Small nanoparticles bring big prospect: The synthesis, modification, photoluminescence and sensing applications of carbon dots

Pengli Gao Zhigang Xie Min Zheng

Citation:  Pengli Gao, Zhigang Xie, Min Zheng. Small nanoparticles bring big prospect: The synthesis, modification, photoluminescence and sensing applications of carbon dots[J]. Chinese Chemical Letters, 2022, 33(4): 1659-1672. doi: 10.1016/j.cclet.2021.09.085 shu

Small nanoparticles bring big prospect: The synthesis, modification, photoluminescence and sensing applications of carbon dots

    作者简介: Pengli Gao received hr B.S. degree from Xinzhou Normal University (2017). She received her master's degree from the School of Chemistry and Life Sciences, Changchun University of Technology (2020). She is currently a Ph.D. candidate at the School of Biological Science and Medical Engineering, Beihang University. Her research interests focus on molecular imaging;
    Zhigang Xie is a full professor at Changchun Institute of Applied Chemistry (CIAC), Chinese Academy of Science (CAS). Dr Xie obtained his B.S. degree in Chemistry from Henan University in 2002 and his Ph.D. degree in Polymer Chemistry from CIAC in 2007. Before joining CIAC in 2012, he worked as a postdoctoral fellow at the Polytechnic University at Brooklyn and the University of North Carolina at Chapel Hill. His research focuses on polymeric nanomedicines and fluorescent nanoparticles;

English

  • Optical nanomaterials have high sensitivity and selectivity for sensing variety of substances (such as ions [1-4], biological substances [5-9] and organic pollutants [10-13]), microenvironments (pH [14-16] and temperature [17, 18]) and fungus [19-22]. In the early stage of the development of nanomaterials, semiconductor quantum dots were first prepared [23-25], but their intrinsic toxicity has become a stumbling block for the further biomedical applications [24]. With the continuous research of nanomaterials, low toxic nanomaterials have been developed, such as metal nanoparticles [26, 27], silicon quantum dots [28] and carbon-based nanomaterials [29-31]. Especially, carbon dots (CDs) have been widely concerned and studied, due to their unique properties, performances and diverse applications [32-38]. Generally, CDs are monodispersed spherical nanoparticles with the average size of 1–10 nm. CDs are generally composed of carbon core with π-conjugated domain as well as amorphous region that is constructed by various functional groups on the surface (such as carboxyl, hydroxyl, and amino groups), which endow CDs with various advantages including: superior stability, high water solubility, tunable photoluminescence, exceptional biocompatibility and low toxicity. Based on the above advantages, CDs can be employed as promising probes for sensing.

    Since CDs were discovered by accident in 2004 [39], their diverse properties have attracted more and more attention. In recent years, some reviews on the preparation, properties and applications of CDs have been published [40-44]. Among them, though several outstanding reviews have discussed about CDs as probes in the field of sensing. For example, Liu et al. summarized the sensing of CDs in solution and solid state [42], Sharma et al. summarized the sensing application of different substances based on green CDs [43]. However, there are no reviews focusing on summarizing the research progress of CDs nanosensors that based on different sensing mechanisms in recent years. In this review, first, the PL mechanisms of CDs are briefly discussed, which lay solid foundation for the development of CDs-based sensing systems. Next, we focus on introducing CDs-based nanosensors for the detection of various substances, microenvironments and fungus through different mechanisms. Finally, we look forward to the future research and development of CDs.

    The starting materials of CDs can be generally divided into three categories: inorganic, organic and biomass materials.

    The main paragraph text follows directly on here. Most inorganic precursors for preparing CDs are limited to carbon-based materials, such as carbon fibers, carbon nanotubes, candle soot, and graphene [45-47]. These large-sized carbon materials have a perfect sp2 hybrid carbon structure, but lack effective transition energy levels to produce fluorescence. In order to confer luminescent property to CDs, it is required to modify their size, symmetry and surface chemistry. The most common method is to treat these carbon-based materials in strong oxidizing acidic environment.

    It is universal to synthesize fluorescent CDs by using organics as starting materials. Commonly used organic precursors include small organic molecules, biological macromolecules, polymers, amino acids and carbohydrate compounds, which undergo dehydration polymerization and carbonization in the process of forming CDs [38, 48-50]. Since these precursors themselves contain abundant functional groups, such as amino, carboxyl and hydroxyl groups, which endow the as-prepared CDs with interesting properties and potential applications. Among them, citric acid is the most commonly used organic precursor [50].

    In addition to inorganic and organic precursors mentioned above, some environmentally friendly materials have been utilized as carbon sources to prepare CDs. CDs synthesized from natural products have been reported, including milk, eggs, fruits, honey, green leaves etc. [51-53]. Because this type of natural products is common in nature, CDs obtained from biomass are very popular. Moreover, biomass materials usually contain a large amount of nitrogen and oxygen elements, which are very effective for preparing heteroatom-doped CDs with high PL efficiency.

    The synthetic methods of CDs have gone through two phases. The first phase was in 2004–2010, researchers are mainly committed to the development of top-down synthetic routes. The other stage was from 2010, CDs were frequently synthesized from different starting materials via bottom-up routes. These researches not only simplified the preparation process, but also significantly improved optical properties of CDs. Generally, the preparation methods of CDs fall into two broad categories: top-down and bottom-up routes (Fig. 1). The top-down technique is "breaking down" bulk carbon sources into smaller-sized CDs through arc discharge [39], laser ablation [54] or electrochemical oxidation [55, 56]. CDs prepared by this method usually have a perfect sp2 hybrid structure, but they have poor PL performance and strict preparation conditions. In order to obtain exceptional PL CDs, it basically involves complicated purification, separation procedures and subsequent modification processes, which are not conducive to large-scale production. Therefore, it is imperative to modify the surface of CDs to improve their physical, chemical and optical properties. In order to overcome the disadvantages of the top-down method, researchers have proposed bottom-up route, which includes hydrothermal [57], solvothermal [58, 59], microwave [60] and pyrolysis methods [61, 62]. The bottom-up route has the advantages of relatively simple preparation process, mild reaction conditions and low equipment requirements, which is widely used for preparing CDs from organics or biomass.

    Figure 1

    Figure 1.  Top-down and bottom-up synthetic methods of CDs.

    The biggest advantage of using the top-down method to synthesize CDs is that carbonization and surface modification can be completed by one step to achieve special customized CDs. Generally, pre-modified CDs are synthesized via selecting specific starting materials by means of pyrolysis, microwave, hydrothermal or solvothermal one-pot synthetic method. Zheng et al. successfully prepared multicolor fluorescent CDs that can cross the blood-brain barrier and target glioma cells (C6) by pyrolysis using d-glucose and l-aspartic acid as raw materials [63]. After that, our group synthesized a series of CDs by changing the molar ratio of d-glucose and l-aspartic acid, and determined the optimal synthesis ratio for targeting C6 cells [61]. Beforehand, our group used cyanine dyes (CyOH) [58], diketopyrrolopyrrole (DPP) [64], porphyrin [65] or dopamine [35] as starting materials via one-pot synthetic method, to synthesize CDs with high photothermal or photodynamic efficiency. Heteroatom doping is also an effective way to tune the PL performance of CDs and endow CDs with special functions. Heteroatom-doped CDs are prepared from the precursors that contain heteroelements, including non-metal elements (N [66], S [67], P [68], B [69], F [70], Cl [71], Br [72], Si [73], Se [74]) and metal elements (Cu [75], Zn [76], Mn [77, 78], Fe [79], Al [80], Zr [50]). Among them, N and S have become common doped elements [81-84], because they have rich valence electrons. Recently, our group reported concentration-modulated PL of CDs [50] which were synthesized by using citric acid and thiourea as starting materials. With the increase of concentrations, the band gap of CDs gradually decreases, resulting in the red-shift of fluorescence (Fig. 2). Qu et al. successfully prepared N-doped CDs from citric acid and ethylenediamine via hydrothermal method and the PL quantum yield (PQY) of CDs reached up to 94% [85]. During the reaction, citric acid is dehydrated to form CDs, and the adjacent amide and carboxylic acid groups form pyrrole-N in the carbon skeleton through intramolecular dehydrogenation, and further converted into graphite-N under heating conditions, thus improving the PQY of CDs. At the same time, they also proved that different types of amino groups can achieve different doping effects. Some studies reported that surface defects that caused by the passivation of heteroatoms can act as excitation energy traps and further influence the PL property of CDs. For example, Xu et al. synthesized S-doped CDs with PQY of 67% using citric acid and sodium thiosulfate as starting materials [86]. S atoms play a role in catalyzing redox during the formation of doped CDs and introduce deeper degree of surface defects on S-CDs surface. With the development of research on CDs, researchers are not satisfied with the preparation of single element doped CDs. Co-doping of N and S or other elements can effectively enhance the PL of CDs. Dong et al. used citric acid and cysteine as precursors to prepare N, S co-doped fluorescent CDs with PQY of 73% by hydrothermal method [87]. N, S-CDs have excitation-independent behavior, and they proved that the PL of CDs originates from the radiation recombination of holes and electrons that were trapped by the surface of CDs. N atoms introduce a new surface state, while the introduction of S atoms enhances the new surface state. Ding et al. prepared S-CDs and N, S-CDs, which have similar sizes but different optical properties [88]. The PQY of S-CDs is 5.5%, while the PQY of N, S-CDs is 54.4%. With prolonging the reaction time to increase N content, the PQY of N, S-CDs also gradually increased. They proved that N existed in the form of C=N and C-N, which significantly improved PQY. At the same time, the doping of S element has a synergistic effect in N, S-CDs. Therefore, heteroatom doping can effectively modify the PL of CDs. So far, heteroatoms doping has become an effective method to modulate the properties of CDs, and the development of heteroatom-doped CDs has become a hot topic in the field of CDs.

    Figure 2

    Figure 2.  Top-down and bottom-up synthetic methods of CDs.

