F. Liu, M. Y. Han, and Z. P. Zhang, “Fluorescence “Turn On” detection of mercuric ion based on bis(dithiocarbamato)copper(II) complex functionalized carbon nanodots,” Anal. Chem., vol. 86, pp. 1123–1130, 2014, https://doi.org/10.1021/ac402894z.Search in Google Scholar PubMed

[75] X. H. Gao, Y. Z. Lu, R. Z. Zhang, et al.., “One-pot synthesis of carbon nanodots for fluorescence turn-on detection of Ag+ based on the Ag+-induced enhancement of fluorescence,” J. Mater. Chem. C, vol. 3, pp. 2302–2309, 2015, https://doi.org/10.1039/c4tc02582b.Search in Google Scholar

[76] A. Bigdeli, F. Ghasemi, S. Abbasi-Moayed, et al.., “Ratiometric fluorescent nanoprobes for visual detection: design principles and recent advances – A review,” Anal. Chim. Acta, vol. 1079, pp. 30–58, 2019, https://doi.org/10.1016/j.aca.2019.06.035.Search in Google Scholar PubMed

[77] S. Chowdhury, B. Rooj, A. Dutta, and U. Mandal, “Review on recent advances in metal ions sensing using different fluorescent probes,” J. Fluoresc., vol. 28, pp. 999–1021, 2018, https://doi.org/10.1007/s10895-018-2263-y.Search in Google Scholar PubMed

[78] H. Chen, Y. Xie, A. M. Kirillov, et al.., “A ratiometric fluorescent nanoprobe based on terbium functionalized carbon dots for highly sensitive detection of an anthrax biomarker,” Chem. Commun., vol. 51, pp. 5036–5039, 2015, https://doi.org/10.1039/c5cc00757g.Search in Google Scholar PubMed

[79] X. Yan, H. X. Li, X. S. Han, and X. G. Su, “A ratiometric fluorescent quantum dots based biosensor for organophosphorus pesticides detection by inner-filter effect,” Biosens. Bioelectron., vol. 74, pp. 277–283, 2015, https://doi.org/10.1016/j.bios.2015.06.020.Search in Google Scholar PubMed

[80] X. Yan, H. X. Li, W. S. Zheng, and X. G. Se, “Visual and fluorescent detection of tyrosinase activity by using a dual-emission ratiometric fluorescence probe,” Anal. Chem., vol. 87, pp. 8904–8909, 2015, https://doi.org/10.1021/acs.analchem.5b02037.Search in Google Scholar PubMed

[81] J. J. Zhao, M. J. Huang, L. L. Zhang, et al.., “Unique approach to develop carbon dot-based nanohybrid near-infrared ratiometric fluorescent sensor for the detection of mercury ions,” Anal. Chem., vol. 89, pp. 8044–8049, 2017, https://doi.org/10.1021/acs.analchem.7b01443.Search in Google Scholar PubMed

[82] Y. Jiao, Y. F. Gao, Y. T. Meng, et al.., “One-step synthesis of label-free ratiometric fluorescence carbon dots for the detection of silver ions and glutathione and cellular imaging applications,” ACS Appl. Mater. Interfaces, vol. 11, pp. 16822–16829, 2019, https://doi.org/10.1021/acsami.9b01319.Search in Google Scholar PubMed

[83] Z. Q. Sun, Z. Y. Yang, L. Zhao, et al.., “Microwave-assisted fabrication of multicolor photoluminescent carbon dots as a ratiometric fluorescence sensor for iron ions,” New J. Chem., vol. 43, pp. 853–861, 2019, https://doi.org/10.1039/c8nj05324c.Search in Google Scholar

[84] Y. F. Wang, L. Yang, B. H. Liu, S. M. Yu, and C. L. Jiang, “A colorimetric paper sensor for visual detection of mercury ions constructed with dual-emission carbon dots,” New J. Chem., vol. 42, pp. 15671–15677, 2018, https://doi.org/10.1039/c8nj03683g.Search in Google Scholar

[85] C. Liu, D. H. Ning, C. Zhang, et al.., “Dual-colored carbon dot ratiometric fluorescent test paper based on a specific spectral energy transfer for semiquantitative assay of copper ions,” ACS Appl. Mater. Interfaces, vol. 9, pp. 18897–18903, 2017, https://doi.org/10.1021/acsami.7b05827.Search in Google Scholar PubMed

[86] S. Gogoi, R. Devi, H. S. Dutta, M. Bordoloi, and R. Khan, “Ratiometric fluorescence response of a dual light emitting reduced carbon dot/graphene quantum dot nanohybrid towards As(iii),” J. Mater. Chem. C, vol. 7, pp. 10309–10317, 2019, https://doi.org/10.1039/c9tc02199j.Search in Google Scholar

[87] B. Y. Han, Y. Li, X. X. Hu, et al.., “Paper-based visual detection of silver ions and l-cysteine with a dual-emissive nanosystem of carbon quantum dots and gold nanoclusters,” Analytical Methods, vol. 10, pp. 1–7, 2018, https://doi.org/10.1039/c8ay01044g.Search in Google Scholar

[88] S. M. Lu, D. Wu, G. L. Li, et al.., “Carbon dots-based ratiometric nanosensor for highly sensitive and selective detection of mercury(II) ions and glutathione,” RSC Adv., vol. 6, pp. 103169–103177, 2016, https://doi.org/10.1039/c6ra21309j.Search in Google Scholar

[89] G. Q. Xiang, Y. Ren, H. Zhang, et al.., “Carbon dots based dual-emission silica nanoparticles as ratiometric fluorescent probe for chromium speciation analysis in water samples,” Can. J. Chem., vol. 96, pp. 72–77, 2018, https://doi.org/10.1139/cjc-2017-0472.Search in Google Scholar

