Next Article in Journal
Highly Active Amino-Fullerene Derivative-Modified TiO2 for Enhancing Formaldehyde Degradation Efficiency under Solar-Light Irradiation
Next Article in Special Issue
Semiconductor-to-Insulator Transition in Inter-Electrode Bridge-like Ensembles of Anatase Nanoparticles under a Long-Term Action of the Direct Current
Previous Article in Journal
Correction: Xie et al. Synergistic Effect of MoS2 and SiO2 Nanoparticles as Lubricant Additives for Magnesium Alloy–Steel Contacts. Nanomaterials 2017, 7, 154
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review on Carbon Dots: Synthesis, Characterization and Its Application in Optical Sensor for Environmental Monitoring

by
Nur Alia Sheh Omar
1,2,
Yap Wing Fen
1,2,*,
Ramli Irmawati
1,
Hazwani Suhaila Hashim
1,
Nur Syahira Md Ramdzan
1 and
Nurul Illya Muhamad Fauzi
2
1
Faculty of Science, Universiti Putra Malaysia (UPM), Serdang 43400, Selangor, Malaysia
2
Institute of Nanoscience and Nanotechnology, Universiti Putra Malaysia (UPM), Serdang 43400, Selangor, Malaysia
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(14), 2365; https://doi.org/10.3390/nano12142365
Submission received: 9 May 2022 / Revised: 11 June 2022 / Accepted: 15 June 2022 / Published: 11 July 2022
(This article belongs to the Special Issue Properties and Potential Applications of Nanoparticles)

Abstract

:
The development of carbon dots (CDs), either using green or chemical precursors, has inevitably led to their wide range application, from bioimaging to optoelectronic devices. The reported precursors and properties of these CDs have opened new opportunities for the future development of high-quality CDs and applications. Green precursors were classified into fruits, vegetables, flowers, leaves, seeds, stem, crop residues, fungi/bacteria species, and waste products, while the chemical precursors were classified into acid reagents and non-acid reagents. This paper quickly reviews ten years of the synthesis of CDs using green and chemical precursors. The application of CDs as sensing materials in optical sensor techniques for environmental monitoring, including the detection of heavy metal ions, phenol, pesticides, and nitroaromatic explosives, was also discussed in this review. This profound review will offer knowledge for the upcoming community of researchers interested in synthesizing high-quality CDs for various applications.

1. Introduction

Fluorescent carbon dots (CDs), also known as carbon quantum dots (CQDs) or carbon nanodots (CNs), have drawn a great deal of attention in recent years, owing to their high photostability, excellent water solubility, tuneable fluorescence and optical properties, low toxicity, good biocompatibility, and environmental friendliness [1]. To date, a variety of precursors have been utilized to prepare CDs through “bottom-up” or “top-down” approaches [2]. Among these approaches, hydrothermal/carbonization treatment is frequently applied for the preparation CDs because of the outstanding advantages, such as high yield, simple manipulation, easy control, uniform products, lower air pollution, low energy consumption and so on [3]. Despite these advantages, CDs cannot be produced without the presence of starting materials, also known as precursors. In this regard, the development of green and chemical synthesis methods for producing high fluorescent CDs has gathered the focus of researchers. The green synthesis methods by means of green precursors of synthesis involves the usage of inexpensive or recycled materials, while the chemical synthesis methods involve toxic chemical reagents or organic solvents as precursors. CDs synthesized using these methods usually contain a certain amount of oxygen, hydrogen, and nitrogen. With all these options to prepare CDs under certain experimental conditions, it is no surprise that differential fluorescence properties of CDs were easily acquired. In addition, surface passivation (generally involves functional groups such as amine groups and hydroxyl groups) and the heteroatom doping (n-type doping: nitrogen, phosphorus, sulfur, and chlorine or p-type doping: boron) approach is another factor that paves the way to efficiently improve the fluorescent properties, quantum yield, and other physicochemical properties of CDs [4]. Such intriguing properties have triggered new capabilities in the field of sensing and bioimaging.
To date, many review articles on the preparation of CDs, heteroatom doped CDs, and their applications have been published [5,6,7,8]. However, there is no comprehensive review exploring various green and chemical precursors as the carbon source in CDs and their application in environmental monitoring. The purpose of this review is, therefore, to update and organize the green and chemical precursors used in CD preparation. The final section will give a concise overview of their application in the optical sensing of heavy metal ions, phenol, pesticides, and nitroaromatic explosives.

2. CDs Synthesized from Green Precursors

Carbon dots (CDs) can be synthesized using two precursors, either from green or chemical sources. As of now, researchers focus more on simple, low-cost, and greenway synthesis for the large-scale production of high-quality CDs. Numerous green sources, such as fruits, vegetables, flowers, leaves, seeds, stems, crop residues, fungi/bacteria species and waste products, have been used as a carbon source for the preparation of CDs. In this section, all green sources used for the synthesis of CDs are briefly arranged and compared.

2.1. Fruits

Orange juice and watermelon peels were among the very first precursors used as a carbon source to synthesize CDs. The studies reveal the as-prepared CDs have different fluorescence properties, due to the presence of different chemical groups in the raw materials and synthesis method. The synthesized CDs from orange juice showed the quantum yield of 26% at an emission of 441 nm, whereas a 7.1% quantum yield was demonstrated by watermelon peel-derived CDs at an emission of 490–580 nm [9,10]. In 2014, Du et al. reported renewable wastes of sugar cane bagasse as a new precursor for fluorescent CDs [11]. It, thus, demonstrated that such bagasse-derived CDs could function as highly effective fluorescent sensing probes for labeling and imaging in biomedical applications.
Lemon peels, prunus avium extract, cornstalk, corn bract, dried lemon peels, pulp-free lemon juice, citrus lemon peels, citrus sinensis peels, etc. are another precursor that have been used in the development of CDs, as shown in Table 1 [12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41]. Among these precursors, acidic fruits, such as lemon peels, lemon juice, and citrus sinensis, were frequently chosen. This is because the juice extract is rich in sucrose, glucose, fructose, citric acid, and ascorbic acid, while the peels are mainly composed of proteins, fibers, and less of oils and antioxidants. Consequently, CDs from juice extract exhibit higher fluorescence properties than peels, due to the high acid and sugar contents that provide a considerable amount of carbon and hydrogen elements. A comparison study of the different precursors used in hydrothermal synthesis was then carried out to understand the role of citric acid [30]. It was found that lemon juice has higher photoluminescence (PL) emission than ripe lemon juice and orange juice. This result is because lemon juices have higher concentrations of citric acid than orange juices. Moreover, the decrease in PL emission of ripe lemon juice was due to a significant decrease in the grade of constituents and destroyed their surface structure.
Instead of just relying on carbon sources to produce high quantum yields of CDs, it has been found that by passivating CDs with amine-containing molecules, such as ethylenediamine, one can increase the quantum yield and selectively sensing analytes. Figure 1 shows the hydrothermal reaction of CDs from citrus lemon juice in the presence of ethylenediamine to produce nitrogen-doped CDs. The results showed that the prepared nitrogen-doped CDs had the quantum yield of 31% under bright blue emission [28]. However, there are limits to these blue emissive CDs, especially in the field of bioimaging because of strong tissue autofluorescence at low wavelength emissions. Herein, Ding et al. (2017) have heated an ethanol solution of pulp-free lemon juice to produce very high red-luminescent CDs, which hold the promise for in vitro and in vivo bio-imaging [17].
At present, Bael patra fruit-derived CDs turned out to be an excellent carbon source without the need of chemical additives [39]. Typically, three different components of Bael patra fruit, i.e., hard shell, edible pulp, and a mixture of pulp and gum were reported for the successful synthesis of emissive CDs, and it was found that CDs derived from the hard shell had the highest quantum yield of 59.39%. This is because the synthesized CDs via the carbonization approach resulted in more production of carbon content and micropores over its surface. These micropores can provide a higher ratio of surface-active sites for the adsorption of toxins from wastewater samples, thus significantly improving the sensing performance.

2.2. Vegetables

The exploration of new carbon sources that possess abundant reserve, green, simple and high quality CDs has drawn tremendous attention in the area of kitchen waste, such as vegetables. Various green and non-green vegetables, such as celery leaves, sweet pepper, lemon grass, tomato, carrot, rose-heart radish, turmeric, cinnamon, red chili, black pepper, hongcaitai, cauliflower, kelp, tomato, crown daisy leaves, cabbage, cherry tomatoes, scallion leaves, and red beet, have been reported, as shown in Table 2 [42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58]. Both green and non-green vegetables have many properties and their own advantages. For instance, the family of green vegetables, such as celery leaves, lemon grass, hongcaitai, kelp, cabbage, crown daisy leaf, and scallion, often contain many organic compounds, such as organic acids, amides, amino acids, proteins, saccharides, carbohydrates, chlorophyll, etc., which can bring good physical and chemical properties to CDs. Meanwhile, non-green vegetables, such as tomato, red chili, turmeric, black pepper, cinnamon, and red beet, are plants rich in various bioactive compounds of lycopene, capsaicin, curcumin, piperine, and cinnamaldehyde, respectively, which enable their application in the biomedical field. It was reported that bioactive compounds will partially remain inside or at the surface of the CDs after the hydrothermal process, leading to different photoluminescent and biomedical properties. Vasimalai et al. (2018) have demonstrated the uses of cinnamon, red chili, turmeric, and black pepper as CD precursors for biomedical applications [48]. They found that black pepper CDs have the highest quantum yield of 43.6% due to the various functional groups present in the sample, namely O–H, C–H, C–O–N, C=O, C–O, and C–N vibrational stretching peaks. All in all, the CDs prepared using celery leaves have contributed to the highest quantum yield of 53%. It was discovered that celery leaves contain abundant folic acid with affluent -COOH and the addition of L-glutathione as N, S- dopant has enriched their surface groups, which are beneficial for improving the quantum yield of CDs [42].