    The surface of CDs contains prolific active groups (such as amino, carboxyl and hydroxyl groups), which can be passivated with various agents and functional polymers through covalent or non-covalent interactions. Post-modification can reduce the defects and non-radiative recombination on the surface of CDs and improve the performance of CDs. Polyethylene glycol (PEG) is the most common surface passivators. Sun et al. prepared CDs with no visible fluorescence. After decorated with PEG at 120 ℃ for 72 h, the passivated CDs emitted bright fluorescence [54]. Peng and Travas-Sejdic passivated CDs with 4, 7, 10-trioxa-1, 13-tridecanediamine to enhance the fluorescence of CDs (Fig. 3a) [89]. The above results demonstrate that the PL property of CDs can be adjusted by surface passivation. After that, plenty of researches have shown that organic molecules (such as PEG [90], mercaptosuccinic acid [91], polyethyleneimine [92]) can act as passivating agent to improve the optical property of CDs. CDs with abundant amine groups were successfully conjugated with oxaliplatin to construct theranostic nanomedicine (CD-Oxa, Fig. 3b) [93]. The PL of blue fluorescent CDs (CDs-B, Fig. 3c) could be modulated by modification with vitamin C and acetaldehyde to obtain green (CDs-V) and red emissive CDs (CDs-A), respectively [94]. CDs can also be modified CDs with (Fig. 3d) drugs [70], organic dyes (Fig. 3e) [95, 96], and proteins [97, 98] via non-covalent interactions. Post-modification opens up an avenue to improve the performance of CDs and broaden their applications.

    Figure 3

    Figure 3.  (a) Schematic diagram of improving the photoluminescence of CDs by passivation. Copied with permission [89]. Copyright 2009, American Chemical Society. (b) Synthetic scheme for CD-Oxa and its applications in theranostics. Copied with permission [93]. Copyright 2014, Wiley. (c) Schematic syntheses and applications of CD-B, CD-V and CD-A. Copied with permission [94]. Copyright 2019, Royal Society of Chemistry. (d) Fabrication of FCD-DOX and FCD-BODIPY and the cellular uptake of FCD-DOX. Copied with permission [70]. Copyright 2019, American chemical society. (e) Schematic illustration of CDsG-AIE preparation and cellular imaging. Copied with permission [96]. Copyright 2016, American Chemical Society.

    Generally, CDs have strong absorption in the ultraviolet region and can extend into the visible light range [99]. In the UV-vis absorption spectrum, the absorption peak at 210–230 nm is ascribed to the π-π* electron transition of the C=C bond in carbon core, while the absorption peak at 330–360 nm may be due to n-π* electronic transition of C=N/C=O bond [100, 101]. The absorption spectrum of CDs can generally reflect the optimal PL excitation wavelength. It is worth noting that surface passivation or heteroatoms doping will change the surface groups and intrinsic structures of CDs, and ultimately adjust the absorption property of CDs. Correspondingly, different UV-vis absorption spectra reflect the different structures and compositions of CDs.

    Since the advent of CDs in 2004, researchers have successfully prepared CDs with different PL, which covers almost the entire spectral range from UV to near infrared region. There is no doubt that the development of CDs with various PL has become one of the research hotspots that researchers are most concerned on. The PL of CDs can generally be divided into two categories. The first type is excitation-dependent PL. With the change of excitation wavelength, the emission peak will red-shift or blue-shift. Other inorganic or organic luminescent materials are not provided with this type of PL property. Arcudi et al. prepared CDs by microwave-assisted methods [101]. They purified the crude product through dialysis and low-pressure gel permeation chromatography to obtain four types of CDs with different fluorescent properties. The other type is excitation-independent PL behavior, where emission peak will not move with the change of excitation wavelength. Jiang et al. reported red, green, and blue fluorescent CDs derived from o-phenylenediamine, p-phenylenediamine, and m-phenylenediamine, which have excellent excitation-independent PL properties [100].

    Abundant starting materials, diverse synthesis methods, varied reaction conditions make the PL of CDs present diversity and complexity [102]. Until now, explanation on the PL mechanisms of fluorescent CDs has been an open topic that is debated by researchers. According to the previous literature, the PL of CDs mainly includes several mechanisms: size effect, molecular state, surface state, defect state, environmental effect and cross-linking enhance emission effect.

    Some researchers believe that the PL of CDs is determined by the sp2 hybrid structure that distributes in carbon cores, which can be explained by size effect. Changing the sizes of CDs will affect the distribution of sp2 hybrid structure in carbon core, and further impact the corresponding energy band gap, which eventually results in the changes of CDs fluorescence. Sk et al. studied the relationship between the size effect of CDs and PL property through theoretical simulation [103]. They constructed CDs by density functional theory to explain that size changes will affect the emission of CDs. With the internal conjugated framework of CDs gradually extends outwards, the size of CDs increases and the emission appears obvious redshift. Meanwhile, multicolor fluorescent CDs derived from glutamic acid and p-phenylenediamine were prepared, by changing reaction solvent, precursors and pH [104]. With the aid of transmission electron microscopy (TEM), researchers found that the average size of blue, green, yellow, and red fluorescent CDs sequentially increased from 1.8 nm to 7.6 nm. They believed that the dehydration and carbonization processes were controlled by changing reaction conditions, which can promote the increase of sp2 hybrid framework structure in carbon core, causing the emission red shift. Therefore, the above results prove that PL property of CDs is affected by size.

    Molecular state is employed to explain the mechanism of CDs with high PL efficiency. It refers to that the PL property of CDs depends on the molecular residues on their surface. Molecular luminescence generally occurs during the synthesis of CDs [105]. In the initial stage of carbonization, due to the low temperature, starting materials undergo an intermolecular or intramolecular dehydration process, in which some small organic molecules with strong fluorescence may be generated. They are free in solution, or participate the formation of carbon core during the subsequent carbonization process and connect to the inside and surface of carbon skeleton. Variety of starting materials and synthetic conditions are capable of producing a plethora of chromophores, leading to multifarious PL [106]. Song et al. discovered a fluorescent molecule named IPCA during the synthesis of CDs from citric acid and ethylenediamine [107]. They verified that IPCA is the true emissive center of CDs, which connects on the surface and inside of the carbon core to make CDs emit fluorescence. Subsequently, CDs derived from citric acid and mercaptoethylamine were characterized by mass spectroscopy and 1H NMR, and a luminophore M1 was found (Fig. 4a) [108]. Combining the results of density functional theory calculation (Figs. 4b and c), they certified that the luminophore M1 is the origin of the CDs fluorescence.

    Figure 4

    Figure 4.  (a) Preparation of CDs and luminescent mechanism diagram. (b) The theoretically simulated optimal configuration of M1 and (c) the HOMO and LUMO energy level arrangement about its ground and excited states. Copied with permission [108]. Copyright 2016, American Chemical Society.

    Surface state is widely used to explain the mechanism of CDs PL. The abundant surface states of CDs are excited to undergo radiation recombination, which makes CDs emit different fluorescence [109]. Therefore, the functional groups or emissive centers on CDs have a great influence on their PL. The surface state of CDs could be modified chemically, which can change or generate new energy levels and further modulate PL property [110]. Li et al. utilized hydrothermal methods to change the passivated degree of CDs surface and study the luminescence mechanism of CDs [111]. They certificated that the contents of amino groups on the surface of CDs are the factors that affect the luminescence of CDs (Figs. 5ae). CDs were successfully synthesized from thiourea and p-phenylenediamine [110], and purified by column chromatography to obtain multicolor CDs from blue to red. Since the CDs have similar size distribution and graphite structures, size effect was excluded. Through structural analysis of these samples, the researchers found that different extent of surface oxidation caused the band gap of CDs gradually decrease, leading to the fluorescence shift from blue to red (Fig. 5f). After that, they also studied the discrepancy of CDs with different surface states, with the aid of single-particle fluorescence imaging technology. The result revealed the rationality about the surface states mechanism of CDs [112]. At the same time, model CDs with different contents of amino groups were constructed via the density functional theory (Fig. 5g). It was found that the more amino groups on the surface of CDs, the smaller the energy level, which caused the red shift of emission. The above results demonstrate that luminescence is related to the surface states of CDs.

    Figure 5

    Figure 5.  (a) Preparation of CDs and luminescent mechanism diagram. (b) The theoretically simulated optimal configuration of M1 and (c) the HOMO and LUMO energy level arrangement about its ground and excited states. Copied with permission [108]. Copyright 2016, American Chemical Society.

    The PL property of CDs is affected by the dangling bonds of sp3 hybrid on the edge of carbon skeleton structure or the defects in the internal structure. In the early development of CDs, bulk carbon materials such as carbon fibers, carbon nanotubes and graphene were flaked by arc discharge or laser ablation to obtain CDs (Fig. 6) [54]. The as-synthesized CDs have weak fluorescence. After passivated by strong reducing agents, the surface defects of CDs will be more stable to promote effective radiation recombination and get higher PQY. Genc et al. verified that CDs have internal defects by Raman spectroscopy and electron spin resonance spectroscopy [113]. They believed that the PL of CDs derived from surface or internal defects of CDs.

    Figure 6

    Figure 6.  The emission of CDs from passivated surface by PEG1500N. Copied with permission [54]. Copyright 2006, American Chemical Society.

    The environmental factors include temperature, solvent and pressure etc., which also affect the luminescence properties of CDs. Song's group discovered that CDs with amino groups [114], can form hydrogen bonds with solvent molecules, which is the main reason for the solvent discoloration of CDs. Our group prepared a kind of CDs with solvent discoloration [115], due to the interaction of CDs with solvents. Besides solvent, temperature and pressure also affect the luminescent properties of CDs.

    CEE is to explore the origin of strong luminescence of non-conjugated polymer dots with non-emissive or weakly emissive sub-chromophores. Zhu et al. used PEI as a model non-conjugated polymer to prepare polymer dots by cross-linking reaction [116]. They found that the decreased vibration and rotation of the cross-linked PEI is the main reason for the luminescence enhancement. In addition, the other three cross-linked polymer dots also proved the universality of CEE. Since then, another literature demonstrated that the luminescent properties of CDs were caused by CEE [117].