[90] Q. Q. Tan, R. R. Zhang, G. Y. Zhang, X. Y. Liu, F. L. Qu, and L. M. Lu, “Embedding carbon dots and gold nanoclusters in metal-organic frameworks for ratiometric fluorescence detection of Cu2+,” Anal. Bioanal. Chem., vol. 412, pp. 1317–1324, 2020, https://doi.org/10.1007/s00216-019-02353-5.Search in Google Scholar PubMed

[91] F. Ghasemi, M. R. Hormozi-Nezhad, and M. Mahmoudi, “A new strategy to design colorful ratiometric probes and its application to fluorescent detection of Hg(II),” Sensor. Actuator. B Chem., vol. 259, pp. 894–899, 2018, https://doi.org/10.1016/j.snb.2017.12.141.Search in Google Scholar

[92] X. X. Zhu, H. Jin, C. L. Gao, R. J. Gui, and Z. H. Wang, “Ratiometric, visual, dual-signal fluorescent sensing and imaging of pH/copper ions in real samples based on carbon dots-fluorescein isothiocyanate composites,” Talanta, vol. 162, pp. 65–71, 2017, https://doi.org/10.1016/j.talanta.2016.10.015.Search in Google Scholar PubMed

[93] Z. Q. Ye, R. Tang, H. Wu, B. B. Wang, M. Q. Tan, and J. L. Yuan, “Preparation of europium complex-conjugated carbon dots for ratiometric fluorescence detection of copper(II) ions,” New J. Chem., vol. 38, pp. 5721–5726, 2014, https://doi.org/10.1039/c4nj00966e.Search in Google Scholar

[94] P. J. Li, A. N. Ang, H. T. Feng, and S. F. Y. Li, “Rapid detection of an anthrax biomarker based on the recovered fluorescence of carbon dot-Cu(II) systems,” J. Mater. Chem. C, vol. 5, pp. 6962–6972, 2017, https://doi.org/10.1039/c7tc01058c.Search in Google Scholar

[95] D. Gu, L. Hong, L. Zhang, H. Liu, and S. M. Shang, “Nitrogen and sulfur co-doped highly luminescent carbon dots for sensitive detection of Cd (II) ions and living cell imaging applications,” J. Photochem. Photobiol. B Biol., vol. 186, pp. 144–151, 2018, https://doi.org/10.1016/j.jphotobiol.2018.07.012.Search in Google Scholar PubMed

[96] P. Devi, P. Rajput, A. Thakur, K. H. Kim, and P. Kumar, “Recent advances in carbon quantum dot-based sensing of heavy metals in water,” Trac. Trends Anal. Chem., vol. 114, pp. 171–195, 2019, https://doi.org/10.1016/j.trac.2019.03.003.Search in Google Scholar

[97] A. Pankaj, K. Tewari, S. Singh, and S. P. Singh, “Waste candle soot derived nitrogen doped carbon dots based fluorescent sensor probe: An efficient and inexpensive route to determine Hg(II) and Fe (III) from water,” J. Environ. Chem. Eng., vol. 6, pp. 5561–5569, 2018, https://doi.org/10.1016/j.jece.2018.08.059.Search in Google Scholar

[98] K. Rajendran and N. Rajendiran, “Bluish green emitting carbon quantum dots synthesized from jackfruit (Artocarpus heterophyllus) and its sensing applications of Hg (II) and Cr (VI) ions,” Mater. Res. Express, vol. 5, pp. 1–8, 2018, https://doi.org/10.1088/2053-1591/aaae4b.Search in Google Scholar

[99] S. Pawar, S. Kaja, and A. Nag, “Red-emitting carbon dots as a dual sensor for In3+ and Pd2+ in water,” ACS Omega, vol. 5, pp. 8362–8372, 2020, https://doi.org/10.1021/acsomega.0c00883.Search in Google Scholar PubMed PubMed Central

[100] Y. B. Wang, Q. Chang, and S. L. Hu, “Carbon dots with concentration-tunable multicolored photoluminescence for simultaneous detection of Fe3+ and Cu2+ ions,” Sensor. Actuator. B Chem., vol. 253, pp. 928–933, 2017, https://doi.org/10.1016/j.snb.2017.07.031.Search in Google Scholar

[101] H. Z. Lu, S. F. Xu, and J. Q. Liu, “One pot generation of blue and red carbon dots in one binary solvent system for dual channel detection of Cr3+ and Pb2+ based on ion imprinted fluorescence polymers,” ACS Sens., vol. 4, pp. 1917–1924, 2019, https://doi.org/10.1021/acssensors.9b00886.Search in Google Scholar PubMed

[102] F. Qu, S. Wang, D. Y. Liu, and J. M. You, “Differentiation of multi-metal ions based on fluorescent dual-emission carbon nanodots,” RSC Adv., vol. 5, pp. 82570–82575, 2015, https://doi.org/10.1039/c5ra16373k.Search in Google Scholar

[103] L. Y. Yu, G. J. Ren, M. Y. Tang, et al., “Effective determination of Zn2+, Mn2+, and Cu2+ simultaneously by using dual-emissive carbon dots as colorimetric fluorescent probe,” Eur. J. Inorg. Chem., pp. 3418–3426, 2018, https://doi.org/10.1002/ejic.201800474.Search in Google Scholar

[104] J. P. Song, Q. Ma, Y. Q. Liu, Y. Guo, F. Feng, and S. M. Shuang, “Novel single excitation dual-emission carbon dots for colorimetric and ratiometric fluorescent dual mode detection of Cu2+ and Al3+ ions,” RSC Adv., vol. 9, pp. 38568–38575, 2019, https://doi.org/10.1039/c9ra07030c.Search in Google Scholar PubMed PubMed Central