2.3. Flowers

Flowers such as Selenicereus grandifloras, water hyacinth, Osmanthus fragrans, rose flowers, and Tagetes erecta have also shown promise as carbon precursors, which were subsequently used to bind pesticides and metal ions. From Table 3, we can observe that there is an increase in quantum yield from 3.8% in 2019 to 63.7% in 2021. This significant improvement was attained by the presence of surface-active organic groups of C–N, –NH, and –OH, as shown in the FTIR analyses [59,60,61,62,63].
In another study, Shekarbeygi et al. (2020) evaluated the effect of rose pigments (blue, red, and yellow) and the effect of their extraction methods (aqueous and alcohol) on optical properties of the synthesized CDs [62]. The results indicated that the quantum yields obtained for all the CDs were not affected by the rose pigments and extraction methods, as the yields were almost the same. However, such rose pigments and extraction methods affected the CDs’ thermal stability and emission wavelengths. In the case of thermal stability, the fluorescence intensities were reduced in both aqueous and alcoholic CDs, with an increase in temperature ranging from 17 to 57 °C, yet aqueous CDs showed a higher decreasing rate than alcoholic CDs. As for the emission wavelengths, the alcohol extract had a larger wavelength than the aqueous extract, which may be due to the change in dielectric constant of solvent and the presence of more phenolic groups in the alcoholic extract. Thus, this work showed that CDs prepared with alcoholic extract with yellow petals has higher stability and a longer emission wavelength with better quantum yield than the others.

2.4. Leaves, Seeds, and Stems

In this section, we focused on the parts of plant-derived CDs, namely leaves (such as Ocimum sanctum leaves, bamboo leaves, gingko leaves, Gynostemma leaves, betel leaves, Calotropis procera leaves, Elettaria cardamomum leaves, Cornus walteri leaves, tea leaves, Kentucky bluegrass), seeds (such as Acacia concinna seeds, fennel seeds, Pearl millet seeds), and stem (lotus root), as presented in Table 4 [64,65,66,67,68,69,70,71,72,73,74,75,76,77,78]. Most of them have excellent healing properties, such as anti-inflammatory, antioxidant, antimicrobial, and antidote properties. Due to those health benefits, the synthesized CDs could serve simultaneously in the fluorescent sensing of metal ions and as contrast agents for live cells [64,65,66]. Nonetheless, more efforts are still needed to explore the effect of reaction temperature, time, and pH, despite the remarkable use of carbon precursors. The work of Dager et al. (2019) witnessed complete carbonization of fennel seeds (hydrocarbon converted to the graphitic structure) when the reaction temperature increased to the highest temperature at 500 °C for 3 h [70]. It was also found that by heating the sample longer (for 5 h) did not result in any significant change in either PL or crystallinity. Furthermore, when the synthesized CDs were dispersed in water at different pHs ranging from 3 to 13, the emission intensity gradually increased from acidic to basic media and eventually decreased when the reaction pH was higher than 9.
Leaves are recognized as the most excellent carbon precursors by reason of exhibiting higher quantum yields than seeds and stem. Most interestingly, the Calotropis procera leaf-derived CDs provided an excellent quantum yield (71.95%) without any toxic agents or surface passivation chemicals [73]. It is noteworthy that an excellent quantum yield resulted from the functional groups (OH, N–H, C=O, and C=N bond) derived from the carbon precursors itself. Furthermore, the synthesized CDs have a strong peak around 320 nm in the UV–Vis spectra of CDs (Figure 2), implying that it was spawned by the n → π* transition of C=O bonds over the surface of CDs.

2.5. Crop Residues

Crop residues, such as sago waste, palm kernel shell, and wheat straw, are another potential carbon source used in the green preparation of CDs (Table 5). In the year of 2014, Tan et al. successfully demonstrated the conversion of sago waste into fluorescent CDs via thermal pyrolysis without any surface passivation [79]. The heating temperature of pyrolysis was found to alter the degree of carbonization of bulk sago waste into carbonaceous residues. It was observed that heating treatments of lower than 400 °C can lead to incomplete carbonization, while higher temperatures can cause severe decomposition of the organic structures in sago waste into ashes, which leads to the loss of the fluorescing property. The optimum temperature of carbonization was found to be 400 °C. However, the synthesized CDs may suffer from low quantum yields due to the absence of solvent.
Solvents, especially solvents with high boiling points, play an important role in providing better carbonization rates for the formation of CDs. Of these, controlling the reaction conditions by adding solvent is of particular importance. Thus, to date, diethylene glycol [80], ultrapure water and ethanol [81], and deionized water [82] have been used as solvents to improve carbonization efficiency. The results showed that the reaction of diethylene glycol (DEG) on palm kernel shell-derived CDs had the highest quantum yield of 44%. This advancement is due to the higher boiling point of DEG than the others, thus allowing a more complete reaction for the formation of CDs at higher temperatures.

2.6. Fungi/Bacteria Species

The conversion of fungi/bacteria species into a value-added product such as carbonaceous nanomaterials has contributed to the green and sustainable improvement. Algal blooms [83], yogurt [84], enokitake mushroom [85], microalgae biochar [86], mushroom [87], agarose waste [88], and Shewanella oneidensis [89] have been previously chosen as a carbon resource (Table 6). A study conducted by Pacquiao et al. (2018) demonstrated the increment of quantum yield of 11% to 39% upon passivation with tetraethylenepentamine in the presence of 5% v/v sulfuric acid [85]. The XPS spectrum revealed the presence of small signals of nitrogen and sulfur at 399 eV (N1s) and 168 eV (S2p), respectively, thus confirming the successful heteroatoms doping on CDs and sulfuric acid as a reagent. Furthermore, the effect of temperatures (180, 200, and 250 °C) and reaction times (4, 6, and 8 h) on the quantum yield was evaluated, yielding the optimum temperature and reaction time at 250 °C for 4 h. The research, thus, concluded that higher temperatures and shorter reaction times can produce higher quantum yields. Even so, CDs prepared from agarose waste-derived CDs without using other passivation chemicals exhibited the highest quantum yield of 62% [88]. This quantum efficiency can be attributed to the presence of different functional groups (–C=O, –OH, and N–H) on the surface of CDs, in addition to being related to the surface defects and the particle size of CDs.

2.7. Waste Products

There are ongoing efforts to improve the photoluminescent properties while reducing the production cost and protecting the environment. Of the different plant and fungi sources, the reuse of waste (Table 7), such as frying oil [90], biocrude oil [91], polystyrene [92], tea powders [93], expired milk [94], coal powder [95], papers [96], polyolefin [97], polypropylene plastic waste [98], bike soot [99], etc. [100,101,102,103,104,105,106,107,108,109,110,111], has become one of the hottest scientific research topics nowadays for the development of CDs. Among these materials, plastic waste and heavy oil products offer a promising precursor, as they have produced high quantum yields of more than 60%. In the case of plastic waste, cup-derived CDs have displayed the highest quantum yield value of 65% at 310 nm compared to bottles (64%) and polybags (62%) [108]. The high quantum yield of CDs is highly dependent on the carbonyl groups in CDs, revealing the presence of NH2. Furthermore, all the three prepared CDs displayed a slight red shift in the emission peak, with respect to the excitation wavelength. These behavioral variations were mainly due to the presence of epoxy and hydroxyl functional groups in the CDs prepared from the plastic waste, which resulted in creating new energy levels between n-π* gaps. Due to this reason, the energy gaps of the CDs between the lowest unoccupied molecular orbital (LUMO) and the highest occupied molecular orbital (HOMO) became reduced with the increasing degree of surface oxidation, thereby contributing to the enhanced fluorescence property (Figure 3).
Recently, the precursors with higher molecular weight, higher heteroatoms and higher aromatic structures were shown to be more conducive to the production of CDs. Ma et al. (2021) observed that CDs synthesized with asphalt as the precursor have the highest fluorescence quantum yield compared to heavy oil, light deasphalted oil (LDAO), and heavy deasphalted oil (HDAO) [110]. The reason is that the CDs from asphalt had the highest aromatic carbon ratio and the lowest naphthenic carbon ratio and alkyl carbon ratio. In addition, the asphalt had higher molecular weight, more oxygen, nitrogen and sulfur contents, and higher carbon/hydrogen atomic ratio than the other precursors. This result was further confirmed by the presence of the doublet γ peak and (002) peak in the X-ray diffraction pattern of asphalt, which indicated that the asphalt had a great quantity of graphite-like ordered structures.

3. CDs Synthesized from Chemical Precursors

Although many natural precursors have been used in the preparation of low-cost CDs, chemical precursors are still being explored to this day to produce high quantum yields of CDs. This section reviews two types of reaction chemical precursors used in CD preparation, namely acid and non-acid reagents, which were briefly summarized in Table 8 and Table 9.