    Whether in chemistry or biology field, it is indispensable to detect environmental pollutants, bioactive substances, biological microenvironments, bacteria and viruses (Table 1). Usually, most of analytes exist in food, daily necessities, environment or soil, which are absorbed by human body through the food chain, and ultimately influence human health. Therefore, it is extremely important to develop nanosensors with non-toxicity, excellent biological safety and water solubility. So far, plenty of probes have been developed, such as semiconductor quantum dots [23-25] and fluorescent dyes [118-120]. Compared with these materials, CDs have advantages of good water solubility, high stability, low-toxicity and simple preparation [32-35, 57-62], which make them great potential for sensing.

    Table 1

    Table 1.  Summary of representative CDs used in various sensing applications.
    DownLoad: CSV

    The surface of CDs contains abundant functional groups. These functional groups act as active sites to react with analytes, and further change the PL of CDs to a large extent. Based on the changed PL signal, CDs can be employed as an effective fluorescence probe for sensing. In summary, the quenching or suppressed quenching mechanisms of CDs can be divided into eight types: static quenching, dynamic quenching, aggregation-induced quenching (AIQ), Förster energy resonance transfer (FERT), inner filter effect (IFE), photoinduced electron transfer (PET), oxidation/reduction and coordination reaction.

    6.1.1   Fluorescence quenching

    Static quenching-based sensing: The complexes between CDs and analytes are formed through intermolecular interactions, resulting in the fluorescence quenching of CDs. Static quenching has several obvious characteristics: (1) The fluorescence lifetime of CDs remains changed before and after quenching, (2) the absorption spectrum of CDs changes after adding analytes, (3) temperature affects the stability of complex, leading to reduced quenching efficiency.

    Based on the mechanism of static quenching, the fluorescence of CDs is quenched by various analytes [121-125]. Lu et al. synthesized bright yellow fluorescent CDs (Y-CDs) and used them as a multi-functional nanosensor for the detection of fluoroquinolones (FQs) (Fig. 7a) [123]. The fluorescence of Y-CDs can be effectively quenched by FQs. The addition of FQs leaded to the absorption peak of Y-CDs at 427 nm disappeared, while a new absorption peak appeared at 451 nm (Fig. 7b), indicating that FQs formed complexes with the functional groups on the surface of Y-CDs. The fluorescence lifetime did not change in the presence of FQs (Fig. 7c). At the same time, they also studied the quenching efficiency at different temperature (Fig. 7d), and the results proved that the higher the temperature, the lower the quenching efficiency. The results verified that the quenching mechanism was static quenching. Our group designed a kind of chiral CDs from l-aspartic acid and citric acid [126]. The chiral CDs have good selectivity on Sn2+ ions (Figs. 7e and f). The addition of Sn2+ had no effect on the fluorescence lifetime of CDs, demonstrating that the fluorescence quenching of chiral CDs induced by Sn2+ was static quenching. CDs prepared by Liu et al. have the property of strong acid-induced fluorescence enhancement. Fe3+ could coordinate with amino groups or amide groups on the surface of CDs to form CDs/Fe3+ hybrids (Fig. 7g) [122]. The fluorescence lifetime of CDs and CDs/Fe3+ hybrids is 8.06 ns and 7.85 ns, respectively, certifying that the quenching mechanism of CDs belonged to static quenching. Ma et al. prepared B-CDs by using citric acid and sodium tetraphenylborate as starting materials via hydrothermal method. When the ortho hydroxyl groups of catechol cross-linked with the boron hydroxyl groups of B-CDs, the fluorescence of B-CDs was quenched [127]. The bimodal quenching constant proves that fluorescence quenching of B-CDs is a static quenching process. When the concentration of catechol is 1–50 nmol/L, the limit of detection (LOD) is 0.25 nmol/L.

    Figure 7

    Figure 7.  (a) Schematic illustrating the sensing mechanism of Y-CDs to FQs. (b) UV-vis absorption. (c) Fluorescence decay curves of Y-CDs before and after adding FQs. (d) The quenching efficiency at 25 ℃ and 37 ℃. Copied with permission [123]. Copyright 2018, American Chemical Society. (e) Schematic illustration for fabricating chiral CDs-based nanoprobes for assaying Sn2+. (f) Fluorescence decay curves of chiral CDs before and after adding Sn2+. Copied with permission [126]. Copyright 2020, Elsevier. (g) Schematic representation of the sensing process of CD to Fe3+. Copied with permission [122]. Copyright 2018, American Chemical Society.

    Dynamic quenching-based sensing: When the excited state of CDs collides with quenchers, it returns to the ground state through energy or charge transfer, this process is dynamic quenching. The characteristics of dynamic quenching are the exact opposite of static quenching: (1) The fluorescence lifetime of CDs will change upon the addition of quenchers, (2) the UV-vis absorption spectrum of CDs is not affected, and (3) quenching efficiency is enhanced by increasing temperature.

    Hu et al. prepared N, S, Cl-CDs from salty food via acid hydrolysis, (Fe(CN)6)4- can effectively quench the fluorescence of N, S, Cl-CDs by dynamic quenching (Figs. 8ac) [128]. Thus, N, S, Cl-CDs can be successfully used to detect (Fe(CN)6)4- in food samples with high sensitivity. Liu et al. prepared N-CDs by oxidizing cork materials. Fe3+ can effectively quench the fluorescence of N-CDs [129]. They suggested that the excited electrons of CDs were transferred to the 3d orbit of Fe3+ (Fig. 8d), and these non-radiative electron holes cause fluorescence quenching. At the same time, after adding Fe3+, the fluorescence lifetime of CDs was attenuated, which further proved the dynamic quenching mechanism. Song et al. also demonstrated the multicolor N-CDs have similar quenching mechanism for detecting Fe3+ in living cells (Figs. 8e and f) [130]. Pb2+ ions could effectively quench the fluorescence of CDs synthesized from bamboo [131], with LOD less than 0.14 nmol/L (Fig. 8g).

    Figure 8

    Figure 8.  (a) Preparation and schematic illustration of N, S, Cl-CDs-based nanosensors for detection of [Fe(CN)6]4− in food samples. (b) The photoluminescence spectra of the N, S, Cl-CDs-based sensing system containing various concentrations of [Fe(CN)6]4−. (c) The fluorescence lifetime of N, S, Cl-CDs before and after adding [Fe(CN)6]4−. Copied with permission [128]. Copyright 2020, Elsevier. (d) Mechanism diagram of N-CDs coordinated with Fe3+ ions. Copied with permission [129]. Copyright 2020, Elsevier. (e) Images of cells incubated with different concentrations of NCDs/Fe3+ (f) The fluorescence lifetime of NCDs before and after adding Fe3+. Copied with permission [130]. Copyright 2017, American Chemical Society. (g) Schematic illustration of CDs for the detection of Pb2+. Copied with permission [131]. Copyright 2019, Elsevier.

    AIQ-based sensing: The AIQ effect is almost common for most fluorophores. When the fluorescent probe aggregates, the excited state energy of the probe is transferred to the ground state through a non-radiative transition, resulting in sharp decrease or even complete disappearance in the fluorescent signal. The coordination of CDs with analytes makes CDs aggregate, which causes the fluorescence quenching. The main manifestation is that the average size becomes larger after adding quenchers.

    Yue et al. prepared ethylenebis-(oxyethylenetriaza)tetraacetic acid (EGTA) modified CDs by hydrothermal method for the detection of Ca2+ [132]. After adding Ca2+, EGTA on the surface of CDs can capture Ca2+ (Figs. 9ac) and form the aggregates, leading to the fluorescence quenching of CDs. CDs synthesized from l-threonine and P2O5 have high sensitivity to Au3+ [133]. The addition of Au3+ makes the average size of CDs increase from 0.62 ± 0.4 nm to 21.1 ± 0.8 nm (Fig. 9d), proving that Au3+ induces the aggregation of CDs and fluorescence quenching. Sun et al. successfully prepared CDs that have different fluorescence quenching to Ce3+, Fe3+ and Cu2+ (Figs. 9eh) [134].

    Figure 9

    Figure 9.  (a) Schematic diagram of Ca2+ ions induce CDs aggregation. The hydrodynamic diameters of CDs before (b) and after (c) the addition of Ca2+ ions. Copied with permission [132]. Copyright 2019, American Chemical Society. (d) The hydrodynamic diameters of CDs and CDs+Au3+. Copied with permission [133]. Copyright 2018, Elsevier. (e) Schematic illustration of CDs for detection of Ce3+, Fe3+ and Cu2+. TEM images of CDs with the addition of (f) Ce3+, (g) Fe3+, and (h) Cu2+. Copied with permission [134]. Copyright 2018, American Chemical Society

    FRET-based sensing: FRET needs to satisfy the following conditions between the excited state CDs and the ground state quencher: (1) The emission spectrum of CDs (energy donor) overlaps with the absorption spectrum of quencher (energy accepter), (2) the distance between CDs and quencher range from 1 nm to 10 nm. When the energy resonance transfer occurs, it is often accompanied by the attenuation of the fluorescence lifetime of CDs.

    FRET is also one of the common quenching mechanisms for CDs-based nanosensors. For example, our group developed amphiphilic CDs for sensing environmental nitro-aromatic explosives (Fig. 10a) based on FRET [135]. Compared with other nitroaromatics (Figs. 10b and c), we observed that the fluorescence spectrum of CDs had a better overlap with the UV-vis absorption spectrum of 2, 4, 6-trinitrophenol (TNP), which resulted in the fluorescence of amphiphilic CDs was effectively quenched. Hu et al. designed red dual-emission CDs that can sensitively detect methyl blue [125]. They found that fluorescence spectrum of CDs partially overlapped with absorption spectrum of methyl blue, which promoted non-radiative recombination and quenched the fluorescence of CDs. Kudr et al. used microwave pyrolysis to prepare fluorescent CDs from citric acid and diethylenetriamine for detecting DNA damage [136]. They explored the competition experiment of the binding of CDs and ethidium bromide (EtBr) with DNA. CDs acted as energy donors, while EtBr acted as energy acceptors, promoting the binding of CDs with DNA though FRET. Effective energy resonance transfer occurred between 3-aminophenylboronic acid-derived CDs and MnO2, as evidenced by quenched fluorescence [137]. With the concentration of MnO2 increased, the quenching efficiency of CDs gradually increased, but the fluorescence lifetime of CDs markedly decreased.