[105] Z. Y. Liu, W. Y. Jin, F. X. Wang, et al.., “Ratiometric fluorescent sensing of Pb2+ and Hg2+ with two types of carbon dot nanohybrids synthesized from the same biomass,” Sensor. Actuator. B Chem., vol. 296, pp. 1–8, 2019, https://doi.org/10.1016/j.snb.2019.126698.Search in Google Scholar

[106] D. J. Li, Y. Sun, Q. R. Shen, et al.., “Smartphone-based three-channel ratiometric fluorescent device and application in filed analysis of Hg2+, Fe3+ and Cu2+ in water samples,” Microchem. J., vol. 152, pp. 1–10, 2020, https://doi.org/10.1016/j.microc.2019.104423.Search in Google Scholar

[107] L. Y. Yu, L. Y. Zhang, G. J. Ren, et al.., “Multicolorful fluorescent-nanoprobe composed of Au nanocluster and carbon dots for colorimetric and fluorescent sensing Hg2+ and Cr6+,” Sensor. Actuator. B Chem., vol. 262, pp. 678–686, 2018, https://doi.org/10.1016/j.snb.2018.01.192.Search in Google Scholar

[108] K. Patir and S. K. Gogoi, “Nitrogen-doped carbon dots as fluorescence ON-OFF-ON sensor for parallel detection of copper) and mercury(II) ions in solutions as well as in filter paper-based microfluidic device,” Nanoscale Adv., vol. 1, pp. 592–601, 2019, https://doi.org/10.1039/c8na00080h.Search in Google Scholar PubMed PubMed Central

[109] B. Li, H. Ma, B. Zhang, et al.., “Dually emitting carbon dots as fluorescent probes for ratiometric fluorescent sensing of pH values, mercury(II), chloride and Cr(VI) via different mechanisms,” Microchim. Acta, vol. 186, pp. 1–10, 2019, https://doi.org/10.1007/s00604-019-3437-2.Search in Google Scholar PubMed

[110] Y. L. Wang, S. Y. Lao, W. J. Ding, Z. D. Zhang, and S. Y. Liu, “A novel ratiometric fluorescent probe for detection of iron ions and zinc ions based on dual-emission carbon dots,” Sensor. Actuator. B Chem., vol. 284, pp. 186–192, 2019, https://doi.org/10.1016/j.snb.2018.12.139.Search in Google Scholar

[111] C. Li, W. J. Liu, Y. J. Ren, X. B. Sun, W. Pan, and J. P. Wang, “The selectivity of the carboxylate groups terminated carbon dots switched by buffer solutions for the detection of multi-metal ions,” Sensor. Actuator. B Chem., vol. 240, pp. 941–948, 2017, https://doi.org/10.1016/j.snb.2016.09.068.Search in Google Scholar

[112] C. L. Shen, L. X. Su, J. H. Zang, X. J. Li, Q. Lou, and C. X. Shan, “Carbon nanodots as dual-mode nanosensors for selective detection of hydrogen peroxide,” Nanoscale Res. Lett., vol. 12, p. 447, 2017, https://doi.org/10.1186/s11671-017-2214-6.Search in Google Scholar PubMed PubMed Central

[113] F. Yarur, J. R. Macairan, and R. Naccache, “Ratiometric detection of heavy metal ions using fluorescent carbon dots,” Environ. Sci. Nano, vol. 6, pp. 1121–1130, 2019, https://doi.org/10.1039/c8en01418c.Search in Google Scholar

[114] P. J. Li, Y. Y. Hong, H. T. Feng, and S. F. Y. Li, “An efficient “off-on” carbon nanoparticle-based fluorescent sensor for recognition of chromium(VI) and ascorbic acid based on the inner filter effect,” J. Mater. Chem. B, vol. 5, pp. 2979–2988, 2017, https://doi.org/10.1039/c7tb00017k.Search in Google Scholar PubMed

[115] F. P. Shi, Y. Zhang, W. D. Na, X. Y. Zhang, Y. Li, and X. G. Su, “Graphene quantum dots as selective fluorescence sensor for the detection of ascorbic acid and acid phosphatase via Cr(VI)/Cr(III)-modulated redox reaction,” J. Mater. Chem. B, vol. 4, pp. 3278–3285, 2016, https://doi.org/10.1039/c6tb00495d.Search in Google Scholar PubMed

[116] S. A. Feng, Z. Gao, H. Liu, J. F. Huang, X. L. Li, and Y. L. Yang, “Feasibility of detection valence speciation of Cr(III) and Cr(VI) in environmental samples by spectrofluorimetric method with fluorescent carbon quantum dots,” Spectrochim. Acta Mol. Biomol. Spectrosc., vol. 212, pp. 286–292, 2019, https://doi.org/10.1016/j.saa.2018.12.055.Search in Google Scholar PubMed

[117] Y. Y. Xu, Z. Z. Zhang, P. Y. Lv, Y. H. Duan, G. P. Li, and B. X. Ye, “Ratiometric fluorescence sensing of Fe2+/3+ by carbon dots doped lanthanide coordination polymers,” J. Lumin., vol. 205, pp. 519–524, 2019, https://doi.org/10.1016/j.jlumin.2018.10.001.Search in Google Scholar

[118] I. Costas-Mora, V. Romero, I. Lavilla, and C. Bendicho, “In situ building of a nanoprobe based on fluorescent carbon dots for methylmercury detection,” Anal. Chem., vol. 86, pp. 4536–4543, 2014, https://doi.org/10.1021/ac500517h.Search in Google Scholar PubMed

[119] B. L. Xue, Y. Yang, R. Tang, et al.., “One-step hydrothermal synthesis of a flexible nanopaper-based Fe3+ sensor using carbon quantum dot grafted cellulose nanofibrils,” Cellulose, vol. 27, pp. 729–742, 2020, https://doi.org/10.1007/s10570-019-02846-7.Search in Google Scholar