3.1. Acid Reagents

Since 2012, carbon sources based on acid reagents, such as citric acid, phosphoric acid, acetic acid, folic acid, ascorbic acid, sodium citrate, ascorbic acid, glutamic acid, malonic acid, maleic anhydride, boric acid, pyrogallic acid, phthalic acid, 3-aminobenzeneboronic acid, succinic acid, p-aminosalicylic acid, diethylenetriamine-pentacetate acid, maleic anhydride, DL-thioctic acid, sulfamic acid, tartaric acid, 2-aminoterephthalic acid, trans-aconitic acid, dehydroabietic acid, dithiosalicylic acid, etc., have been intensively employed in the production of CDs [112,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142,143,144,145,146,147,148,149,150,151,152,153,154,155,156,157,158,159,160,161,162,163,164,165,166,167,168,169,170,171,172,173,174,175,176,177,178,179,180,181,182,183,184,185,186,187,188,189,190,191,192,193,194,195,196,197].
Dong et al. (2013) first compared the fluorescence properties of the following three types of CDs: (i) O-CDs synthesized from citric acid only, (ii) N-CDs synthesized from citric acid and glycine, and (iii) N,S-CDs synthesized from citric acid and L-cysteine [113]. It was observed that the doping of nitrogen into O-CDs can introduce a new kind of surface state (labelled as the N-state). Electrons trapped by the new formed surface states are able to facilitate a high yield of radiative recombination. Although the quantum yield of N-CDs was found to be higher than that of O-CDs, the fluorescence spectra were still broad and excitation-dependent. In such cases, the introduction of sulfur atoms into CDs could lead to more significant enhancement, offering higher yields and excitation-independent emission. The introduced sulfur atoms seem to be able to eliminate the O-states and enhance the N-state, meaning that the original surface states are almost neglected in the N,S-CDs. However, when the authors tuned the ratio of L-cysteine into citric acid, from 1 g:2 g into 0.125 g:2 g, the quantum yield of N,S-CDs decreased from 73% to 37% and the emission wavelength was dependent on excitation and yielded a red-shift from 415 to 540 nm under the excitation wavelength of 375 to 480 nm. Interestingly, when the excitation wavelength was set lower than 375 nm, the emission wavelength was excitation independent. Although the aforementioned CDs were highly enhanced by nitrogen and sulfur co-doping, the chemistry involved therein is extremely challenging due to the involvement of oxygen, let alone the interference from defects. Therefore, until single nitrogen doping-sources, such as urea, ethylenediamine, PEG diamine, melamine, etc., have been widely reported, as shown in Table 8.
On the other hand, citric acid as the carbon source was commonly used in the chemical synthesis of CDs. Citric acid has considerable advantages over other acid reagents, such as being relatively cheap, more sour, less harmful to the environment, and readily available in large commercial quantities. Consequently, two highest fluorescence quantum yields were attributed to the reflux reaction of citric acid and diethylenetriamine [133] and hydrothermal reaction of citric acid and ethylenediamine [176] at 82.40% and 85.69%, respectively. As expected, ethylenediamine as N-doping agents plays a major role in improving the fluorescence properties in CDs. By way of example, Wang et al. (2015) have reported the influence of three different polyethylenic amine molecules, i.e., ethylenediamine (EDA), diethylenetriamine (DETA), and tetraethylenepentamine (TEPA) with a combination of citric acid on photoluminescence performance [129]. The results showed that the CDs-EDA has the highest PL quantum yield at 69.3%, followed by CDs-DETA (68%) and CDs-TEPA (33.4%). The results imply that the increasing presence of cyclic imines (C=N) with the enhancement of conjugated π-domains in CDs imparts superior PL efficiency.
There is another study that reported the fabrication of CDs from different nitrogen sources, such as ethylenediamine (e), hexamethylenetetramine (h), and triethanolamine (t) [130]. From the X-ray photoelectron spectroscopy analyses, the ratios of carbon to nitrogen in these CDs were determined, namely, 83:17 for e-CDs, 86.14 for h-CDs, and 96:4 for t-CDs. Because of the high nitrogen content in e-CDs, the quantum yield was found to be highest at 53%. The amount of nitrogen can be typically correlated with a high photoluminescence quantum yield and confirms similar features in the absorption spectra (a shoulder at 234 nm and a broad peak at 340 nm) between e-CDs and pure citrazinic acid, the basic unit from the presumed class of fluorophores. However, at wavelengths longer than 400 nm, pure citrazinic acid has no absorption characteristic, unlike e, h and t-CDs, as shown in Figure 4a. The presence of broad absorption in this energy is commonly assigned to surface states related to the functional surface groups in CDs, which form low-energy sub-band gaps within the n−π* band gap.
Furthermore, Chang et al. (2017) reported that the phosphoric acid (H3PO4) or sucrose solution alone could not produce a significant fluorescence emission under the excitation of 423 nm [137]. However, when H3PO4 and sucrose solutions were mixed and incubated in the oven, the fluorescence light was emitted at the excitation wavelength of 423 nm. It is thought that the mild heating of H3PO4 has the power to hydrolyze sugar into simple carbon-rich molecules. Of note, to deactivate the carbonization process, the authors then added sodium hydroxide into CDs, but this appeared to cause a high salt content that was not favourable to many applications. Hence, acetone with a 1:1 volume-to-volume ratio was used to overcome this problem, resulting in two phases, as shown in Figure 4b. The highly ionic salt species remained in the water mother-liquor, while the CDs that were less ionic were partitioned into the acetone phase. The removal of CDs from acetone can be easily achieved via evaporation, since acetone is quite volatile. Based on this, the fluorescence signal of CDs was found to be increased by approximately 10%.
Recently, other carbon sources derived from amino acids, namely arginine, lysine, histidine, cysteine, and methionine have been reported [168,193,194,195]. This is because of their several advantages, for instance, non-toxic, biocompatible, and eco-friendly nature. Most importantly, each of the amino acids have different numbers of -COOH, -NH, and -SH functional groups, which can effectively improve the optical and chemical properties of CDs. Upon mixing these amino acids with citric acid via the microwave method, N and/or S-doped CDs were successfully synthesized in a very short time for the first time, which only took 1–4 min [168]. The reaction mechanism was proposed to include the following two steps: (i) reaction of carboxylic acid of citric acid with the -NH and –SH groups of amino acids via condensation polymerization and then the esterification of citric acid with amino acids, (ii) carbonization of these polymer clusters during irradiation by total dehydration to generate N and/or S doped CDs under microwave exposure. The research concluded that only cysteine-CDs were found to possess the highest quantum yield of 89.5% upon the N and S atom containing CDs.
The highest quantum yield of 93% was achieved when sodium citrate dihydrate and urea were used as precursors [179]. Sodium citrate dihydrate plays an important role as a self-assembly trigger for a carbon-based structure, due to the intermolecular H-bonding, while urea acts as nitrogen-doping precursors, which both are responsible for the increase in quantum yield. In another study, instead of using common synthesis methods such as oxidation, combustion, and hydrothermal, Ji et al. (2021) established another method, which is simpler and timesaving. Briefly, 0.34 g of L-cysteine, 0.15 g of urea, and 8.5 g of diphosphorus pentoxide were mixed, followed by the addition of 6 mL of water with a rapid stirring process [193]. These CDs were then embedded in in polyvinyl alcohol (PVA) gel substrate to form a smooth and high fluorescent CD/PVA film for further use in hexavalent chromium detection.
Table 8. Summary of the synthesis of CDs from acid reagents.
Table 8. Summary of the synthesis of CDs from acid reagents.
PrecursorTechniquePropertiesYearReference
Carbon SourcePassivation/SolventParticle SizeFluorescenceQuantum Yield
Citric acidTerbium (III) nitrate pentahydrateCarbonizationTEM- 3 nmλem- 450 nm
λex- 320 nm
-2012[112]
Citric acid monohydrateL-cysteineHydrothermal treatmentHRTEM- 7 nmλem- 415 nm
λex- 345 nm
73%2013[113]
--λem- 435 nm
λex- 345 nm
5.3%
Glycine-λem- 415 nm
λex- 345 nm
16.9%
Citric acidUreaMicrowaveTEM- 4 to 6 nmλem- 460 nm
λex- 360 nm
0.13%2013[114]
Ethyleneglycol bis-(2-aminoethyl ether)- N,N,N′,N′-tetraacetic acidTris(hydroxymethyl)aminomethaneThermal carbonizationTEM- 5 nmλem- 425 nm
λex- 310 nm
28%2014[115]
Citric acidPEG-diamineSolid-phaseTEM- 1.7 nmλem- 435 nm
λex- 360 nm
31%2014[116]
Citric acidPoly(ethylenimine)PyrolysisHRTEM- 3.5–4.5 nm-42.5%2014[117]
Folic acidEthylene glycol and nanopure waterHydrothermalTEM- 4.5 nmλem- 470 nm
λex- 395 nm
15.7%2014[118]
Poly(ethylene glycol) and ascorbic acidDistilled waterMicrowaveTEM- 2.3 nmλem- 450 nm
λex- 373 nm
-2014[119]
Citric acidEthylenediamineMicrowave-assisted pyrolysisTEM- 3 nmλem- 455 nm
λex- 280 nm
-2015[120]
Citric acidL-Tyrosine methyl ester hydrochlorideHydrothermalTEM- 3.7 nmλem- 433 nm