    Figure 10

    Figure 10.  (a) Schematic diagram for synthesizing amphiphilic fluorescent CDs and their application in TNP detection. (b) Fluorescence quenching efficiency of CDs upon the addition different nitroaromatics. (c) Absorption spectra of various nitroaromatic compounds and the emission spectrum of CDs. Copied with permission [135]. Copyright 2019, Elsevier.

    IFE-based sensing: When the UV-vis spectrum of quencher overlaps with the excitation or emission spectrum of CDs, inner filter effect (IFE) will occur. Quenchers can not only shield the excitation light of CDs, but also can absorb the emission of CDs. The enhanced absorption of quencher indicates that the fluorescence of CDs is quenched. Most importantly, the IFE-based fluorescence quenching mechanism does not affect the fluorescence lifetime of CDs.

    CDs sensing application that based on IFE for the detection of Cr2O72- was first developed by Zheng et al. in 2013 [138]. After adding Cr2O72-, the fluorescence of CDs was completely quenched under 365 nm excitation (Fig. 11a), and because the excitation spectrum of CDs completely overlapped with the absorption spectrum of Cr2O72- (Fig. 11b), which met with the condition of IFE. Xu et al. prepared CDs via hydrothermal method by using (3-aminopropyl)triethoxysilane and glycerin as starting materials. CDs were employed as a ratiometric nanosensor for the detection of cholinesterases [5]. Huang et al. developed fluorescent N, Co-CDs that can be sensitively quenched by 2, 3-diaminophenazine through IFE [139]. N, Co-CDs can be employed as a ratiometric fluorescent probe to effectively detect cholesterol and uric acid in human serum with LOD of 3.6 nmol/L and 3.4 nmol/L, respectively.

    Figure 11

    Figure 11.  (a) Schematic illustration of fluorescence assays of CDs for the detection of Cr2O72− based on IFE. (b) Fluorescence excitation (red solid line) and emission (blue solid line) spectra of CDs and UV-vis absorption spectrum of Cr2O72− (black solid line). Copied with permission [138]. Copyright 2013, American Chemical Society.

    PET-based sensing: In the process of PET quenching, CDs act as electron donors or acceptors, quenchers cause fluorescence quenching of CDs through electron transfer, this quenching manner is defined as PET quenching.

    Wu et al. constructed a mitochondrial-targeted CDs-based nanosensor to detect peroxynitrite (ONOO-) in living cells [140]. The addition of ONOO- caused the fluorescence of CDs drop off gradually. When its concentration was in the range of 0.2–10 µmol/L, the LOD is 13.5 nmol/L. Experiments showed that ONOO- reacted with o-diaminobenzene on the surface of CDs to generate benzotriazole, and the surface electrons of CDs were transferred to benzotriazole, resulting in the quenched fluorescence (Fig. 12a). Li et al. developed red-emissive CDs for detecting organophosphorus pesticides (OPs) (Fig. 12b) [141]. In the presence of dopamine (DA), acetylcholine (ATCh) and acetylcholinesterase (AChE), OPs inhibited the activity of AChE that blocked the hydrolysis of ATCh, which promoted the polymerization of DA to form polydopamine (PDA). PDA could effectively absorb the electrons of CDs, and then transfered them from the highest occupied molecular orbital (HOMO) to the lowest unoccupied molecular orbital (LUMO). The results indicated that the quenching was attributed to PET mechanism. He et al. synthesized blue fluorescent CDs by using citric acid and EDTA as starting materials [142]. A series of experiments were carried out to prove that Hg2+ quenched the fluorescence of CDs through PET mechanism (Fig. 12c). In addition to the analytes mentioned above, testosterone [143], water [144], hydrogen peroxide [145], and rhodamine B were detected via PET mechanism.

    Figure 12

    Figure 12.  (a) Schematic illustration of preparation of CDs and a suggested PET mechanism of fluorescence quenching of C-dots-TPP in the presence of ONOO. Copied with permission [140]. Copyright 2017, Elsevier. (b) Synthesis route of CDs and schematic illustration of h-CDs/DA-based sensor for OPs detection. Copied with permission [141]. Copyright 2020, American Chemical Society. (c) Schematic of CDs as sensors for the sensing process of Hg2+ ions. Copied with permission [142]. Copyright 2016, Elsevier.

    Fluorescence recovery: "On-Off" nano-hybrid systems have been constructed between fluorescent CDs and quenchers through different interactions, in which quenchers can be decomposed or complexed by analytes with the aid of various chemical reactions, and restoring the fluorescence of CDs. This type of systems has been widely reported.

    Oxidation/reduction reaction-based sensing: The fluorescence of 2, 5-diaminobenzenesulfonic acid-derived RD-CDs can be quenched by hypochlorite (ClO-) by means of oxidation reaction as well as forming nonluminescent complex. Then the oxidant of RD-CDs was deoxygenized by ascorbic acid (AA), leading to the recovery of RD-CDs fluorescence (Fig. 13a) [121]. Zheng et al. prepared CDs from citric acid and diethylenetriamine via pyrolysis. The fluorescence of CDs was quenched via IFE. After adding AA, Cr(VI) was reduced to Cr(III), IFE was eliminated, and the fluorescence of CDs was restored (Fig. 13b) [138]. Xu et al. constructed a nonluminescent pCDs-MnO2 for detecting GSH [146]. After adding GSH to pCDs-MnO2 solution, pCDs were released through a series of reduction processes, and their fluorescence was restored.

    Figure 13

    Figure 13.  (a) Reversible sensing performance of RD-CDs for ClO and AA in colorimetric and fluorescence modes. Copied with permission [121]. Copyright 2019, American Chemical Society. (b) Schematic illustration of fluorescence assays for AA based on reduction reaction of CD-Cr(VI). Copied with permission [138]. Copyright 2013, American Chemical Society.

    Coordination reaction-based sensing: Gao et al. constructed CDs/Fe3+ complex through PET mechanism, which could sensitively detect phytic acid (PA). Since PA has a higher affinity with Fe3+ than CDs, Fe3+ preferentially coordinated with PA, causing the remove of Fe3+ from the surface of CDs. Thus, the fluorescence of CDs was restored [147]. Similarly, Hu et al. developed CDs/Fe3+ nanosensors to effectively detect pyrophosphate. The fluorescence of CDs was restored by the coordination between Fe3+ and pyrophosphate [148]. In the concentration range of 2–600 µmol/L, the LOD of pyrophosphate reached 0.86 µmol/L. Meanwhile, our group constructed chiral CD-Sn hybrids for sensing lysine enantiomers. l-Lysine (l-Lys) could complex with Sn2+ to prompt l-Lys to flee from the surface of CDs, recovering the fluorescence of chiral CDs [126]. However, d-Lys had no effect on the fluorescence of chiral CDs. Thus, CD-Sn was used for sensing Lys enantiomers and the LOD of l-Lys was 3.34 µmol/L.

    6.2.1   CDs for microenvironment sensing

    It is essential to develop nanosensors for detecting microenvironmental conditions. Since chemical processes and major diseases are closely related to abnormal pH fluctuations, it is essential to develop CDs-based pH nanosensors. Our group constructed two kinds CDs (CD-A and CD-B) derived from BODIPY dyes A/B with PEG1500. Both CD-A and CD-B can be used for pH sensing in extreme acidic environments in bacteria (Fig. 14a) [14]. Guo et al. prepared N, S-CDs from p-aminobenzenesulfonic acid and o-phenylenediamine by solvothermal method. The pH titration experiment proved that N, S-CDs could act as a ratiometric fluorescent probe to detect pH with good reversibility [15]. The protonation and deprotonation of pyridine N and pyrrole N in CDs manifested the response to pH. When pH dropped from 3.0 to 1.0, a new emission peak appeared at 634 nm, which could be attributed to the charge transfer of functional groups on the surface of CDs in an acidic environment. Our group has developed lysosome-targeting probes based on CDs to detect pH changes in vitro and in vivo (Fig. 14b) [149]. CDs have a good pH-sensitivity in pH range from 4.0 to 8.0. CDs exhibit good performance in quantitatively measuring pH changes in living cells and monitoring pH fluctuations in organisms. In short, the design and exploitation of CDs-based pH nanosensors has opened up a new path for disease diagnosis.

    Figure 14

    Figure 14.  (a) Synthetic scheme for CDs and their application as pH probes. Copied with permission [14]. Copyright 2018, Royal Society of Chemistry. (b) Schematic illustrating the fabrication of CDs, their lysosome targeting ability and applications for probing pH changes in vitro and in vivo. Copied with permission [149]. Copyright 2020, Elsevier. (c) Confocal microscopy images of N, S, I-CDs-colon cancer cell HT-29 with corresponding fluorescence field at 15, 25, 35 and 15 ℃, respectively. Copied with permission [17]. Copyright 2020, Elsevier. (d) Schematic mechanism for the temperature-dependent fluorescence intensity of N-CDs. Copied with permission [16]. Copyright 2015, American Chemical Society. (e) The linear calibration of fluorescence intensities for the W-CDs at 1.17 mg/mL in the presence of various amount of E. coli O157:H7 within a concentration range of 105–108 cfu/mL. Copied with permission [19]. Copyright 2016, Elsevier.

    Whether in chemical process or living system, temperature is also one of the important parameters. Accurately detecting the temperature of various environments has also become top priority. Compared with other nanomaterials, CDs have greater advantages for temperature sensing. Kalytchuk et al. constructed N, S, I-CDs from C3N3S3, potassium iodate (KIO3) and ethylenediamine (EDA), which were successfully employed for sensing the temperature in HT-29 cells (Fig. 14c) [17]. Yang et al. synthesized N-CDs from C3N4 and ethanediamine [16]. When temperature increased from 20 ℃ to 80 ℃, the fluorescence of N-CDs gradually decreased (Fig. 14d).