[120] K. H. Lu, J. H. Lin, C. Y. Lin, C. F. Chen, and Y. C. Yeh, “A fluorometric paper test for chromium(VI) based on the use of N-doped carbon dots,” Microchim. Acta, vol. 186, pp. 1–7, 2019, https://doi.org/10.1007/s00604-019-3337-5.Search in Google Scholar PubMed

[121] Y. H. Yan, H. Yu, K. Zhang, et al.., “Dual-emissive nanohybrid of carbon dots and gold nanoclusters for sensitive determination of mercuric ions,” Nano Res., vol. 9, pp. 2088–2096, 2016, https://doi.org/10.1007/s12274-016-1099-5.Search in Google Scholar

[122] Y. J. Zhou, X. Y. Huang, C. Liu, et al.., “Color-multiplexing-based fluorescent test paper: dosage-sensitive visualization of arsenic(III) with discernable scale as low as 5 ppb,” Anal. Chem., vol. 88, pp. 6105–6109, 2016, https://doi.org/10.1021/acs.analchem.6b01248.Search in Google Scholar PubMed

[123] J. W. Ye, Y. J. Geng, F. H. Cao, et al.., “A smartphone-assisted paper-based analytical device for fluorescence assay of Hg2+,” Chem. Res. Chin. Univ., vol. 35, pp. 972–977, 2019, https://doi.org/10.1007/s40242-019-9234-y.Search in Google Scholar

[124] B. Ninwong, P. Sangkaew, P. Hapa, et al.., “Sensitive distance-based paper-based quantification of mercury ions using carbon nanodots and heating-based preconcentration,” RSC Adv., vol. 10, pp. 9884–9893, 2020, https://doi.org/10.1039/d0ra00791a.Search in Google Scholar PubMed PubMed Central

[125] S. W. Hu, S. Qiao, B. Y. Xu, X. Peng, J. J. Xu, and H. Y. Chen, “Dual-functional carbon dots pattern on paper chips for Fe3+ and ferritin analysis in whole blood,” Anal. Chem., vol. 89, pp. 2131–2137, 2017, https://doi.org/10.1021/acs.analchem.6b04891.Search in Google Scholar PubMed

[126] Y. P. Wu, X. Liu, Q. H. Wu, J. Yi, and G. L. Zhang, “Differentiation and determination of metal ions using fluorescent sensor array based on carbon nanodots,” Sensor. Actuator. B Chem., vol. 246, pp. 680–685, 2017, https://doi.org/10.1016/j.snb.2017.02.132.Search in Google Scholar

[127] Z. Wang, C. Xu, Y. X. Lu, et al.., “Fluorescence sensor array based on amino acid derived carbon dots for pattern-based detection of toxic metal ions,” Sensor. Actuator. B Chem., vol. 241, pp. 1324–1330, 2017, https://doi.org/10.1016/j.snb.2016.09.186.Search in Google Scholar

[128] C. G. Cheng, M. Xing, and Q. L. Wu, “Green synthesis of fluorescent carbon dots/hydrogel nanocomposite with stable Fe3+ sensing capability,” J. Alloys Compd., vol. 790, pp. 221–227, 2019, https://doi.org/10.1016/j.jallcom.2019.03.053.Search in Google Scholar

[129] A. Truskewycz, S. A. Beker, A. S. Ball, B. Murdoch, and I. Cole, “Incorporation of quantum carbon dots into a PVP/ZnO hydrogel for use as an effective hexavalent chromium sensing platform,” Anal. Chim. Acta, vol. 1099, pp. 126–135, 2020, https://doi.org/10.1016/j.aca.2019.11.053.Search in Google Scholar PubMed

[130] L. L. Guo, Y. H. Song, K. Y. Cai, and L. Wang, “‘On-off’ ratiometric fluorescent detection of Hg2+ based on N-doped carbon dots-rhodamine B@TAPT-DHTA-COF,” Spectrochim. Acta Mol. Biomol. Spectrosc., vol. 227, pp. 1–8, 2020, https://doi.org/10.1016/j.saa.2019.117703.Search in Google Scholar PubMed

[131] Q. Wu, X. J. Wang, S. A. Rasaki, et al.., “Yellow-emitting carbon-dots-impregnated carboxy methyl cellulose/poly-vinyl-alcohol and chitosan: stable, freestanding, enhanced-quenching Cu2+-ions sensor,” J. Mater. Chem. C, vol. 6, pp. 4508–4515, 2018, https://doi.org/10.1039/c8tc00660a.Search in Google Scholar

[132] D. D. Zhang, W. Z. Jiang, Y. P. Zhao, Y. Dong, X. Feng, and L. Chen, “Carbon dots rooted PVDF membrane for fluorescence detection of heavy metal ions,” Appl. Surf. Sci., vol. 494, pp. 635–643, 2019, https://doi.org/10.1016/j.apsusc.2019.07.141.Search in Google Scholar

[133] R. J. Ying, H. Z. Lu, and S. F. Xu, “Ion imprinted dual reference ratiometric fluorescence probe for respective and simultaneous detection of Fe3+ and Cu2+,” New J. Chem., vol. 43, pp. 6404–6410, 2019, https://doi.org/10.1039/c9nj01356c.Search in Google Scholar

[134] H. Z. Lu, C. W. Yu, and S. F. Xu, “A dual reference ion-imprinted ratiometric fluorescence probe for simultaneous detection of silver (I) and lead (II),” Sensor. Actuator. B Chem., vol. 288, pp. 691–698, 2019, https://doi.org/10.1016/j.snb.2019.03.069.Search in Google Scholar