λex- 348 nm
3.8%2015[121]
Citric acidEthylenediamine and double distilled waterHydrothermal--75.0%2015[122]
Citric acid1-Aminopropyl-3-methy-imidazolium bromidePyrolysisHRTEM- 0.6–1.6 nmλem- 440 nm
λex- 380 nm
2.03–27.66%2015[123]
Glacial acetic acidN-Acetyl-L-cysteine, diphosphorus pentoxide and distilled deionized waterSimple mixingTEM- 2.51–3.44 nmλem- 480 nm
λex- 300 nm
4.65%2015[124]
Citric acidL-cysteine, urea and ultrapure waterMicrowaveTEM- 1.1 nmλem- 450 nm
λex- 353 nm
25.2%2015[125]
Citric acidDithiooxamide and distilled waterMicrowave-assisted hydrothermalSTEM- 2 nmλem- 448 nm
λex- 360 nm
17.6%2015[126]
Citric acid monohydrateAmmonia and double distilled waterHydrothermalTEM- 3.7 nmλem- 442 nm
λex- 350 nm
40.5%2015[127]
Sodium citrateUrea and ultrapure waterElectrochemical carbonizationTEM- 2.4 nmλem- 433 nm
λex- 351 nm
11.9%2015[128]
Citric acid anhydrousEthelendiamineCondensation carbonizationTEM- 3.9 nmλem- 445 nm
λex- 365 nm
69.3%2015[129]
DiethylenetriamineTEM- 3.7 nm68%
TetraethylenepentamineTEM- 4.1 nm33.4%
Citric acid anhydrousEthylenediamine and deionized waterHydrothermal-λem- 440 nm
λex- 320 nm
53%2016[130]
Hexamethylenetetramine and deionized waterλem- 420 nm
λex- 320 nm
17%
Triethanol-amine and deionized waterλem- 420 nm
λex- 320 nm
7%
Citric acidEthylenediamine and deionized waterHydrothermalTEM- 5 to 7 nmλem- 443 nm
λex- 365 nm
-2016[131]
Citric acidBranched polyethylenimineCondensationDLS- 1.9 nmλem- 450 nm
λex- 350 nm
-2016[132]
Citric acidDiethylenetriamineReflux treatmentTEM- 5 -7 nm-82.40%2016[133]
Ascorbic acid and valineEthanol and distilled waterHydrothermalTEM- 4 nmλem- 430 nm
λex- 352 nm
4.8%2016[134]
L-glutamic acidSilica gel powders and waterMicrowaveTEM- 1.64 nmλem- 450 nm
λex- 370 nm
41.2%2016[135]
Malonic acidUrea and ultrapure waterHydrothermalTEM- 2.5 nmλem- 397 nm
λex- 320 nm
12.6%2017[136]
Sucrose and phosphoric acidSodium hydroxideCarbonizationSEM- 10 nmλem- 524 nm
λex- 423 nm
-2017[137]
Citric acidSodium phosphateSolid-phaseTEM- 1.7 nmλem- 435 nm
λex- 360 nm
-2017[138]
Citric acidEthylenediamine and ultrapure waterHydrothermalTEM- <10 nmλem- 431 nm
λex- 337 nm
32.25%2017[139]
Citric acid monohydrateThioureaMicrowave solid-phase pyrolysisTEM- 2 nmλem- 436 nm
λex- 358 nm
23.6%2017[140]
Citric acidSilkHydrothermalTEM- 5.6 nmλem- 425 nm
λex- 360 nm
61.1%2017[141]
Citric acidMelamineHydrothermalTEM- 1.8 nmλem- 422 nm
λex- 320 nm
8.11%2017[142]
Citric acidTartaric acid, ethanediamine and oleic acidSolvothermalTEM- 2.66 nmλem- 460 nm
λex- 360 nm
42.2%2017[143]
Citric acid and ureaWaterSolvothermalTEM- 1.7 nmλem- 448–638 nm
λex- 375 nm
-2017[144]
GlycerolTEM- 2.8 nm
DimethylformamideTEM- 4.5 nm
Maleic anhy- dride and tetraethylenepentamineSulfuric acid and deionized waterPyrolysisTEM- 20 nm
DLS- 8 nm
λem- 450 nm
λex- 360 nm
21%2017[145]
D-(+)-maltose monohydrate,
boric acid and thiocarbamide
-HydrothermalTEM- 2.0 nmλem- 415 nm
λex- 326 nm
8.9%2017[146]
Pyrogallic acidN-N-dumethylformamideSolvothermalTEM- 11.9 nmλem- 520 nm
λex- 360–450 nm
16.8%2018[147]
L-histidine and citric acidEthylene glycolPolyol microwaveTEM- 19 nmλem- 430–511 nm
λex- 350 nm
-2018[148]
Phthalic acid and triethylenediamine hexahydrateDeionized waterMicrowaveTEM- 2–6 nmλem- 520–542 nm
λex- 360–440 nm
16.1%2018[149]
Citric acidUreaSolvothermalTEM- 1.87 nmλem- 590 nm
λex- 540 nm
43%2018[150]
Citric acidLysine and ultrapure waterHydrothermalTEM- 10 nm--2018[151]
Citric acid monohydrate-Thermal treatmentTEM- 3.5 nmλem- 450 nm
λex- 360 nm
3.54%2018[152]
Citric acidAmmonium thiocyanate and deionized waterMicrowave-assistedHRTEM- 30 nmλem- 490 nm
λex- 410 nm
-2018[153]
Folic acid and p-phenylenediamineSodium hydroxideHydrothermalTEM- 2 nmλem- 505 nm
λex- 420 nm
8.4%2018[154]
3-Aminobenzeneboronic acidDeionized waterHydrothermalTEM- 3 nmλem- 504 nm
λex- 400 nm
-2018[155]
Succinic acidDeionized water and glycerolHydrothermalTEM- 2.3 nmλem- 410 nm
λex- 280 nm
11%2018[156]
TEM- 4.6 nmλem- 525 nm
λex- 480 nm
7%
Phosphoric acidEthylenediamineSimple heating (180 °C, 2 h)TEM- 3.2 nmλem- 430 nm
λex- 340 nm
5.17%2018[157]
Simple heating (280 °C, 2 h)TEM- 6.4 nmλem- 413 nm
λex- 340 nm
21.8%
Phosphoric acid and ethanolamineWaterMicrowave irradiationTEM- 3.4 nmλem- 417 nm
λex- 340 nm
20.52%2018[158]
P-aminosalicylic acidEthyleneglycol dimethacrylate and double distilled waterHydrothermalTEM- 3 nm
AFM- 1.6 nm
DLS- 11.7 nm
λem- 520 nm
λex- 390 nm
27.2%2018[159]
Sodium citrateUrea and dimethylformamideSolvothermalTEM- 3.52 nmλem- 446 nm
λex- 370 nm
67%2018[160]
Citric acid monohydrate3-(Aminopro- pyl)triethoxysilane (APTES)Thermal decompositionTEM- 5–15 nmλem- 416 and 480 nm-2019[161]
Citric acidUrea and deionized waterHydrothermal---2019[162]
Glycine and deionized water-
Citric acid, deionized water, ethylene glycol, N,N’- bis(2-aminoethyl)-1,3-propanediamineHRTEM- 5–6 nm
Citric acid and ureaN,N-dimethylformamideSolvothermal-λem- 450, 550, 630 nm
λex- 400–450 nm
-2019[163]
N,N-dimethylformamide, NaOH and HClTEM- 3.7 nmλem- 630 nm
λex- 550 nm
-
N,N-dimethylformamide, NaOH and waterTEM- 2.1 nm--
Procaine hydrochloride and citric acidDouble distilled water and ethylenediamineHydrothermalTEM- 3.3 nmλem- 440 nm
λex- 360 nm
47.1%2019[164]
Anhydrous citric acidN-(β-aminoethyl)-γ-aminopropyl-methyldimethoxysilaneHydrothermalTEM- 2.22 nmλem- 460 nm
λex- 370 nm
51.8%2019[165]
Citric acidThiourea and deionized waterMicrowave-assisted pyrolysisTEM- 3.3 nm--2019[166]
Sodium
citrate and aminopyrazine
Ultrapure waterHydrothermalTEM- 2.38 nmλem- 389 nm
λex- 310 nm
11.8%2019[167]
Citric acidDeionized water and arginineMicrowaveTEM- 11 ± 4 nmλem- 330 nm
λex- 430 nm
3.9 ± 0.4%2019[168]
Deionized water and lysineTEM- 17 ± 2 nmλem- 330 nm
λex- 430 nm
4.2 ± 1.9%
Deionized water and histidineTEM- 6 ± 5 nmλem- 330 nm
λex- 433 nm
2.8 ± 0.2%
Deionized water and cysteineTEM- 10 ± 7 nmλem- 330 nm
λex- 420 nm
89.5 ± 2.3%
Deionized water and methionineTEM- 9 ± 5 nmλem- 330 nm
λex- 407 nm
2.5 ± 0.6%
Citric acid and phenylalanineUltrapure waterHydrothermalTEM- 2–3 nmλem- 330 nm
λex- 310 nm
-2020[169]
Polyacrylamide and citric acidUltrapure waterHydrothermalTEM- 4.1 nmλem- 330 nm
λex- 310 nm
12.6%2020[170]
Citric acid and urea-Infrared carbonizationTEM- 5–10 nmλem- 475 nm
λex- 360 nm
22.2%2020[171]
Citric acid and urea-Hydrothermal
(180 °C, 20 min)
TEM- 2–7 nm-46%
26%
4%
2020[172]
(230 °C, 20 min)λem- 394, 440, 523 nm
λex- 350 nm
23%
35%
36%
Citric acidMelamine and formaldehydeHydrothermalTEM- 3.7 nmλem- 425 nm
λex- 350 nm
63.7%2020[173]
Citric acid monohydrateUreaMicrowave irradiationTEM- 6 nmλem- 536 and 532 nm
λex- 350 nm
-2020[174]
Citric acidPhenylalanineHydrothermalTEM- 11.9 nmλem- 413 nm
λex- 350 nm
65%2020[175]
Citric acid monohydrateEthylenediamineHydrothermalTEM- 5–10 nmλem- blue
λex- 305–395 nm
85.69%2020[176]
Citric acidEthylenediamine and ultrapure waterHydrothermalTEM- 5 nmλem- 444 nm
λex- 360 nm
-2020[177]
Ascorbic acidUrea and deionized waterMicrowave irradiationTEM- 2 nmλem- 415 nm
λex- 340 nm
7%2020[178]
Sodium citrate dihydrateUrea and deionized waterThermal pyrolysisTEM- 2.75 nmλem- 525 nm
λex- 400 nm
93%2020[179]
Citric acid and 3-aminobenzeneboronicDimethylformamideHydrothermalHRTEM- 3.4 nm--2020[180]
Diethylenetriamine- pentacetate acidUltrapure waterCarbonizationHRTEM- 2.85 nm--2020[181]
Maleic anhydride and triethylenetetramineDeionized water and nitric acidPyrolysisTEM- 5.9 nmλem- 400 nm
λex- 320 nm
6.3%2021[182]
DL-thioctic acidDimethylformamide, trisodium citrate dihydrate, sodium hydroxide, double deionized waterHydrothermalHRTEM- 2.52 nmλem- 438 nm
λex- 340 nm
-2021[183]
Citric acid and sulfamic acidPolyethyleneimineTwo-step hydrothermalTEM- 5.1 nmλem- 460 nm
λex- 355 nm
29.1%2021[184]
Tartaric acidUreaSolid-phase thermalTEM- 4.13 nmλem- 537 nm
λex- 460 nm
10.5%2021[185]
2-aminoterephthalic acid and polyethylene glycolOrthophosphoric acidMicrowave-assisted pyrolysisTEM- 3–10 nmλem- 470 nm
λex- 410 nm
67%2021[186]
Citric acidUltrapure water and ethylenediamineHydrothermalTEM- 3.1 nmλem- 445 nm
λex- 356 nm
-2021[187]
Citric acidEthylenediamine and waterMicrowaveTEM- 2.3 nmλem- 450 nm
λex- 360 nm
-2021[188]
Trans-aconitic acidDiethylenetriamine and distilled waterHydrothermalHRTEM- 2–8 nmλem- 435 nm
λex- 345 nm
81%2021[189]
Dehydroabietic acidEthanolamineHydrothermalTEM- 3.2 nmλem- 433 nm
λex- 365 nm
10%2021[190]
Citric acidL-glutamineHydrothermalTEM- 3.5 nmλem- 450 nm
λex- 360 nm
-2021[191]
D-glutamineTEM- 3–4 nm
Dithiosalicylic acidAcetic acid and o-phenylenediamineSolvothermalTEM- 4.5 nmλem- 620 nm
λex- 560 nm
4.05%2021[192]
Acetic acid and m-phenylenediamineTEM- 4.0 nmλem- 560 nm
λex- 460 nm
20.77%
Acetic acid and p-phenylenediamineTEM- 3.5 nmλem- 478 nm
λex- 460 nm
1.76%
L-cysteine and ureaDiphosphorus pentoxide and waterOne-pot synthesisTEM- 4.5 nmλem- 445 nm
λex- 362 nm
17%2021[193]
Methyl cellulose and L-cysteineEthylenediamineHydrothermalTEM-19 nmλem- 370 nm
λex- 330 nm
12.3%2021[194]
Ce (NO3)3·6H2O and L-histidineSodium hydroxide and deionized waterOne-pot hydrothermalSEM- 46 nm--2021[195]
Polyethylenimine and citric acidHot waterMicrowave-assistedSTEM- 12 nmλem- 442 nm
λex- 354 nm
54%2022[196]
Citric acid and ureaUltrapure waterSolvothermalHRTEM- 3.18 nmλem- 470 nm
λex- 330 nm
20.1%2022[197]
20 mL dimethylformamideHRTEM- 3.25 nmλem- 500 nm
λex- 330 nm
22.1%
10 mL dimethylformamide and ethanolHRTEM- 3.47 nmλem- 539 nm
λex- 330 nm
21.9%
10 mL dimethylformamide and acetic acidHRTEM- 3.68 nmλem- 595 nm
λex- 330 nm
24.2%