    6.2.2   CDs for subcellular organelle and bacteria sensing

    CDs have the advantages of good biocompatibility, excellent resistance to photobleaching and non-toxicity, and have been employed for sensing subcellular organelles and bacteria. Since organelles play a key role in cell function, the imaging and location of certain subcellular structures have an important guiding role in the origin of some diseases. Wu's group has studied the imaging of subcellular structures [150-154]. For example, Hua et al. prepared red fluorescent CDs that were doped with metals, and they selected nickel-doped CDs to achieve high-resolution imaging of cell nucleus [153]. CDs can be employed not only for nucleus imaging, but also for other organelles imaging, such as mitochondria and lysosomes [154]. The fluorescent CDs prepared from glycerol and silane possess mitochondrial targeting function, which can distinguish cancer cells from normal cells based on the difference in mitochondrial membrane potential [151].

    Bacteria and fungi are widespread in nature, and their infection is one of the biggest challenges to human health globally. Thus, it is urgent to develop sensors for detecting bacteria and fungi with high sensitivity and selectivity. CDs-based nanosensors for rapidly and accurately detecting bacteria or fungi have been reported [155-157]. Wang et al. used papaya powder as carbon source to synthesize W-CDs by hydrothermal method. W-CDs exhibit excellent fluorescence response to E. coli O157:H7 [19]. As the concentration of E. coli O157:H7 increased, the fluorescence of W-CDs gradually enhanced (Fig. 14e), and the LOD is 9.5 × 104 cfu/mL. They believed that W-CDs could interact with the FimH protein in E. coli O157:H7, resulting in the fluorescence enhancement of W-CDs. On the contrary, the obtained E-CDs from ethanol have no fluorescent response to E. coli O157:H7. Furthermore, Wang et al. constructed a class of aptamer-conjugated CDs (CD-apt). CD-apt can specifically detect Salmonella typhimurium with high selectivity [20]. Hua et al. prepared fluorescent CDs from bacteria (S. aureus or E. coli) for the first time [21], which can selectively stain dead bacteria and fungi. It has also been reported that NPS-CDs prepared from yeast extracts can be used to stain dead bacteria [158]. CDs-605 synthesized from L. plantarum can be able to image biofilm-encased microorganisms (e.g., E. coli, S. oneidensis, P. aeruginosa, S. aures, and T. reesei) [22].

    6.2.3   CDs for sensing volatile organic compounds

    Previously, our group developed hydrophilic CDs with solvatochromism, which can be applied as optical noses for sensing volatile organic compounds (VOCs) [115]. We dropped the CDs solution onto the indicator paper to obtain a paper-based sensor for detecting VOCs (Fig. 15). Moreover, the color change of indicator paper is reversible, proving that the paper-based sensor has the advantage of fast response. Furthermore, Wang et al. designed fluorescent complexes (MCM/CDs) from mesoporous materials and CDs, which integrated the physical adsorption of mesoporous materials and the fluorescent property of CDs [159]. MCM/CQDs could sensitively detect acetic acid gas with the LOD of 0.2 µmol/L.

    Figure 15

    Figure 15.  (a) The reversible emission color changes of CDs indicator paper under UV lamp (365 nm). (b) Photographs of the paper sensor printed with CDs after exposed to the VOCs under the UV lamp (365 nm). Copied with permission [115]. Copyright 2016, Royal Society of Chemistry.

    Survival of the fittest is not only the law of survival in nature, but also the law in the field of scientific research. With the continuous development of nanotechnology, nanomaterials are constantly being updated. Compared with other organic or inorganic nanomaterials, CDs possess ultrasmall size, remarkable photoluminescence, high water solubility, low toxicity, robust stability, extraordinary biocompatibility, and easy surface functionalization, which make CDs superior in a variety of research fields.

    Here, we emphasized the modification of CDs, including pre-modification and post-modification. For pre-modification, in the special customization process of CDs, CDs are endowed with multiple luminescent properties and functions by selecting various starting materials and reaction conditions. On the other hand, for post-modification, the surface of CDs contains plentiful functional groups, which can react with a multitude of functional materials through covalent or non-covalent interactions.

    Due to the diversity of preparation conditions, the obtained CDs are invested with various PL properties, so the corresponding luminescence mechanisms are also different. According to the currently reported literature, the luminescence mechanism of CDs can be classified into four categories: size effect, molecular state, surface state and defect state. In this review, we have explained the luminescence mechanisms of CDs in detail. However, by changing the reaction conditions during the synthesis of CDs, the corresponding luminescence mechanisms will also be very different, and in many cases they will appear synergistically. Because of the infinite possibilities of CDs, their PL has been continuously exploited by researchers. Continuous exploration the PL mechanism of CDs is a top priority, because CDs may have many interesting and meaningful properties that we need to discover.

    When CDs come to sensing, we summarized CDs nanosensors with different mechanisms, including static quenching, dynamic quenching, AIQ, FERT, IFE, PET, oxidation/reduction and coordination reaction. Second, CDs-based nanosensors for probing microenvironments, bacteria and virus, and VOCs were stated. However, the potential applications of CDs are far more than the content of this review. We believe that CDs-based nanosensors still have a long way to go.

    Although CDs-based fluorescent nanosensors have made great progress, it is worth noting that there are very few reports on CDs for sensing viruses. However, viruses have seriously affected human health. Therefore, designing and meticulously customizing CDs-based nanosensors for the detection of viruses are still a stupendous challenge. There is no doubt that overcoming this challenge will open up a brand-new research path for the development of sensors, and will further promote the development of CDs. We firmly believe that CDs, such small-sized fluorescent nanomaterials, have a broader development prospect in the future.

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

    The financial support from the National Natural Science Foundation of China (No. 51873023), the Jilin Province Science and Technology Research Project (No. 20200201088JC).


    1. [1]

      S.R. Ankireddy, J. Kim, Sens. Actuators B: Chem. 255 (2018) 3425-3433. doi: 10.1016/j.snb.2017.09.172

    2. [2]

      V.M. Naik, D.B. Gunjal, A.H. Gore, et al., Anal. Bioanal. Chem. 412 (2020) 2993-3003. doi: 10.1007/s00216-020-02550-7

    3. [3]

      G. Zuo, A. Xie, J. Li, et al., J. Phys. Chem. C 121 (2017) 26558-26565. doi: 10.1021/acs.jpcc.7b10179

    4. [4]

      Y. Jiao, Y. Gao, Y. Meng, et al., ACS Appl. Mater. Interfaces 11 (2019) 16822-16829. doi: 10.1021/acsami.9b01319

    5. [5]

      X. Xu, Y. Cen, G. Xu, et al., Biosens. Bioelectron. 131 (2019) 232-236. doi: 10.1016/j.bios.2019.02.031

    6. [6]

      B. Kong, T. Yang, P. Hou, et al., Luminescence 35 (2020) 222-230. doi: 10.1002/bio.3717

    7. [7]

      W. Yang, J. Ni, F. Luo, et al., Anal. Chem. 89 (2017) 8384-8390. doi: 10.1021/acs.analchem.7b01705

    8. [8]

      C.C. Wang, P.T. Huang, H. Shang Kou, S.M. Wu, Sens. Actuators B: Chem. 311 (2020) 127916.

    9. [9]

      R.F. Zhu, W.Y. Huang, X.F. Ma, et al., Anal. Chim. Acta 1089 (2019) 131-143. doi: 10.1016/j.aca.2019.08.061

    10. [10]

      N.K.R. Bogireddy, R. Cruz Silva, M.A. Valenzuela, V. Agarwal, J. Hazard. Mater. 386 (2020) 121643.

    11. [11]

      A. Singh, E. Eftekhari, J. Scott, et al., Sustain. Mater. Technol. 25 (2020) e00159.

    12. [12]

      N. Thongsai, P. Jaiyong, S. Kladsomboon, I. In, P. Paoprasert, Appl. Surf. Sci. 487 (2019) 1233-1244. doi: 10.1016/j.apsusc.2019.04.269

    13. [13]

      I. Costas-Mora, V. Romero, I. Lavilla, C. Bendicho, Anal. Chem. 86 (2014) 4536-4543. doi: 10.1021/ac500517h

    14. [14]

      L. Qiao, J. Wang, M. Zheng, Z. Xie, Anal. Methods 10 (2018) 1863-1869. doi: 10.1039/C8AY00320C

    15. [15]

      Z. Guo, Y. Jiao, F. Du, et al., Talanta 216 (2020) 120943.

    16. [16]

      Y. Yang, W. Kong, H. Li, et al., ACS Appl. Mater. Interfaces 7 (2015) 27324-27330. doi: 10.1021/acsami.5b08782

    17. [17]

      S. Kalytchuk, K. Polakova, Y. Wang, et al., ACS Nano 11 (2017) 1432-1442. doi: 10.1021/acsnano.6b06670

    18. [18]

      Z. Mu, J. Hua, Y. Yang, Spectrochim. Acta A 224 (2020) 117444.

    19. [19]

      N. Wang, Y. Wang, T. Guo, et al., Biosens. Bioelectron. 85 (2016) 68-75. doi: 10.1016/j.bios.2016.04.089

    20. [20]

      R. Wang, Y. Xu, T. Zhang, Y. Jiang, Anal. Methods 7 (2015) 1701-1706. doi: 10.1039/C4AY02880E

    21. [21]

      X.W. Hua, Y.W. Bao, H.Y. Wang, Z. Chen, F.G. Wu, Nanoscale 9 (2017) 2150-2161. doi: 10.1039/C6NR06558A

    22. [22]

      F. Lin, C. Li, L. Dong, D. Fu, Z. Chen, Nanoscale 9 (2017) 9056-9064. doi: 10.1039/C7NR01975K

    23. [23]

      X. Gao, Y. Cui, R.M. Levenson, L.W.K. Chung, S. Nie, Nat. Biotechnol. 22 (2004) 969-976. doi: 10.1038/nbt994

    24. [24]

      A.M. Derfus, W.C.W. Chan, S.N. Bhatia, Nano Lett. 4 (2004) 11-18. doi: 10.1021/nl0347334

    25. [25]

      A.D. Yoffe, Adv. Phys. 50 (2001) 1-208.

    26. [26]

      M.C. Daniel, D. Astruc, Chem. Rev. 104 (2004) 293-346. doi: 10.1021/cr030698+

    27. [27]

      K.L. Kelly, E. Coronado, L.L. Zhao, G.C. Schatz, J. Phys. Chem. B 107 (2003) 668-677.