[135] S. M. Song, F. Liang, M. L. Li, et al.., “A label-free nano-probe for sequential and quantitative determination of Cr(VI) and ascorbic acid in real samples based on S and N dual-doped carbon dots,” Spectrochim. Acta Mol. Biomol. Spectrosc., vol. 215, pp. 58–68, 2019, https://doi.org/10.1016/j.saa.2019.02.065.Search in Google Scholar PubMed

[136] A. K. Singh, V. K. Singh, M. Singh, et al.., “One pot hydrothermal synthesis of fluorescent NP-carbon dots derived from Dunaliella salina biomass and its application in on-off sensing of Hg (II), Cr (VI) and live cell imaging,” J. Photochem. Photobiol. A Chem., vol. 376, pp. 63–72, 2019, https://doi.org/10.1016/j.jphotochem.2019.02.023.Search in Google Scholar

[137] M. Cao, Y. Li, Y. Z. Zhao, C. Y. Shen, H. Y. Zhang, and Y. F. Huang, “A novel method for the preparation of solvent-free, microwave-assisted and nitrogen-doped carbon dots as fluorescent probes for chromium(vi) detection and bioimaging,” RSC Adv., vol. 9, pp. 8230–8238, 2019, https://doi.org/10.1039/c9ra00290a.Search in Google Scholar PubMed PubMed Central

[138] F. Huo, W. F. Liang, Y. R. Tang, et al.., “Full-color carbon dots with multiple red-emission tuning: on/off sensors, in vitro and in vivo multicolor bioimaging,” J. Mater. Sci., vol. 54, pp. 6815–6825, 2019, https://doi.org/10.1007/s10853-019-03370-6.Search in Google Scholar

[139] M. Zhang, W. T. Wang, P. Yuan, C. Chi, J. Zhang, and N. L. Zhou, “Synthesis of lanthanum doped carbon dots for detection of mercury ion, multi-color imaging of cells and tissue, and bacteriostasis,” Chem. Eng. J., vol. 330, pp. 1137–1147, 2017, https://doi.org/10.1016/j.cej.2017.07.166.Search in Google Scholar

[140] X. X. Yang, F. C. Cui, R. Ren, et al.., “Red-emissive carbon dots for “Switch-On” dual function sensing platform rapid detection of ferric ions and L-cysteine in living cells,” ACS Omega, vol. 4, pp. 12575–12583, 2019, https://doi.org/10.1021/acsomega.9b01019.Search in Google Scholar PubMed PubMed Central

[141] G. Q. Chen, Z. Guo, G. M. Zeng, and L. Tang, “Fluorescent and colorimetric sensors for environmental mercury detection,” Analyst, vol. 140, pp. 5400–5443, 2015, https://doi.org/10.1039/c5an00389j.Search in Google Scholar PubMed

[142] A. Hasan, N. M. Q. Nanakali, A. Salihi, et al.., “Nanozyme-based sensing platforms for detection of toxic mercury ions: An alternative approach to conventional methods,” Talanta, vol. 215, pp. 1–12, 2020, https://doi.org/10.1016/j.talanta.2020.120939.Search in Google Scholar PubMed

[143] C. Herrero-Latorre, J. Barciela-Garcia, S. Garcia-Martin, and R. M. Pena-Crecente, “Graphene and carbon nanotubes as solid phase extraction sorbents for the speciation of chromium: A review,” Anal. Chim. Acta, vol. 1002, pp. 1–17, 2018, https://doi.org/10.1016/j.aca.2017.11.042.Search in Google Scholar PubMed

[144] T. Yang, X. Y. Wang, L. Y. Wang, M. L. Chen, and J. H. Wang, “Biological cells in the speciation analysis of heavy metals,” Analytical Methods, vol. 8, pp. 8251–8261, 2016, https://doi.org/10.1039/c6ay02324j.Search in Google Scholar

[145] X. Zhao, J. S. Gao, X. He, et al.., “DNA-modified graphene quantum dots as a sensing platform for detection of Hg2+ in living cells,” RSC Adv., vol. 5, pp. 39587–39591, 2015, https://doi.org/10.1039/c5ra06984j.Search in Google Scholar

[146] D. Li, X. Yuan, C. N. Li, Y. H. Luo, and Z. L. Jiang, “A novel fluorescence aptamer biosensor for trace Pb(II) based on gold-doped carbon dots and DNAzyme synergetic catalytic amplification,” J. Lumin., vol. 221, pp. 1–9, 2020, https://doi.org/10.1016/j.jlumin.2020.117056.Search in Google Scholar

[147] L. P. Lin, Y. H. Wang, Y. L. Xiao, and W. Liu, “Hydrothermal synthesis of carbon dots codoped with nitrogen and phosphorus as a turn-on fluorescent probe for cadmium(II),” Microchim. Acta, vol. 186, pp. 1–7, 2019, https://doi.org/10.1007/s00604-019-3264-5.Search in Google Scholar PubMed

[148] K. M. Omer and A. Q. Hassan, “Chelation-enhanced fluorescence of phosphorus doped carbon nanodots for multi-ion detection,” Microchim. Acta, vol. 184, pp. 2063–2071, 2017, https://doi.org/10.1007/s00604-017-2196-1.Search in Google Scholar

[149] L. Sciortino, F. Messina, G. Buscarino, S. Agnello, M. Cannas, and F. M. Gelardi, “Nitrogen-doped carbon dots embedded in a SiO2 monolith for solid-state fluorescent detection of Cu2+ ions,” J. Nanopart. Res., vol. 19, pp. 1–11, 2017, https://doi.org/10.1007/s11051-017-3915-6.Search in Google Scholar

[150] Y. X. Ma, Y. L. Chen, J. J. Liu, Y. X. Han, S. D. Ma, and X. G. Chen, “Ratiometric fluorescent detection of chromium(VI) in real samples based on dual emissive carbon dots,” Talanta, vol. 185, pp. 249–257, 2018, https://doi.org/10.1016/j.talanta.2018.03.081.Search in Google Scholar PubMed