3.2. Non-Acid Reagents

Carbon sources based on non-acid reagents, namely graphite oxide [198], poly(ethylene glycol) [199,200], 3-(3,4-dihydroxyphenyl)-L-alanine [201], ethanolamine [202], polyimide [203], polyethyleneglycol bis(3-aminopropyl) [204], ethanol [205], 3-bromophenol [206], azidoimidizole [207], activated carbon [208], glucose [209], N-methylethanolammonium thioglycolate [209], chitosan [210], chlorophyll [211], microcrystalline cellulose [212], o-phenylenediamine [213], carbon paper [214], papain [215], dextrose solution [216], etc. [217,218,219,220,221,222,223,224,225,226,227,228,229,230], have also been reviewed, as shown in Table 9. 3-Bromophenol is the first insecticide adopted to be a potential precursor, thus suggesting that carbon sources are not limited in typical chemical reagents with high purity or low toxicity [206]. In this study, the effect of mixed solvent of ethanol and water ratio (v/v) such as 1:9, 3:7, 5:5, 7:3 and 9:1 and the effect of temperatures such as 150, 160, 170 °C were evaluated. The results indicate that both much thicker organic solvents and much higher carbonized temperatures are in favour of the formation of C-dots featuring mono-dispersion, high yield, and smaller size. The proposed mechanism behind the fabrication of insecticide-based CDs was verified by the occurrence of aromatic C-Br and Br anion, Csp2 C-C peak and Csp3 C-C peak found on the XPS spectrum. The three possible main classes of the reaction that happened among the 3-bromophenol molecules are given in Figure 5.
Efforts to develop and improve the fluorescence of the CDs have been made toward polymer compounds, including synthetic and natural polymers. Thanks to the toxicity of synthetic polymers, natural polymers extracted from plants have been used as a carbon source with the characteristics of being environment-friendly, non-toxic, low price, good biocompatibility, and stable chemical properties. For this reason, Wu et al. (2017) reported microcrystalline cellulose as a carbon precursor to prepare a facile and low-cost one-step hydrothermal treatment of CQDs [212]. The modification of these CQDs with ethylenediamine not only allowed nitrogen atoms on its surface but also introduced them into the carbon nuclear lattice to obtain better optical properties and higher quantum yield than CQDs. Fortunately, nitrogen doped CQDs (NCQDs) have a fluorescence emission intensity of 7.1 times stronger than the CQDs, and it can be explained by the reduction of epoxy, ether, and carboxyl groups on the surface of NCQDs via nucleophilic reactions with amidogen, thus forming a few non-radiative recombination centers in NCQDs. The amination and/or amidation also create radiative recombination centers, such as amine (–C–N–) and amide groups (R2-NCOR), which belong to electron-donating groups. These electron-donating groups and π-electrons that existed in the –C=N group have contributed to the improvement of the fluorescence intensity of a substance. All in all, these NCQDs successfully provided the highest quantum yield of 55%.
The synthesis of CDs using non-acid reagents can be further explored by attempting to use three different aniline compounds, i.e., 1,2,4,5-benzenetetramine tetrahydrochloride, 1,2,4-benzenetriamine dihydrochloride, and o-phenylenediamine, as carbon sources and ethanol as a solvent [217]. In an approach where more amino groups and a high nitrogen content are present in CDs-benzenetetramine tetrahydrochloride, the excitation and emission wavelengths all exhibited a redshift with an increase in the CD particle size. The quantum yield also greatly increased up to 30.2%. Additionally, the fluorescence intensity of the CDs was observed to gradually increase with the increasing amount of raw material from 0.005 g to 0.025 g. It is important to note that the number of amino groups and particle size of CDs have a great influence on the emission wavelength, while the amount of raw material influences the color and intensity of the fluorescence. In another study, the CDs were synthesized using three types of aldehydes, such as glutaraldehyde, nitrobenzaldehyde and benzaldehyde, via the solvothermal method [225]. The resulting CDs were then composited with chitosan, functionalized by a DNA probe and utilized as a detecting platform for rapid detection of microRNA-21. The function of CDs is further classified as a cross-linking agent and fluorescent donor, while chitosan plays an important role in the preparation of three-dimensional hydrogel frames.
Table 9. Summary of the synthesis of CDs from non-acid reagents.
Table 9. Summary of the synthesis of CDs from non-acid reagents.
PrecursorTechniquePropertiesYearReference
Carbon SourcePassivation/SolventParticle SizeFluorescenceQuantum Yield
Graphite oxideNitric acidMicrowave-hydrothermalTEM- 4 nmλem- 520 nm
λex- 470 nm
2.72%2011[198]
Poly(ethylene glycol)Sodium hydroxide and distilled waterReflux methodTEM- 5 nmλem- bright blue
λex- 350 nm
λem- cyan
λex- 390 nm
λem- yellow
λex- 470 nm
λem- red
λex- 540 nm
-2013[199]
Polyethylene glycol 1500Serine and glycerinMicrowave pyrolysis---2013[200]
3-(3,4-dihydroxyphenyl)-L- alanine-Carbonization-oxidationTEM- 3.64 nmλem- 500 nm
λex- 400 nm
6.3%2013[201]
Nitric acidTEM- 4.31 nmλem- 475 nm
λex- 360 nm
1%
Ethanolamine-PyrolysisTEM- 2.7 nmλem- 450 nm
λex- 365 nm
7%2014[202]
Hydrogen peroxideTEM- 8.3 nm10.3%
Polyimide-HydrothermalTEM- 4 nmλem- 490 nm
λex- 365 nm
20.9%2015[203]
Polyethyleneglycol bis(3-aminopropyl)6-Bromohexylboronic acidThermal carbonizationTEM- 5 nmλem- 440 nm
λex- 362 nm
0.3%2015[204]
EthanolHydrogen peroxide and deionized waterHydrothermalTEM- 4.8 nmλem- 456 nm
λex- 400 nm
38.7%2015[205]
3-BromophenolEthanol and deionized waterCarbonizationTEM- 5.2 nmλem- 440 nm
λex- 367 nm
19.6%2016[206]
AzidoimidizoleEthanol-AFM- 5–10 nmλem- 515 nm
λex- 460 nm
-2016[207]
Activated carbonPotassium permanganate, sulfuric acid, deionized water, hydrogen peroxideExhausted oxidationTEM- 12 nmλem- 465 nm
λex- 350 nm
3.94%2016[208]
Potassium permanganate, sulfuric acid, deionized water, hydrogen peroxide, PAMAM-NH2TEM- 65 nm6.93%
GlucoseWater and sodium hydroxideUltrasonic- 7%2016[209]
N-Methylethanolammonium thioglycolateWater and hydrogen peroxideHRTEM- 3–8 nm12.5%
Chitosan-CarbonizationTEM- 1–6 nmλem- 390 nm
λex- 310 nm
4.34%2016[210]
ChlorophyllWaterHydrothermalDLS- 18 nmλem- 520 nm
λex- 440 nm
-2017[211]
Microcrystalline celluloseEthylenediamineHydrothermalTEM- 3.2 nmλem- 426–436 nm
λex- 360 nm
55%2017[212]
o-PhenylenediamineEthanolHydrothermalTEM- 1–2 nmλem- 400–600 nm
λex- 350–500 nm
20%2018[213]
Carbon paperNitric acidHydrothermalTEM- 4.8 nmλem- 450 nm
λex- 350 nm
5.1%2018[214]
Papain and PEG6000Ultrapure waterHydrothermalTEM- 2–3 nmλem- 420 nm
λex- 320 nm
9.45%2018[215]
Dextrose solutionHydrochloric acidMechano-chemicalTEM- 10 nmλem- 456 nm
λex- 390 nm
40%2018[216]
1,2,4,5- Benzenetetramine tetrahydrochlorideEthanolSolvothermalTEM- 9.39 nmλem- 605 nm
λex- 540 nm
30.2%2019[217]
1,2,4-Benzenetriamine dihydrochlorideTEM- 8.60 nmλem- 598 nm
λex- 510 nm
13.4%
o-PhenylenediamineTEM- 6.50 nmλem- 538 nm
λex- 420 nm
16.7%
Copper (II) chloride dihydrateEthanediamineHydrothermalTEM- 1.8 nmλem- 380 nm
λex- 320 nm
7.8%2019[218]
Glucose and taurineDistilled waterHydrothermalTEM- 3 nmλem- 410 nm
λex- 340 nm
11%2019[219]
Polyethylene glycol-PyrolysisDLS- 10 nmλem- ~380 nm
λex- 340 nm
16%2020[220]
m-PhenylenediamineDeionized waterHydrothermalTEM- 5.1 nmλem- 420 nm
λex- 340 nm
12%2020[221]
GlucosamineEthylenediamine and waterMicrowave digestionTEM- 4.45 nmλem- 466 nm
λex- 384 nm
25.38%2020[222]
LactoseHydrochloric acidHydrothermalTEM- 7 to 8 nm--2021[223]
Selenourea and o-phenylenediamineHydrochloric acidHydrothermalTEM- 3 nmλem- 625 nm
λex- 564 nm
23.6%2021[224]
GlutaraldehydeEthanolSolvothermalTEM- 1 nmλem- 453 nm
λex- 360 nm
-2021[225]
NitrobenzaldehydeTEM- 5 nmλem- 421 nm
λex- 360 nm
Benzaldehyde-λem- 430 nm
λex- 360 nm
Diphenyl etherp-Phenylenediamine TEM- 2.8 nmλem- ultraviolet
λex- 285 nm
8%2021[226]
p-Phenylenediamine, dopamine and tris(hydroxymethylaminomethane)TEM- 10–18 nmλem- red
λex- 285 nm
15.5%
o-PhenylenediamineEthanolTwo separate solutions mixed in one-pot hydrothermalTEM- 5 nmλem- ~570 nm
λex- 430 nm
-2022[227]
Ammonium sulfateDeionized water
GlucoseDeionized waterHydrothermalTEM- 8.9 nmλem- 450 nm
λex- 350 nm
-2022[228]
Boric acidTEM- 6.2 nmλem- 400 nm
λex- 320 nm
Sodium persulfateTEM- 6.9 nmλem- 400 nm
λex- 320 nm
UreaTEM- 5.6 nmλem- 450 nm
λex- 390 nm
p-Phenylenediamine and thylenediamineAnhydrous ethanolHydrothermalTEM- 2.76 nm--2022[229]
m-PhenylenediamineEthanolSolvothermalTEM- 6.9 nmλem- 440 nm
λex- 380 nm
11%2022[230]
o-PhenylenediamineTEM- 7.8 nmλem- 550 nm
λex- 380 nm
17%