    28. [28]

      M.V. Wolkin, J. Jorne, P.M. Fauchet, G. Allan, C. Delerue, Phys. Rev. Lett. 82 (1999) 197-200. doi: 10.1103/PhysRevLett.82.197

    29. [29]

      M.S. Mauter, M. Elimelech, Environ. Sci. Technol. 42 (2008) 5843-5859. doi: 10.1021/es8006904

    30. [30]

      A. Magrez, S. Kasas, V. Salicio, et al., Nano Lett. 6 (2006) 1121-1125. doi: 10.1021/nl060162e

    31. [31]

      C. Cha, S.R. Shin, N. Annabi, M.R. Dokmeci, A. Khademhosseini, ACS Nano 7 (2013) 2891-2897. doi: 10.1021/nn401196a

    32. [32]

      S. Lu, L. Sui, J. Liu, et al., Adv Mater. 29 (2017) 1603443.

    33. [33]

      W. Li, Y. Liu, B. Wang, et al., Chin. Chem. Lett. 30 (2019) 2323-2327. doi: 10.1016/j.cclet.2019.06.040

    34. [34]

      C. Ji, Y. Zhou, R.M. Leblanc, Z. Peng, ACS Sens. 5 (2020) 2724-2741. doi: 10.1021/acssensors.0c01556

    35. [35]

      B. Wang, J. Li, Z. Tang, B. Yang, S. Lu, Sci. Bull. 64 (2019) 1285-1292. doi: 10.1016/j.scib.2019.07.021

    36. [36]

      Y. Wang, G. Guo, J. Gao, et al., Chem. Mater. 32 (2020) 8146-8157. doi: 10.1021/acs.chemmater.0c01391

    37. [37]

      Y. Han, B. Tang, L. Wang, et al., ACS Nano 14 (2020) 14761-14768. doi: 10.1021/acsnano.0c01899

    38. [38]

      S. Chen, T. Sun, M. Zheng, Z. Xie, Adv. Funct. Mater. 30 (2020) 2004680.

    39. [39]

      X. Xu, R. Ray, Y. Gu, et al., J. Am. Chem. Soc. 126 (2004) 12736-12737. doi: 10.1021/ja040082h

    40. [40]

      M. Li, T. Chen, J.J. Gooding, J. Liu, ACS Sens. 4 (2019) 1732-1748. doi: 10.1021/acssensors.9b00514

    41. [41]

      Z. Peng, X. Han, S. Li, et al., Coordin. Chem. Rev. 343 (2017) 256-277. doi: 10.1016/j.ccr.2017.06.001

    42. [42]

      M.L. Liu, B.B. Chen, C.M. Li, C.Z. Huang, Green Chem. 21 (2019) 449-471. doi: 10.1039/C8GC02736F

    43. [43]

      V. Sharma, P. Tiwari, S.M. Mobin, J. Mater. Chem. B 5 (2017) 8904-8924. doi: 10.1039/C7TB02484C

    44. [44]

      Y. Zhou, K.J. Mintz, S.K. Sharma, R.M. Leblanc, Langmuir 35 (2019) 9115-9132. doi: 10.1021/acs.langmuir.9b00595

    45. [45]

      J. Lan, C. Liu, M. Gao, C. Huang, Talanta 144 (2015) 93-97. doi: 10.1016/j.talanta.2015.05.071

    46. [46]

      H. Liu, T. Ye, C. Mao, Angew. Chem. Int. Ed. 46 (2007) 6473-6475. doi: 10.1002/anie.200701271

    47. [47]

      L. Zheng, Y. Chi, Y. Dong, J. Lin, B. Wang, J. Am. Chem. Soc. 131 (2009) 4564-4565. doi: 10.1021/ja809073f

    48. [48]

      F. Li, Y. Li, X. Yang, et al., Angew. Chem. Int. Ed. 57 (2018) 2377-2382. doi: 10.1002/anie.201712453

    49. [49]

      L. Ðorđević, F. Arcudi, A. D'Urso, et al., Nat. Commun. 9 (2018) 3442.

    50. [50]

      Y. Su, Z. Xie, M. Zheng, J. Colloid Interface Sci. 573 (2020) 241-249. doi: 10.1016/j.jcis.2020.04.004

    51. [51]

      W. Meng, X. Bai, B. Wang, et al., Energy Environ. Mater. 2 (2019) 172-192. doi: 10.1002/eem2.12038

    52. [52]

      J. Zhang, Y. Yuan, G. Liang, S.H. Yu, Adv. Sci. 2 (2015) 1500002.

    53. [53]

      S. Sahu, B. Behera, T.K. Maiti, S. Mohapatra, Chem. Commun. 48 (2012) 8835-8837. doi: 10.1039/c2cc33796g

    54. [54]

      Y.P. Sun, B. Zhou, Y. Lin, et al., J. Am. Chem. Soc. 128 (2006) 7756-7757. doi: 10.1021/ja062677d

    55. [55]

      Z.A. Qiao, Y. Wang, Y. Gao, et al., Chem. Commun. 46 (2010) 8812-8814. doi: 10.1039/c0cc02724c

    56. [56]

      H. Li, X. He, Z. Kang, et al., Angew. Chem. Int. Ed. 49 (2010) 4430-4434. doi: 10.1002/anie.200906154

    57. [57]

      Z.C. Yang, M. Wang, A.M. Yong, et al., Chem. Commun. 47 (2011) 11615-11617. doi: 10.1039/c1cc14860e

    58. [58]

      M. Zheng, Y. Li, S. Liu, et al., ACS Appl. Mater. Interfaces 8 (2016) 23533-23541. doi: 10.1021/acsami.6b07453

    59. [59]

      J. Wang, Y. Yang, G. Sun, M. Zheng, Z. Xie, Environ. Res. 177 (2019) 108621.

    60. [60]

      H. Zhu, X. Wang, Y. Li, et al., Chem. Commun. (2009) 5118-5120.

    61. [61]

      L. Qiao, T. Sun, X. Zheng, M. Zheng, Z. Xie, Mater. Sci. Eng. C 85 (2018) 1-6. doi: 10.1016/j.msec.2017.12.011

    62. [62]

      R.S. Li, P.F. Gao, H.Z. Zhang, et al., Chem. Sci. 8 (2017) 6829-6835. doi: 10.1039/C7SC01316G

    63. [63]

      M. Zheng, S. Ruan, S. Liu, et al., ACS Nano 9 (2015) 11455-11461. doi: 10.1021/acsnano.5b05575

    64. [64]

      H. He, X. Zheng, S. Liu, et al., Nanoscale 10 (2018) 10991-10998. doi: 10.1039/C8NR02643B

    65. [65]

      Y. Li, X. Zheng, X. Zhang, et al., Adv. Health Mater. 6 (2017) 1600924.

    66. [66]

      S. Miao, K. Liang, J. Zhu, et al., Nano Today 33 (2020) 100879.

    67. [67]

      A. Cai, Q. Wang, Y. Chang, X. Wang, J. Alloys Compd. 692 (2017) 183-189. doi: 10.1016/j.jallcom.2016.09.030

    68. [68]

      S. Chandra, D. Laha, A. Pramanik, et al., Luminescence 31 (2016) 81-87. doi: 10.1002/bio.2927

    69. [69]

      Y. Liu, W. Duan, W. Song, et al., ACS Appl. Mater. Interfaces 9 (2017) 12663-12672. doi: 10.1021/acsami.6b15746

    70. [70]

      P. Gao, S. Liu, Y. Su, M. Zheng, Z. Xie, Bioconjug. Chem. 31 (2020) 646-655. doi: 10.1021/acs.bioconjchem.9b00801

    71. [71]

      X. Li, S.P. Lau, L. Tang, et al., J. Mater. Chem. C 1 (2013) 7308-7313. doi: 10.1039/c3tc31473a

    72. [72]

      W.S. Zou, W.L. Kong, Q.C. Zhao, et al., Microchim. Acta 186 (2019) 576.

    73. [73]

      Z. Qian, X. Shan, L. Chai, et al., ACS Appl. Mater. Interfaces 6 (2014) 6797-6805. doi: 10.1021/am500403n

    74. [74]

      F. Li, T. Li, C. Sun, et al., Angew. Chem. Int. Ed. 56 (2017) 9910-9914. doi: 10.1002/anie.201705989

    75. [75]

      Y. Liu, P. Wu, X. Wu, et al., Talanta 210 (2020) 120649.

    76. [76]

      S. Sun, Q. Guan, Y. Liu, et al., Chin. Chem. Lett. 30 (2019) 1051-1054. doi: 10.1016/j.cclet.2019.01.014

    77. [77]

      Z. Ji, P. Ai, C. Shao, et al., ACS Biomater. Sci. Eng. 4 (2018) 2089-2094. doi: 10.1021/acsbiomaterials.7b01008

    78. [78]

      X. Xi, X. Peng, C. Xiong, et al., Microchim. Acta 187 (2020) 383.

    79. [79]

      F. Ayaz, M. Ö. Alaş, M. Oǧuz, R. Genç, Mol. Biol. Rep. 46 (2019) 2405-2415. doi: 10.1007/s11033-019-04701-1

    80. [80]

      Q. Xu, Y. Liu, R. Su, et al., Nanoscale 8 (2016) 17919-17927. doi: 10.1039/C6NR05434J

    81. [81]

      H.X. Wang, J. Xiao, Z. Yang, et al., J. Mater. Chem. A 3 (2015) 11287-11293. doi: 10.1039/C5TA02057C

    82. [82]

      L. Li, X. Wang, Z. Fu, F. Cui, Mater. Lett. 196 (2017) 300-303. doi: 10.1016/j.matlet.2017.03.112

    83. [83]

      D. Sun, R. Ban, P.H. Zhang, et al., Carbon 64 (2013) 424-434.

    84. [84]

      Y. Wang, S.H. Kim, L. Feng, Anal. Chem. Acta 890 (2015) 134-142. doi: 10.1016/j.aca.2015.07.051

    85. [85]

      D. Qu, M. Zheng, L. Zhang, et al., Sci. Rep. 4 (2014) 5294.

    86. [86]