[151] W. J. Zhang, S. G. Liu, L. Han, H. Q. Luo, and N. B. Li, “A ratiometric fluorescent and colorimetric dual-signal sensing platform based on N-doped carbon dots for selective and sensitive detection of copper (II) and pyrophosphate ion,” Sensor. Actuator. B Chem., vol. 283, pp. 215–221, 2019, https://doi.org/10.1016/j.snb.2018.12.012.Search in Google Scholar

[152] Y. F. Gao, Y. Jiao, H. L. Zhang, et al., “One-step synthesis of a dual-emitting carbon dot-based ratiometric fluorescent probe for the visual assay of Pb2+ and PPi and development of a paper sensor (vol 7, pg 5502, 2019),” J. Mater. Chem. B, vol. 7, pp. 6643–6643, 2019, https://doi.org/10.1039/c9tb90133g.Search in Google Scholar PubMed

[153] H. Z. Xie, J. Dong, J. L. Duan, G. I. N. Waterhouse, J. Y. Hou, and S. Y. Ai, “Visual and ratiometric fluorescence detection of Hg2+ based on a dual-emission carbon dots-gold nanoclusters nanohybrid,” Sensor. Actuator. B Chem., vol. 259, pp. 1082–1089, 2018, https://doi.org/10.1016/j.snb.2017.12.149.Search in Google Scholar

[154] W. J. Niu, D. Shan, R. H. Zhu, S. Y. Deng, S. Cosnier, and X. J. Zhang, “Dumbbell-shaped carbon quantum dots/AuNCs nanohybrid as an efficient ratiometric fluorescent probe for sensing cadmium (II) ions and L-ascorbic acid,” Carbon, vol. 96, pp. 1034–1042, 2016, https://doi.org/10.1016/j.carbon.2015.10.051.Search in Google Scholar

[155] X. F. Guo, C. Liu, N. Li, S. D. Zhang, and Z. Y. Wang, “Ratiometric fluorescent test paper based on silicon nanocrystals and carbon dots for sensitive determination of mercuric ions,” R. Soc. Open Sci., vol. 5, pp. 1–8, 2018, https://doi.org/10.1098/rsos.171922.Search in Google Scholar PubMed PubMed Central

[156] B. Y. Han, T. T. Peng, M. B. Yu, et al.., “One-pot synthesis of highly fluorescent Fe2+-doped carbon dots for a dual-emissive nanohybrid for the detection of zinc ions and histidine,” New J. Chem., vol. 42, pp. 13651–13659, 2018, https://doi.org/10.1039/c8nj01858h.Search in Google Scholar

[157] M. Y. Deng, S. Wang, C. S. Liang, H. X. Shang, and S. M. Jiang, “A FRET fluorescent nanosensor based on carbon dots for ratiometric detection of Fe3+ in aqueous solution,” RSC Adv., vol. 6, pp. 26936–26940, 2016, https://doi.org/10.1039/c6ra02679f.Search in Google Scholar

[158] Y. S. Kuang, L. F. Chen, J. H. Lu, et al.., “A carbon-dot-based dual-emission probe for ultrasensitive visual detection of copper ions,” New J. Chem., vol. 42, pp. 19771–19778, 2018, https://doi.org/10.1039/c8nj04854a.Search in Google Scholar

[159] H. L. Fu, Z. Y. Ji, X. J. Chen, et al.., “A versatile ratiometric nanosensing approach for sensitive and accurate detection of Hg2+ and biological thiols based on new fluorescent carbon quantum dots,” Anal. Bioanal. Chem., vol. 409, pp. 2373–2382, 2017, https://doi.org/10.1007/s00216-017-0183-3.Search in Google Scholar PubMed

[160] Y. Y. Wang, J. He, M. D. Zheng, M. D. Qin, and W. Wei, “Dual-emission of Eu based metal-organic frameworks hybrids with carbon dots for ratiometric fluorescent detection of Cr(VI),” Talanta, vol. 191, pp. 519–525, 2019, https://doi.org/10.1016/j.talanta.2018.08.078.Search in Google Scholar PubMed

[161] Y. H. Wang, C. Zhang, X. C. Chen, et al.., “Ratiometric fluorescent paper sensor utilizing hybrid carbon dots-quantum dots for the visual determination of copper ions,” Nanoscale, vol. 8, pp. 5977–5984, 2016, https://doi.org/10.1039/c6nr00430j.Search in Google Scholar PubMed

[162] H. Y. Xu, K. N. Zhang, Q. S. Liu, Y. Liu, and M. X. Xie, “Visual and fluorescent detection of mercury ions by using a dually emissive ratiometric nanohybrid containing carbon dots and CdTe quantum dots,” Microchim. Acta, vol. 184, pp. 1199–1206, 2017, https://doi.org/10.1007/s00604-017-2099-1.Search in Google Scholar

[163] C. X. Wang, Y. J. Huang, K. L. Jiang, M. G. Humphrey, and C. Zhang, “Dual-emitting quantum dot/carbon nanodot-based nanoprobe for selective and sensitive detection of Fe3+ in cells,” Analyst, vol. 141, pp. 4488–4494, 2016, https://doi.org/10.1039/c6an00605a.Search in Google Scholar PubMed

[164] B. M. Cao, C. Yuan, B. H. Liu, C. L. Jiang, G. J. Guan, and M. Y. Han, “Ratiometric fluorescence detection of mercuric ion based on the nanohybrid of fluorescence carbon dots and quantum dots,” Anal. Chim. Acta, vol. 786, pp. 146–152, 2013, https://doi.org/10.1016/j.aca.2013.05.015.Search in Google Scholar PubMed