4. Application of CD-Based Optical Sensor for Environmental Monitoring

The incorporation of CDs in optical sensors for environmental monitoring, including the detection of heavy metal ions, phenol, pesticides, and nitroaromatic explosives, are reviewed below (Figure 6). This review aims to survey state-of-the-art CD-based optical sensors with their sensing performance, namely range of detection, limit of detection, and linear correlation coefficient.

4.1. Heavy Metal Ions

Environmental pollutions caused by toxic heavy metal ions (HMI) have resulted in the development of different optical sensors using nanoscale materials [231,232,233,234,235,236,237]. Of late, CD-based optical sensors have garnered tremendous research interest, which is summarized chronologically in Table 10. Herein, an optical fibre sensor for Hg2+ and Cu2+ detection was developed using CDs functionalized with poly(ethylene glycol) and N-acetyl-l-cysteine [238]. This sensor has quite interesting analytical potential, as it allows for higher linear coefficients when detecting Hg2+ in aqueous solutions. Furthermore, the concentration range of Hg2+ was successfully lowered to femtomolar detection using a fluorescent sensor [199]. The outstanding nature of the sensing performance can be probably attributed to the Hg2+ ions, which have stronger affinity toward carboxylic groups on the surface of CDs and stronger quenching effect on the PL of the prepared CDs.
On the other hand, Co2+ and Cu2+ ions have been detected using CD-based composites. In the presence of CTAB to PEG-passivated CDs, a long fluorescence lifetime with higher emission efficiency was noticed due to the rapid production of •OH radicals from the Co(II)-H2O2-OH system, which is responsible for the highly selective response of the present chemiluminescence system towards Co2+ ions [200]. On the other hand, the electrochemiluminescence of the oxidized CDs/K2S2O8 system resulted in higher quenching to 20 nM of Cu2+ compared to 1 µM of Pb2+, Ni2+, Mn2+, Fe2+, and Co2+ [201]. Another HMI of interest is Fe3+, which have been detected three times in a row in the year of 2014 using diverse optical sensors [116,117,119]. Among them, the phenolic hydroxyl of N-doped CDs exhibited higher sensitivity toward Fe3+ quenching, with the linearity range of 0.01−500 μM and detection limit of 2.5 nM [116]. The reason is that the hydroxyl group has a good binding affinity for Fe3+, leading to the splitting of d orbital of Fe3+ and consequently causing significant fluorescent quenching. Similarly, nitrogen-doped CQDs were used as an effective fluorescent sensing platform for label-free sensitive detection of Hg2+ ions in an ultrapure water solution with a detection limit of 0.23 μM [118]. Due to the excellent sensitivity and selectivity of the Hg2+ sensor, this sensor was further applied to the determination of Hg2+ in tap water and real lake water samples. The results showed that the PL intensity (excited at 360 nm) decreased gradually with the increasing concentration of Hg2+ in both tap water and lake water from 5 to 50 μM.
In other work, the sensing potential of fluorescent CDs synthesized through the pyrolysis method was evaluated for Cu (II), Cr (II), Co (II), Ni (II), Al (III), Ca (II), Pb (II), Zn (II), Sn (II) and Hg (II) ions [79]. Among all the metal ions tested, Cu (II) and Pb (II) have the highest sensitivity, while Zn (II), Hg (II) and Ca (II) have the lowest sensitivity. The reason for this relatively low sensitivity might be due to the diamagnetic properties of Zn (II), Hg (II) and Ca (II) ions, which eliminate the paramagnetic quenching mechanism with the CDs.
With the accelerating demand of mercury in the skin lightening industry, a high number of optical techniques for Hg2+ have been reported [15,23,54,65,89,122,124,125,126,127,128,133,142,143,152,153,160,178,207]. Notably, fluorescence techniques are the most employed. Altogether, there are several phenomena about the quenching effect of CDs to Hg2+, which are as follows: (i) electron transfer from excited nitrogen and sulfur atoms in CDs to Hg2+ (nonradiative recombination); (ii) strong binding between Hg2+ and S atoms or carboxylate or hydroxyl groups from CDs; and (iii) Hg2+ has a larger ionic radius and polarization; therefore, deformation happened more easily when it interacted with the nitrogen and sulfur atoms. Fortunately, the Hg2+ can be detected as low as 2.6 nM using CD-labeled oligodeoxyribonucleotide (ODN) and quenched by graphene oxide [122]. In this study, CD-ODN acted as the energy donor and molecular recognition probe and GO served as the fluorescence resonance energy transfer acceptor.
Due to the concerns about improving and protecting living organisms, more optical detection methods that utilize CDs as a sensing element have been developed to detect Fe2+/3+ [13,22,38,40,72,78,87,124,135,142,151,159,166,169,219,226], Cu2+ [53,66,97,141,227,239], Pb2+ [65,87,181,240], Cr6+/4+ [23,85,193,241], and Co2+ [51]. For instance, Shamsipur et al. (2018) prepared green-emitting CDs instead of common blue-emitting CDs for selective and sensitive detection of Fe3+ [159]. This green CDs approach could detect Fe3+ in the range of 0.05–10.0 μM, with a detection limit of 0.0137 μM. The detection limit was found to be significantly lower than the maximum level of Fe3+ (0.54 μM) allowed in drinking water by the U.S. Environmental Protection Agency.
Li et al. (2021) invested in interesting research to detect low concentrations of Pb2+ ions [181]. An electrochemiluminescence (ECL) aptasensor was prepared by modifying a glassy carbon electrode (GCE)/NCQDs with amino aptamers by NHS/EDC linkage for selective and sensitive detection of Pb2+. As shown in Figure 7, the binding of the hairpin aptamer with Pb2+ forms the G-quadruplex, and thus exposes the amino group to the NCQDs, causing significant changes in ECL intensity. The results revealed that the GCE/NCQDs/aptamer sensor can detect Pb2+ quickly and accurately, providing the lowest detection limit of 18.9 pM. Importantly, the proposed aptasensor did not suffer from interference and it had excellent stability.