      Q. Xu, P. Pu, J. Zhao, et al., J. Mater. Chem. A 3 (2015) 542-546. doi: 10.1039/C4TA05483K

    87. [87]

      Y. Dong, H. Pang, H.B. Yang, et al., Angew. Chem. Int. Ed. 52 (2013) 7800-7804. doi: 10.1002/anie.201301114

    88. [88]

      H. Ding, J.S. Wei, H.M. Xiong, Nanoscale 6 (2014) 13817-13823. doi: 10.1039/C4NR04267K

    89. [89]

      H. Peng, J. Travas-Sejdic, Chem. Mater. 21 (2009) 5563-5565. doi: 10.1021/cm901593y

    90. [90]

      J. Zhang, W. Shen, D. Pan, et al., N. J. Chem. 34 (2010) 591-593. doi: 10.1039/b9nj00662a

    91. [91]

      H. Gonçalves, J.C.G. Esteves da Silva, J. Fluoresc. 20 (2010) 1023-1028. doi: 10.1007/s10895-010-0652-y

    92. [92]

      A. Hao, X. Guo, Q. Wu, et al., J. Lumin. 170 (2015) 90-96.

    93. [93]

      M. Zheng, S. Liu, J. Li, et al., Adv. Mater. 26 (2014) 3554-3560. doi: 10.1002/adma.201306192

    94. [94]

      M. Zheng, L. Qiao, Y. Su, P. Gao, Z. Xie, J. Mater. Chem. B 7 (2019) 3840-3845. doi: 10.1039/C9TB00544G

    95. [95]

      Y. Su, S. Lu, P. Gao, M. Zheng, Z. Xie, Mater. Chem. Front. 3 (2019) 1747-1753. doi: 10.1039/C9QM00257J

    96. [96]

      J. Zhang, M. Zheng, F. Zhang, et al., Chem. Mater. 28 (2016) 8825-8833. doi: 10.1021/acs.chemmater.6b04894

    97. [97]

      J. Zhang, M. Zheng, Z. Xie, J. Mater. Chem. B 4 (2016) 5659-5663. doi: 10.1039/C6TB01622G

    98. [98]

      Y. Li, W. Liu, C. Sun, et al., J. Colloid Interface Sci. 523 (2018) 226-233. doi: 10.1016/j.jcis.2018.03.108

    99. [99]

      Z. Wang, P. Long, Y. Feng, C. Qin, W. Feng, RSC Adv. 7 (2017) 2810-2816. doi: 10.1039/C6RA25465A

    100. [100]

      K. Jiang, S. Sun, L. Zhang, et al., Angew. Chem. Int. Ed. 54 (2015) 5360-5363. doi: 10.1002/anie.201501193

    101. [101]

      F. Arcudi, L. Dordevic, M. Prato, Angew. Chem. Int. Ed. 55 (2016) 2107-2112. doi: 10.1002/anie.201510158

    102. [102]

      C.L. Shen, Q. Lou, K.K. Liu, L. Dong, C.X. Shan, Nano Today 35 (2020) 100954.

    103. [103]

      M.A. Sk, A. Ananthanarayanan, L. Huang, K.H. Lim, P. Chen, J. Mater. Chem. C 2 (2014) 6954-6960. doi: 10.1039/C4TC01191K

    104. [104]

      H. Ding, J.S. Wei, P. Zhang, et al., Small 14 (2018) 1800612.

    105. [105]

      J.B. Essner, J.A. Kist, L. Polo-Parada, G.A. Baker, Chem. Mater. 30 (2018) 1878-1887. doi: 10.1021/acs.chemmater.7b04446

    106. [106]

      J. Schneider, C.J. Reckmeier, Y. Xiong, et al., J. Phys. Chem. C 121 (2017) 2014-2022.

    107. [107]

      Y. Song, S. Zhu, S. Zhang, et al., J. Mater. Chem. C 3 (2015) 5976-5984. doi: 10.1039/C5TC00813A

    108. [108]

      J. Zhang, L. Yang, Y. Yuan, J. Jiang, S.H. Yu, Chem. Mater. 28 (2016) 4367-4374. doi: 10.1021/acs.chemmater.6b01360

    109. [109]

      L. Bao, C. Liu, Z.L. Zhang, D.W. Pang, Adv. Mater. 27 (2015) 1663-1667. doi: 10.1002/adma.201405070

    110. [110]

      H. Ding, S.B. Yu, J.S. Wei, H.M. Xiong, ACS Nano 10 (2016) 484-491. doi: 10.1021/acsnano.5b05406

    111. [111]

      X. Li, S. Zhang, S.A. Kulinich, Y. Liu, H. Zeng, Sci. Rep. 4 (2014) 4976.

    112. [112]

      S.H. Jin, D.H. Kim, G.H. Jun, S.H. Hong, S. Jeon, ACS Nano 7 (2013) 1239-1245. doi: 10.1021/nn304675g

    113. [113]

      R. Genc, M.O. Alas, E. Harputlu, et al., Sci. Rep. 7 (2017) 11222.

    114. [114]

      T. Zhang, J. Zhu, Y. Zhai, et al., Nanoscale 9 (2017) 13042-13051. doi: 10.1039/C7NR03570E

    115. [115]

      M. Zheng, Y. Li, Y. Zhang, Z. Xie, RSC Adv. 6 (2016) 83501-83504. doi: 10.1039/C6RA16055G

    116. [116]

      S. Zhu, L. Wang, N. Zhou, et al., Chem. Commun. 50 (2014) 13845-13848. doi: 10.1039/C4CC05806B

    117. [117]

      L. Ai, Y. Yang, B. Wang, et al., Sci. Bull. 66 (2021) 839-856. doi: 10.1016/j.scib.2020.12.015

    118. [118]

      S.L. Shen, X.Q. Huang, Y.Y. Zhang, et al., Sens. Actuators B: Chem. 263 (2018) 252-257. doi: 10.1016/j.snb.2018.02.121

    119. [119]

      Y. Zhang, S. Li, Z. Zhao, Anal. Chem. 88 (2016) 12380-12385. doi: 10.1021/acs.analchem.6b03632

    120. [120]

      R. Gotor, P. Ashokkumar, M. Hecht, K. Keil, K. Rurack, Anal. Chem. 89 (2017) 8437-8444. doi: 10.1021/acs.analchem.7b01903

    121. [121]

      Z. Wei, H. Li, S. Liu, et al., Anal. Chem. 91 (2019) 15477-15483. doi: 10.1021/acs.analchem.9b03272

    122. [122]

      J. Liu, Y. Chen, L. Wang, et al., J. Agric. Food Chem. 67 (2019) 3826-3836. doi: 10.1021/acs.jafc.9b00024

    123. [123]

      W. Lu, Y. Jiao, Y. Gao, et al., ACS Appl. Mater. Interfaces 10 (2018) 42915-42924. doi: 10.1021/acsami.8b16710

    124. [124]

      Y. Hu, L. Zhang, X. Li, et al., ACS Sustain. Chem. Eng. 5 (2017) 4992-5000. doi: 10.1021/acssuschemeng.7b00393

    125. [125]

      W. Lu, Y. Gao, Y. Jiao, et al., Nanoscale 9 (2017) 11545-11552. doi: 10.1039/C7NR02336G

    126. [126]

      P. Gao, Z. Xie, M. Zheng, Sens. Actuators B: Chem. 319 (2020) 128265.

    127. [127]

      Y. Ma, A.Y. Chen, Y.Y. Huang, et al., Carbon 162 (2020) 234-244. doi: 10.1016/j.carbon.2020.02.048

    128. [128]

      Q. Hu, Y. Pan, X. Gong, et al., Food Chem. 308 (2020) 125590.

    129. [129]

      Z. Liu, M. Chen, Y. Guo, et al., Chem. Eng. J. 384 (2020) 123260.

    130. [130]

      Y. Song, C. Zhu, J. Song, et al., ACS Appl. Mater. Interfaces 9 (2017) 7399-7405. doi: 10.1021/acsami.6b13954

    131. [131]

      Z. Liu, W. Jin, F. Wang, et al., Sens. Actuators B: Chem. 296 (2019) 126698.

    132. [132]

      J. Yue, L. Li, L. Cao, et al., ACS Appl. Mater. Interfaces 11 (2019) 44566-44572. doi: 10.1021/acsami.9b13737

    133. [133]

      Z. Chen, S. Wang, X. Yang, J. Photoch. Photobiol. A: Chem. 365 (2018) 178-184. doi: 10.1016/j.jphotochem.2018.08.001

    134. [134]

      S. Sun, K. Jiang, S. Qian, Y. Wang, H. Lin, Anal. Chem. 89 (2017) 5542-5548. doi: 10.1021/acs.analchem.7b00602

    135. [135]

      J. Wang, Y. Yang, G. Sun, M. Zheng, Z. Xie, Environ. Res. 177 (2019) 108621.

    136. [136]

      J. Kudr, L. Richtera, K. Xhaxhiu, et al., Biosens. Bioelectron. 92 (2017) 133-139. doi: 10.1016/j.bios.2017.01.067

    137. [137]

      Y. Wang, K. Jiang, J. Zhu, L. Zhang, H. Lin, Chem. Commun. 51 (2015) 12748-12751. doi: 10.1039/C5CC04905A

    138. [138]

      M. Zheng, Z. Xie, D. Qu, et al., ACS Appl. Mater. Interfaces 5 (2013) 13242-13247. doi: 10.1021/am4042355

    139. [139]

      S. Huang, E. Yang, J. Yao, et al., ACS Omega 4 (2019) 9333-9342. doi: 10.1021/acsomega.9b00874

    140. [140]

      X. Wu, S. Sun, Y. Wang, et al., Biosens. Bioelectron. 90 (2017) 501-507. doi: 10.1016/j.bios.2016.10.060

    141. [141]

      H. Li, D. Su, H. Gao, et al., Anal. Chem. 92 (2020) 3198-3205. doi: 10.1021/acs.analchem.9b04917

    142. [142]

      J. He, H. Zhang, J. Zou, et al., Biosens. Bioelectron. 79 (2016) 531-535. doi: 10.1016/j.bios.2015.12.084

    143. [143]