[165] K. Ahmad, S. K. Gogoi, R. Begum, M. P. Sk, A. Paul, and A. Chattopadhyay, “An interactive quantum dot and carbon dot conjugate for pH-sensitive and ratiometric Cu2+ sensing,” ChemPhysChem, vol. 18, pp. 610–616, 2017, https://doi.org/10.1002/cphc.201601249.Search in Google Scholar PubMed

[166] J. Hao, F. F. Liu, N. Liu, M. L. Zeng, Y. H. Song, and L. Wang, “Ratiometric fluorescent detection of Cu2+ with carbon dots chelated Eu-based metal-organic frameworks,” Sensor. Actuator. B Chem., vol. 245, pp. 641–647, 2017, https://doi.org/10.1016/j.snb.2017.02.029.Search in Google Scholar

[167] P. Y. Lv, Y. Y. Xu, Z. Liu, G. P. Li, and B. X. Ye, “Carbon dots doped lanthanide coordination polymers as dual-function fluorescent probe for ratio sensing Fe-2(+/3+) and ascorbic acid,” Microchem. J., vol. 152, pp. 1–6, 2020, https://doi.org/10.1016/j.microc.2019.104255.Search in Google Scholar

[168] Y. H. Zhang, X. Fang, H. Zhao, and Z. X. Li, “A highly sensitive and selective detection of Cr(VI) and ascorbic acid based on nitrogen-doped carbon dots,” Talanta, vol. 181, pp. 318–325, 2018, https://doi.org/10.1016/j.talanta.2018.01.027.Search in Google Scholar PubMed

[169] H. Y. Zhang, Y. Wang, S. Xiao, H. Wang, J. H. Wang, and L. Feng, “Rapid detection of Cr(VI) ions based on cobalt(II)-doped carbon dots,” Biosens. Bioelectron., vol. 87, pp. 46–52, 2017, https://doi.org/10.1016/j.bios.2016.08.010.Search in Google Scholar PubMed

[170] H. Q. Zhang, Y. H. Huang, Z. B. Hu, C. Q. Tong, Z. S. Zhang, and S. R. Hu, “Carbon dots codoped with nitrogen and sulfur are viable fluorescent probes for chromium(VI),” Microchim. Acta, vol. 184, pp. 1547–1553, 2017, https://doi.org/10.1007/s00604-017-2132-4.Search in Google Scholar

[171] H. Y. Wu, J. Wang, J. C. Xu, et al.., “Environmentally friendly cleaner water-soluble fluorescent carbon dots coated with chitosan: synthesis and its application for sensitivity determination of Cr(VI) ions,” J. Iran. Chem. Soc., vol. 15, pp. 23–33, 2018, https://doi.org/10.1007/s13738-017-1206-x.Search in Google Scholar

[172] J. L. Yu, L. Y. Hao, B. B. Dong, et al.., “Synthesis and characterization of a multi-functional on-off-on fluorescent oxidized graphitic carbon nitride nanosensor for iodide, chromium(vi), and ascorbic acid,” J. Mater. Chem. C, vol. 7, pp. 11896–11902, 2019, https://doi.org/10.1039/c9tc03955d.Search in Google Scholar

[173] J. L. Wang, Z. L. Wu, S. H. Chen, R. Yuan, and L. Dong, “A novel multifunctional fluorescent sensor based on N/S co-doped carbon dots for detecting Cr (VI) and toluene,” Microchem. J., vol. 151, pp. 1–8, 2019, https://doi.org/10.1016/j.microc.2019.104246.Search in Google Scholar

[174] R. Vaz, J. Bettini, J. G. F. Junior, et al.., “High luminescent carbon dots as an eco-friendly fluorescence sensor for Cr(VI) determination in water and soil samples,” J. Photochem. Photobiol. A Chem., vol. 346, pp. 502–511, 2017, https://doi.org/10.1016/j.jphotochem.2017.06.047.Search in Google Scholar

[175] S. K. Tammina and Y. L. Yang, “Highly sensitive and selective detection of 4-nitrophenol, and on-off-on fluorescence sensor for Cr (VI) and ascorbic acid detection by glucosamine derived n-doped carbon dots,” J. Photochem. Photobiol. A Chem., vol. 387, pp. 1–11, 2020, https://doi.org/10.1016/j.jphotochem.2019.112134.Search in Google Scholar

[176] D. Y. Tai, C. F. Liu, and J. S. Liu, “Facile synthesis of fluorescent carbon dots from shrimp shells and using the carbon dots to detect chromium(VI),” Spectrosc. Lett., vol. 52, pp. 194–199, 2019, https://doi.org/10.1080/00387010.2019.1607879.Search in Google Scholar

[177] L. Y. Sheng, B. H. Huangfu, Q. H. Xu, et al.., “A highly selective and sensitive fluorescent probe for detecting Cr(VI) and cell imaging based on nitrogen-doped graphene quantum dots,” J. Alloys Compd., vol. 820, pp. 1–10, 2020, https://doi.org/10.1016/j.jallcom.2019.153191.Search in Google Scholar

[178] B. H. Luo, H. Yang, B. X. Zhou, et al.., “Facile synthesis of luffa sponge activated carbon fiber based carbon quantum dots with green fluorescence and their application in Cr(VI) determination,” ACS Omega, vol. 5, pp. 5540–5547, 2020, https://doi.org/10.1021/acsomega.0c00195.Search in Google Scholar PubMed PubMed Central

[179] Y. S. Liu, Z. J. Chen, W. Li, et al.., “A nanocomposite probe consisting of carbon quantum dots and phosphotungstic acid for fluorometric determination of chromate(VI) with improved selectivity,” Microchim. Acta, vol. 185, pp. 1–10, 2018, https://doi.org/10.1007/s00604-018-2993-1.Search in Google Scholar PubMed