The fabrication of CDs with industrial/agricultural waste has proven to be useful as a detection probe for metal ions. A study conducted by Yan et al. (2020) demonstrated that the synthesized N-CDs from crown daisy leaves had abundant surface functional groups that could selectively and sensitively detect Cu2+ at a very low detection limit of 1.0 nM by the fluorescent quenching effect [53]. This result is due to the selective complexation interaction between Cu2+ and carboxyl and amino groups of the N-CDs.
On the other hand, improving the detection performance of hexavalent chromium (Cr6+) even at the nanomolar level has become an enduring concern of researchers. Rajendran and Rajendiran (2018) reported that the CQDs can be effectively used as fluorescent probes for the detection of Cr6+ at the nanomolar level [23]. This research reported strong fluorescence quenching with increasing Cr6+ concentration, resulting in an improved detection limit of 10 nm. Nevertheless, the detection limit of Co2+ was found to be higher at the micromolar level despite using CD-based optical sensors [51]. The proposed sensor was then applied in the analysis of real river water samples with a recovery value of 95.0–106.8% and a relative standard deviation below 5.3%, indicating that the synthesized CDs had great application prospects in Co2+ detection.
Table 10. Summary of the developed optical sensors for heavy metal ions detection.
Table 10. Summary of the developed optical sensors for heavy metal ions detection.
Heavy Metal IonsMaterialOptical SensorRange of DetectionLimit of DetectionLinear Correlation
Coefficient
YearReference
Hg2+CDs@PEG and N-acetyl-l-cysteineOptical fibre0–2.69 μM-0.9772010[238]
Cu2+0.975
Hg2+CDsFluorescent0–5 fM1 fM-2013[199]
Co2+CTAB@CDsChemiluminescent1.0–1000 nM0.67 nM0.9922013[200]
Cu2+o-CDs/K2S2O8Electrochemiluminescent0–4 nM--2013[201]
Fe3+N-CDsChemiluminescent1.0 × 10−7–1.0 × 10−6 M66.7 nM0.9932014[117]
Fe3+N-doped CDsFluorescent0.01–500 μM2.5 nM-2014[116]
Fe3+CDsElectrochemiluminescent5–80 μM700 nM0.9932014[119]
Hg2+N-CQDsFluorescent0–25 μM0.23 μM0.9942014[118]
Pb2+CDsFluorescent0–47.62 μM7.49 μM-2014[79]
Cu2+7.78 μM
Al3+13.38 μM
Ni2+13.90 μM
Co2+18.07 μM
Cr2+23.69 μM
Sn2+31.51 μM
Ca2+34.79 μM
Hg2+38.02 μM
Zn2+69.64 μM
Fe3+CDsFluorescent0.10–10 μM31.5 nM0.99772015[124]
Hg2+0.01–2.0 μM15.3 nM0.9977
Hg2+CDsFluorescent0.01–10 μM3.3 nM0.9972015[128]
Hg2+ODN-CDsFluorescent5–200 nM2.6 nM0.9742015[122]
Hg2+N-S-CDsFluorescent0–40 μM2.0 μM0.9942015[125]
Hg2+N-CDsFluorescent0–8 μM0.087 μM0.99622015[127]
Hg2+N,S-co-doped CDsFluorescent0–20 μM0.18 μM0.99752015[126]
Hg2+N-rich CDsFluorescent0–20 μM0.63 μM0.9892016[207]
Fe3+N-doped CDsFluorescent0–1000 μM100 μM-2016[135]
Fe3+N-CDsFluorescent0–1000 μM0.96 μM-2016[13]
Hg2+LR-CDsFluorescent0.1–1.5 μM18.7 nM0.99192016[64]
2.0–60.0 μM0.994
Hg2+CDsFluorescent0–80 μM0.201 μM0.99822016[133]
Pb2+CDsFluorescent0.01–1.0 μM0.59 nM0.9982017[65]
Cu2+Nitrogen-doped CDsFluorescent0.5–4 μM0.38 μM0.9982017[141]
Hg2+CDsFluorescent0–0.5 mM0.78 μM0.99442017[142]
Fe3+0–0.15 mM1.17 μM0.9977
Hg2+Nitrogen-doped CQDsFluorescent0–18 μM83.5 nM0.99792017[143]
Hg2+CDsFluorescent0–40 μM9 nM0.98962017[15]
Fe3+N-CQDsFluorescent0–300 μM0.16 μM0.98112018[22]
Cr6+CDsFluorescent0–100 μM0.73 μM0.99032018[85]
Cu2+CQDsFluorescent1–8 μM6.33 nM0.9982018[97]
Cu2+CDsFluorescent0.01–500 μM4.3 nM0.99072018[66]
Cu2+CQDsFluorescent0–100 μM31.5 μM0.98972018[239]
Fe3+CdSe@SiO2-CDsFluorescent9–120 μM0.26 μM0.9952018[151]
Fe3+CDsFluorescent0.05–10.0 μM13.7 nM0.9922018[159]
Hg2+N-CDsFluorescent0.001–5 μM0.65 μM0.9852018[160]
Hg2+CDsFluorescent0–100 μM2.47 μM0.98922018[152]
Hg2+N-S-CDsFluorescent0.01–50 μM0.008 μM0.96222018[153]
Hg2+CQDsFluorescent5–70 nM8 nM0.99702018[23]
Cr6+10 nM0.9956
Fe3+CDsFluorescent1–700 μM<1 μM0.9932019[219]
Fe2+CDsOptical microfiber0–5.372 μM0.179 μM-2019[166]
Co2+CDsFluorescent1–2 μM0.39 μM0.99122019[51]
Pb2+VV-CDsFluorescent1–100 μM12 nM0.998532020[87]
Fe3+16 nM0.99933
Pb2+GCE/NCQDs/aptamersElectrochemiluminescence50–387.9 nM0.0189 nM0.9982020[181]
As3+CDs-MnO2Fluorescent0–200 nM16.8 nM0.9922020[177]
Cu2+CDsFluorescent0–120 nM1.0 nM0.9972020[53]
Fe3+Phe-CDsFluorescent5–500 μM0.720 μM0.99592020[169]
Hg2+N-CDsFluorescent0.15–90 μM0.20 μM0.9932020[178]
Hg2+CDsFluorescent0.01–5 μM6.25 nM0.9912020[54]
Cr4+S, N-CDsFluorescent0.03–50 μM21.14 nM0.9962021[193]
Fe3+N-CDsFluorescent0.3–3.3 μM0.135 μM0.99182021[72]
Fe3+CDs@PDAFluorescent2–27 μM3.75 μM
5.82 μM
0.994
0.991
2021[226]
Pb2+N-CDs/R-CDs@ZIF-8Fluorescent0.05–50 μM4.78 nM0.99522021[240]
Cr6+BNCDsFluorescent0–100 μM0.41 μM0.9992021[241]
Fe3+KBNCDsFluorescent0–25 μM1.2 μM0.9972022[78]
Mn2+1.4 μM0.998
Hg2+CDsFluorescent0–46 μM2 μM0.9972022[89]
Fe3+CDsFluorescent20–100 μM0.07 μM0.99772022[38]
Fe3+M-CDsFluorescent5–30 μM0.47 μM0.9982022[40]
Cu2+NS-CDsColorimetric1–100 μM200 nM0.994812022[227]
Where CDs@PEG: carbon dots functionalized poly(ethylene glycol), CTAB: cationic cetyltri-methylammonium bromide, o-CDs/K2S2O8: oxidized carbon dots/potassium persulfate, N-CDs: nitrogen doped carbon dots, N-CQDs: nitrogen doped carbon quantum dots, ODN-CDs: carbon dots-labeled oligodeoxyribonucleotide, N,S-CDs: nitrogen and sulphur co-doped carbon dots, LR: lotus root, CdSe@SiO2-CDs: silica coated cadmium selenide carbon dots, VV-CDs: Volvariella volvacea-carbon dots, GCE: glassy carbon electrode, MnO2: manganese dioxide, Phe-CDs: phenylalanine-carbon dots, CDs@PDA: p-phenylenediamin-derived carbon dots, N-CDs/R-CDs@ZIF-8: nitrogen-doped carbon dots/red-carbon dots@zeolite imidazole, BNCDs: boron-nitrogen co-doped carbon dots, KBNCDs: Kentucky bluegrass nitrogen-doped carbon dots, M-CDs: Morus nigra carbon dots, Hg2+: mercury ions, Cu2+: copper ions, Co2+: cobalt ions, Fe3+/2+: ferric ions, Pb2+: lead ions, Al3+: aluminum ions, Ni2+: nickel ions, Cr2+/4+/6+: chromium ions, Sn2+: tin ions, Ca2+: calcium ions, Zn2+: zinc ions, As3+: arsenic ions.