      M. Luo, Y. Hua, Y. Liang, et al., Biosens. Bioelectron. 98 (2017) 195-201. doi: 10.1016/j.bios.2017.06.056

    144. [144]

      S. Pawar, U.K. Togiti, A. Bhattacharya, A. Nag, ACS Omega 4 (2019) 11301-11311. doi: 10.1021/acsomega.9b01208

    145. [145]

      M. Lan, Y. Di, X. Zhu, et al., Chem. Commun. 51 (2015) 15574-15577. doi: 10.1039/C5CC05835J

    146. [146]

      Y. Xu, X. Chen, R. Chai, et al., Nanoscale 8 (2016) 13414-13421. doi: 10.1039/C6NR03129C

    147. [147]

      Z. Gao, L. Wang, R. Su, et al., Biosens. Bioelectron. 70 (2015) 232-238. doi: 10.1016/j.bios.2015.03.043

    148. [148]

      Y. Hu, Z. Gao, J. Yang, H. Chen, L. Han, J. Colloid Interface Sci. 538 (2019) 481-488. doi: 10.1016/j.jcis.2018.12.016

    149. [149]

      P. Gao, J. Wang, M. Zheng, Z. Xie, Chem. Eng. J. 381 (2020) 122665.

    150. [150]

      G. Gao, Y.W. Jiang, J. Yang, F.G. Wu, Nanoscale 9 (2017) 18368-18378. doi: 10.1039/C7NR06764J

    151. [151]

      X.W. Hua, Y.W. Bao, Z. Chen, F.G. Wu, Nanoscale 9 (2017) 10948-10960. doi: 10.1039/C7NR03658B

    152. [152]

      X.W. Hua, Y.W. Bao, F.G. Wu, ACS Appl. Mater. Interfaces 10 (2018) 10664-10677. doi: 10.1021/acsami.7b19549

    153. [153]

      X.W. Hua, Y.W. Bao, J. Zeng, F.G. Wu, ACS Appl. Mater. Interfaces 11 (2019) 32647-32658. doi: 10.1021/acsami.9b09590

    154. [154]

      X. Chen, X. Zhang, F.G. Wu, Chin. Chem. Lett. 32 (2021) 3048-3052. doi: 10.1016/j.cclet.2021.03.061

    155. [155]

      G. Gao, Y.W. Jiang, W. Sun, F.G. Wu, Chin. Chem. Lett. 29 (2018) 1475-1485. doi: 10.1016/j.cclet.2018.07.004

    156. [156]

      J. Yang, G. Gao, X. Zhang, et al., Carbon 146 (2019) 827-839. doi: 10.1016/j.carbon.2019.02.040

    157. [157]

      J. Yang, X. Zhang, Y.H. Ma, et al., ACS Appl. Mater. Interfaces 8 (2016) 32170-32181. doi: 10.1021/acsami.6b10398

    158. [158]

      Y. Song, H. Li, F. Lu, et al., J. Mater. Chem. B 5 (2017) 6008-6015. doi: 10.1039/C7TB01092C

    159. [159]

      M. Wang, Y. Xia, J. Qiu, X. Ren, Spectrochim Acta A: Mol. Biomol. Spectrosc. 206 (2019) 170-176. doi: 10.1016/j.saa.2018.08.006

  • Figure 1  Top-down and bottom-up synthetic methods of CDs.

    Figure 2  Top-down and bottom-up synthetic methods of CDs.

    Figure 3  (a) Schematic diagram of improving the photoluminescence of CDs by passivation. Copied with permission [89]. Copyright 2009, American Chemical Society. (b) Synthetic scheme for CD-Oxa and its applications in theranostics. Copied with permission [93]. Copyright 2014, Wiley. (c) Schematic syntheses and applications of CD-B, CD-V and CD-A. Copied with permission [94]. Copyright 2019, Royal Society of Chemistry. (d) Fabrication of FCD-DOX and FCD-BODIPY and the cellular uptake of FCD-DOX. Copied with permission [70]. Copyright 2019, American chemical society. (e) Schematic illustration of CDsG-AIE preparation and cellular imaging. Copied with permission [96]. Copyright 2016, American Chemical Society.

    Figure 4  (a) Preparation of CDs and luminescent mechanism diagram. (b) The theoretically simulated optimal configuration of M1 and (c) the HOMO and LUMO energy level arrangement about its ground and excited states. Copied with permission [108]. Copyright 2016, American Chemical Society.

    Figure 5  (a) Preparation of CDs and luminescent mechanism diagram. (b) The theoretically simulated optimal configuration of M1 and (c) the HOMO and LUMO energy level arrangement about its ground and excited states. Copied with permission [108]. Copyright 2016, American Chemical Society.

    Figure 6  The emission of CDs from passivated surface by PEG1500N. Copied with permission [54]. Copyright 2006, American Chemical Society.

    Figure 7  (a) Schematic illustrating the sensing mechanism of Y-CDs to FQs. (b) UV-vis absorption. (c) Fluorescence decay curves of Y-CDs before and after adding FQs. (d) The quenching efficiency at 25 ℃ and 37 ℃. Copied with permission [123]. Copyright 2018, American Chemical Society. (e) Schematic illustration for fabricating chiral CDs-based nanoprobes for assaying Sn2+. (f) Fluorescence decay curves of chiral CDs before and after adding Sn2+. Copied with permission [126]. Copyright 2020, Elsevier. (g) Schematic representation of the sensing process of CD to Fe3+. Copied with permission [122]. Copyright 2018, American Chemical Society.

    Figure 8  (a) Preparation and schematic illustration of N, S, Cl-CDs-based nanosensors for detection of [Fe(CN)6]4− in food samples. (b) The photoluminescence spectra of the N, S, Cl-CDs-based sensing system containing various concentrations of [Fe(CN)6]4−. (c) The fluorescence lifetime of N, S, Cl-CDs before and after adding [Fe(CN)6]4−. Copied with permission [128]. Copyright 2020, Elsevier. (d) Mechanism diagram of N-CDs coordinated with Fe3+ ions. Copied with permission [129]. Copyright 2020, Elsevier. (e) Images of cells incubated with different concentrations of NCDs/Fe3+ (f) The fluorescence lifetime of NCDs before and after adding Fe3+. Copied with permission [130]. Copyright 2017, American Chemical Society. (g) Schematic illustration of CDs for the detection of Pb2+. Copied with permission [131]. Copyright 2019, Elsevier.

    Figure 9  (a) Schematic diagram of Ca2+ ions induce CDs aggregation. The hydrodynamic diameters of CDs before (b) and after (c) the addition of Ca2+ ions. Copied with permission [132]. Copyright 2019, American Chemical Society. (d) The hydrodynamic diameters of CDs and CDs+Au3+. Copied with permission [133]. Copyright 2018, Elsevier. (e) Schematic illustration of CDs for detection of Ce3+, Fe3+ and Cu2+. TEM images of CDs with the addition of (f) Ce3+, (g) Fe3+, and (h) Cu2+. Copied with permission [134]. Copyright 2018, American Chemical Society

    Figure 10  (a) Schematic diagram for synthesizing amphiphilic fluorescent CDs and their application in TNP detection. (b) Fluorescence quenching efficiency of CDs upon the addition different nitroaromatics. (c) Absorption spectra of various nitroaromatic compounds and the emission spectrum of CDs. Copied with permission [135]. Copyright 2019, Elsevier.

    Figure 11  (a) Schematic illustration of fluorescence assays of CDs for the detection of Cr2O72− based on IFE. (b) Fluorescence excitation (red solid line) and emission (blue solid line) spectra of CDs and UV-vis absorption spectrum of Cr2O72− (black solid line). Copied with permission [138]. Copyright 2013, American Chemical Society.

    Figure 12  (a) Schematic illustration of preparation of CDs and a suggested PET mechanism of fluorescence quenching of C-dots-TPP in the presence of ONOO. Copied with permission [140]. Copyright 2017, Elsevier. (b) Synthesis route of CDs and schematic illustration of h-CDs/DA-based sensor for OPs detection. Copied with permission [141]. Copyright 2020, American Chemical Society. (c) Schematic of CDs as sensors for the sensing process of Hg2+ ions. Copied with permission [142]. Copyright 2016, Elsevier.

    Figure 13  (a) Reversible sensing performance of RD-CDs for ClO and AA in colorimetric and fluorescence modes. Copied with permission [121]. Copyright 2019, American Chemical Society. (b) Schematic illustration of fluorescence assays for AA based on reduction reaction of CD-Cr(VI). Copied with permission [138]. Copyright 2013, American Chemical Society.

    Figure 14  (a) Synthetic scheme for CDs and their application as pH probes. Copied with permission [14]. Copyright 2018, Royal Society of Chemistry. (b) Schematic illustrating the fabrication of CDs, their lysosome targeting ability and applications for probing pH changes in vitro and in vivo. Copied with permission [149]. Copyright 2020, Elsevier. (c) Confocal microscopy images of N, S, I-CDs-colon cancer cell HT-29 with corresponding fluorescence field at 15, 25, 35 and 15 ℃, respectively. Copied with permission [17]. Copyright 2020, Elsevier. (d) Schematic mechanism for the temperature-dependent fluorescence intensity of N-CDs. Copied with permission [16]. Copyright 2015, American Chemical Society. (e) The linear calibration of fluorescence intensities for the W-CDs at 1.17 mg/mL in the presence of various amount of E. coli O157:H7 within a concentration range of 105–108 cfu/mL. Copied with permission [19]. Copyright 2016, Elsevier.

    Figure 15  (a) The reversible emission color changes of CDs indicator paper under UV lamp (365 nm). (b) Photographs of the paper sensor printed with CDs after exposed to the VOCs under the UV lamp (365 nm). Copied with permission [115]. Copyright 2016, Royal Society of Chemistry.

    Table 1.  Summary of representative CDs used in various sensing applications.

    下载: 导出CSV
  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  461
  • HTML全文浏览量:  87
文章相关
  • 发布日期:  2022-04-15
  • 收稿日期:  2021-05-14
  • 接受日期:  2021-09-23
  • 修回日期:  2021-09-23
  • 网络出版日期:  2021-09-30
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章