[180] L. Li, C. Y. Shao, Q. Wu, Y. J. Wang, and M. Z. Liu, “Green synthesis of multifunctional carbon nanodots and their applications as a smart nanothermometer and Cr(VI) ions sensor,” Nano, vol. 13, pp. 1–14, 2018, https://doi.org/10.1142/s1793292018501473.Search in Google Scholar

[181] H. Y. Li, D. Li, Y. Guo, Y. Yang, W. L. Wei, and B. Xie, “On-site chemosensing and quantification of Cr(VI) in industrial wastewater using one-step synthesized fluorescent carbon quantum dots,” Sensor. Actuator. B Chem., vol. 277, pp. 30–38, 2018, https://doi.org/10.1016/j.snb.2018.08.157.Search in Google Scholar

[182] J. Kaur, S. Sharma, S. K. Mehta, and S. K. Kansal, “Highly photoluminescent and pH sensitive nitrogen doped carbon dots (NCDs) as a fluorescent sensor for the efficient detection of Cr (VI) ions in aqueous media,” Spectrochim. Acta Mol. Biomol. Spectrosc., vol. 227, pp. 1–8, 2020, https://doi.org/10.1016/j.saa.2019.117572.Search in Google Scholar PubMed

[183] X. J. Gong, Y. Liu, Z. H. Yang, S. M. Shuang, Z. Y. Zhang, and C. Dong, “An “on-off-on” fluorescent nanoprobe for recognition of chromium(VI) and ascorbic acid based on phosphorus/nitrogen dualdoped carbon quantum dot,” Anal. Chim. Acta, vol. 968, pp. 85–96, 2017, https://doi.org/10.1016/j.aca.2017.02.038.Search in Google Scholar PubMed

[184] L. Y. Fang, L. Zhang, Z. Z. Chen, C. S. Zhu, J. J. Liu, and J. T. Zheng, “Ammonium citrate derived carbon quantum dot as on-off-on fluorescent sensor for detection of chromium(VI) and sulfites,” Mater. Lett., vol. 191, pp. 1–4, 2017, https://doi.org/10.1016/j.matlet.2016.12.098.Search in Google Scholar

[185] B. Bac and R. Genc, “Naked eye and smartphone applicable detection of toxic mercury ions using fluorescent carbon nanodots,” Turk. J. Chem., vol. 41, pp. 931–943, 2017, https://doi.org/10.3906/kim-1701-46.Search in Google Scholar

[186] Y. F. Wang, M. A. Wu, S. M. Yu, and C. L. Jiang, “Semi-quantitative and visual assay of copper ions by fluorescent test paper constructed with dual-emission carbon dots,” RSC Adv., vol. 8, pp. 12708–12713, 2018, https://doi.org/10.1039/c8ra00917a.Search in Google Scholar PubMed PubMed Central

[187] V. K. Singh, V. Singh, P. K. Yadav, et al.., “Bright-blue-emission nitrogen and phosphorus-doped carbon quantum dots as a promising nanoprobe for detection of Cr(VI) and ascorbic acid in pure aqueous solution and in living cells,” New J. Chem., vol. 42, pp. 12990–12997, 2018, https://doi.org/10.1039/c8nj02126k.Search in Google Scholar

[188] V. Roshni, S. Misra, M. K. Santra, and D. Ottoor, “One pot green synthesis of C-dots from groundnuts and its application as Cr(VI) sensor and in vitro bioimaging agent,” J. Photochem. Photobiol. A Chem., vol. 373, pp. 28–36, 2019.10.1016/j.jphotochem.2018.12.028Search in Google Scholar

[189] Y. Liu, X. J. Gong, Y. F. Gao, et al.., “Carbon-based dots co-doped with nitrogen and sulfur for Cr(VI) sensing and bioimaging,” RSC Adv., vol. 6, pp. 28477–28483, 2016, https://doi.org/10.1039/c6ra02653b.Search in Google Scholar

[190] S. Huang, E. L. Yang, J. D. Yao, X. Chu, Y. Liu, and Q. Xiao, “Nitrogen, phosphorus and sulfur tri-doped carbon dots are specific and sensitive fluorescent probes for determination of chromium(VI) in water samples and in living cells,” Microchim. Acta, vol. 186, pp. 1–10, 2019, https://doi.org/10.1007/s00604-019-3941-4.Search in Google Scholar PubMed

[191] S. Huang, E. L. Yang, J. D. Yao, Y. Liu, and Q. Xiao, “Red emission nitrogen, boron, sulfur co-doped carbon dots for “on-off-on” fluorescent mode detection of Ag+ ions and L-cysteine in complex biological fluids and living cells,” Anal. Chim. Acta, vol. 1035, pp. 192–202, 2018, https://doi.org/10.1016/j.aca.2018.06.051.Search in Google Scholar PubMed

[192] Y. Q. Chen, X. B. Sun, X. Y. Wang, W. Pan, G. F. Yu, and J. P. Wang, “Carbon dots with red emission for bioimaging of fungal cells and detecting Hg2+ and ziram in aqueous solution,” Spectrochim. Acta Mol. Biomol. Spectrosc., vol. 233, pp. 1–9, 2020, https://doi.org/10.1016/j.saa.2020.118230.Search in Google Scholar PubMed

[193] M. Tian, Y. M. Liu, Y. T. Wang, and Y. Zhang, “Facile synthesis of yellow fluorescent carbon dots for highly sensitive sensing of cobalt ions and biological imaging,” Analytical Methods, vol. 11, pp. 4077–4083, 2019, https://doi.org/10.1039/c9ay01244c.Search in Google Scholar

Received: 2020-09-04
Accepted: 2020-10-15
Published Online: 2020-11-05

© 2020 Pingjing Li and Sam F. Y. Li, published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 24.4.2024 from https://www.degruyter.com/document/doi/10.1515/nanoph-2020-0507/html
Scroll to top button