4.2. Phenols

The existence of different forms of phenolic in petrochemical engineering, printing, food processing, and dyeing has been continuously reported as the most harmful pollutant due to its high toxicity and low biodegradability [242]. Table 11 summarizes the chronological order of the detection of phenolic compounds using optical sensors. Among the reported studies, 2-4-6-trinitrophenol, also called picric acid, is the most common pollutant monitored by means of fluorescent sensor [72,112,114,134,136,150,170,221]. It is well known that 2-4-6-trinitrophenol is a representative electron-deficient nitroaromatic, due to the electron-withdrawing nature of the three nitro groups. From the perspective of the sensor performance, Niu et al. (2013) explored the feasibility of using amine-capped CDs for the detection of 2-4-6-trinitrophenol in aqueous and ethanol solutions [114]. For the detection of 2-4-6-trinitrophenol in ethanol, the amine-capped CDs demonstrated obvious fluorescent quenching and superior sensitivity to 2-4-6-trinitrophenol in aqueous solution. Since water is much more polar than ethanol, the detection is, therefore, driven by an electrostatic interaction between 2-4-6-trinitrophenol and CDs.
Generally, there are three possible mechanisms causing fluorescence quenching between the phenolic compounds and proposed CDs, which are as follows: (i) electron transfer; (ii) fluorescence energy transfer; and (iii) inner filter effect. Amid all the mechanisms, fluorescence energy transfer (FET) is a dominant quenching pathway due to the electron-rich nature of the CDs and the electron-deficient aromatic group of the phenol [115,134,243]. In contrast to FET, the electron transfer process became possible for phenol fluorescence quenching [112,150,204]. Notably, electron transfer in the excited state of CDs can occur from the lowest unoccupied molecular orbital (LUMO) to the LUMO of phenols, if possible. In addition, there are certain rules that are required to fulfil FET quenching. The rules are as follows: (i) the absorption spectrum of the quencher (phenols) should overlap with the emission spectrum fluorophore and (ii) there is a change in the fluorescence lifetime of CDs before and after the addition of the quencher. Furthermore, the inner filter effect (IFE) is also considered important and has been applied as a developing novel fluorescent assay in recent studies [34,136,170,218,219,226]. This approach offers more simplicity and expediency because IFE-based sensors use fluorophore and the receptor directly rather than chemically interacting with each other.
In another study, Xue et al. (2018) found that phenol can form hydrogen bonds with the carboxyl group on the surface of the fluorescence CQDs, which will facilitate the fluorescence quenching of the CQDs in the excited state, leading to the high sensitivity and selectivity of CQDs for the detection of phenol [95]. In addition, the fluorescence quenching effect can be explained by either a dynamic or static process. Interestingly, a static quenching effect between N-CDs and tannic acid found in the works of [167] can be proved by several measurements, which are as follows: (i) a decrease in the fluorescence lifetime; (ii) a rate that is constant more than 1.0 × 1010 M−1s−1; and (iii) a blue-shift of the absorption peak of CDs. Wang et al. (2019) compared the selectivity of the CDs@MIP and CDs@MIP to 4-nitrophenol, which resulted in good selectivity of CDs@MIP on 4-nitrophenol, due to the hydrogen bonding interactions, leading to the charge transfer from CDs to 4-nitophenol [165]. There is still much room for further development of CD-based optical sensors for the detection of phenolic compounds, despite achieving a high quantum yield and good sensing performance. For instance, Saravanan et al. (2020) successfully synthesized N@CDs from m-phenylenediamine as a single source of carbon and nitrogen to detect 2-4-6-trinitrophenol [221]. The fluorescent N@CD probe showed high selectivity toward 2-4-6-trinitrophenol relative to other nitro explosives, such as 4-nitrophenol, 4-amino-3-nitrophenol, 2,5-dinitrophenol, phenol, 2-chlorophenol, 3-nitrophenol, and 4-amino-triphenol.
In addition, ultra-weak chemiluminescent (CL) systems have become one of the focuses of the increasing attention on the determination of tannic acid [173]. Interestingly, tannic acid could dramatically suppress the CL intensity of the FNCDs-H2O2-K3Fe(CN)6 system with an excellent linear response (R2 = 0.9971) and was further applied for the determination of tannic acid in a red wine sample. The recovery of 98.9–102.1% was obtained, demonstrating its reliability and application potential in real sample analysis. Recently, the aggregated N,S-CDs as a dual-excitation ratiometric fluorescent probe have proven to be an effective approach for the quantitative determination of chlorogenic acid [184]. Apart from dual excitation, Liu et al. (2021) observed the dual-color fluorescence emission of ultraviolet and red light when detecting 4-nitrophenol using CDs@PDA [226]. The detection limits obtained for the ultraviolet and red-light emission were 3.44 µM and 7.29 µM, respectively. In addition, 2-4-6-trinitrophenol has been detected using the fluorescence method with a detection limit of 0.11 μM in the concentration range of 0.3–3.3 μM [72]. The possible mechanism for fluorescence quenching of N-CDs by 2-4-6-trinitrophenol is explained by the electrostatic interaction or hydrogen bonding that occurs between the nitrogen groups of CDs and phenolic group of 2-4-6-trinitrophenol. To be exact, the electrostatic interactions occur when a large number of nitro groups with electron-withdrawing characteristics on benzene interact with nitrogen-containing functional groups of CDs. Hence, it can be concluded that nitro groups play a vital role in fluorescent quenching.

4.3. Pesticides

Pesticides are chemical compounds that are used to kill pests. Generally, there are six classes of pesticides according to their target species, i.e., insecticides, herbicides, rodenticides, bactericides, fungicides, and larvicides. The largest and most widely used in crop protection are insecticides [244]; however, they are classified as an extremely toxic class of chemical compounds by the World Health Organization (WHO).
Organophosphorus (OP) compounds, such as methyl parathion [121,138], paraoxon-ethyl [137], dichlorvos [245], malathion [245], ethion [245], paraoxon [154,155,211], chlorpyrifos [63,96,246], methyl-paraoxon [180], diazinon [50,62], and quinalphos [247], [63], are widely reported insecticides in the development of CD-based optical sensors. Oddly, most OP pesticides are always selected as inhibitors for sensors based on enzyme activity. According to Hou et al. (2016), dichlorvos can inhibit the activity of acetyl cholinesterase (AChE), which can turn the fluorescence of CDs off again [245], which is similar to the detection of paraoxon [155]. However, in a study of Lin et al. (2017), only OP of chlorpyrifos effectively inhibited H2O2 production by destroying the acetylcholinesterase activity, thereby increasing the fluorescence of C-dots after being initially turned off by Fe3+ [96]. Moreover, the proposed sensor exhibits a low selectivity against OP. Therefore, a new facile fluorescence probing based on novel N-doped carbon dots and methyl parathion hydrolase (MPH) was developed to detect methyl parathion selectively [138]. The use of MPH in this study helps to degrade OP compounds and has many advantages over other enzymes, such as less background noise and high turnover number, which is favourable for sensitive detection.
Another class of insecticide is carbamates. Li and co-workers have been devoted to detecting carbaryl in the presence of acetylcholinesterase (ACh) and choline oxidase (ChOx) by using N-CQDs, S-CQDs, and co-doped N, S-CQD-based photoluminescence sensors [209]. The results demonstrated that high sensitivity originates from the N and S dopants, which offer stronger capacities to adsorb hydrogen peroxide (H2O2) and generate local states to trap hot electrons, promoting the electron transfer to H2O2, and thus resulting in the strong quenching of the CQD fluorescence.
Apart from the insecticide organophosphate, insecticides neonicotinoids and organochlorine have become a target pesticide. Mandal et al. (2019) have reported a simple fluorescence-based method using CDs for the detection of commonly used insecticides, such as imidacloprid and lindane [246]. The fluorescence emission of the CDs was found to be enhanced in the presence of the increasing concentration of imidacloprid; however, the emission was quenched in the presence of lindane. The interactions between CDs and insecticides were mainly driven by the high reactive -NO2 and -NH groups in imidacloprid that directly bind to the surface functionality amino groups of CDs and by the H-bond or week ionic interaction or covalent binding between the free surface groups of the CDs and lindane, followed by the substitution reaction. In addition, the researchers also expanded the potential of CDs in sensing unclassified pesticides, namely tetradifon [246]. The leaving groups (-Cl) of tetradifron were found to possessstrong covalent binding with the surface amine group of the CDs with a high affinity. Nonetheless, the results showed that tetradifron has a higher detection limit than imidacloprid, but offers a lower detection limit than lindane. This phenomenon can be attributed to the binding affinity of all the different target pesticides. Their binding affinity can be arranged in the following decreasing order: imidacloprid > tetradifron > lindane.
Until 2021, the fluorescence response of CDs towards herbicides [50,60,213,220,246] and fungicides [73,182,183] has also been explored. It is reported that with increasing concentrations of herbicides, i.e., atrazine [213], pretilachlor [60], glyphosate [50], and fungicide thiophanate methyl [183], this has resulted in an increase in PL emission intensity (turn-on). This fluorescence ‘turn-on’ model was dynamic quenching, whereas, when the concentrations of trifluralin [220], pyrimethanil [182], and isoprothiolane [73] were increased, the fluorescence emission of CDs was found to decrease, leading to static quenching. Supchocksoonthorn et al. (2021) suggested that the fluorescent quenching is closely related to the inner filter effect, due to the π-π interaction between the large aromatic core of the CDs and anilinopyrimidine unit of pyrimethanil and between their polar functional groups [182].
Furthermore, as an improvement to the method with a single fluorescence signal, Zhu et al. (2022) developed a new type of CD-functionalized core-shell nanospheres as a ratiometric fluorescence to detect λ-cyhalothrin (LC), a typical pyrethroid insecticide. Figure 8 shows the main binding reactions in the preparation of the CD-functionalized core-shell nanospheres. Here, silica nanoparticles play a role in preventing the contact between m-CDs and the dispersion solvent so the m-CDs could act as a reliable reference background; o-CDs act as a fluorescence signal; 3-aminopropyrtriethoxy silane (APTES) as a functional monomer to connect with LC through hydrogen bonding; ionic liquids (IL) as a cationic surfactant, which are attached to the system by electrostatic interaction to enhance the fluorescence of the core-shell nanospheres; and tetraethyl orthosilicate (TEOS) is used to form a layer of imprinted silicon shell on the surface of m-CDs@SiO2. Inspired by the role of materials, these newly formed core-shell nanospheres managed to achieve lower detection limits, excellent reusability, and better precision than those of other methods reported in this work [230]. All the proposed CD-based optical sensors for the detection of pesticides are tabulated orderly in Table 12.

4.4. Explosive Compounds

Military-grade explosives such as trinitroluene (TNT), dinitroaniline (DNT), are still a major worldwide concern in terms of terror threat and environmental impact. The only method that detects explosive compounds using CDs is a fluorescent sensor. For instance, Zhang et al. (2015) used N-rich CDs for 2,4,6-trinitrotoluene determination [120]. Apparently, TNT is a typical electron deficient nitro compound, which can selectively interact with amine group on CDs, leading to charge transfer from fluorescent CDs to the aromatic rings of TNT and consequently quenching the fluorescence emission of CDs strongly.
Dai and co-workers found that fluorescence sensors are well suited for detecting nitroaromatic compounds, such as 2,4-dinitrotoluene (DNT), as they can quench the emission of the excited species [132]. In this work, static quenching occurred with an increasing concentration of DNT, as CDs formed stable charge-transfer complexes with DNT molecules. Considering that the selectivity of those sensors were relatively low because some metal ions or other structure analogy can quench those fluorescent nanoparticles, Xu et al. introduced molecularly imprinted polymers (MIPs) to improve the selectivity [131]. However, this causes poor sensitivity because of the highly cross-linked nature of MIPs. Therefore, a novel strategy using periodic mesoporous silica particles as the imprinting matrix and using amino-CDs directly as a “functional monomer” was proposed to improve the sensitivity of M-MIPs@CDs.
In addition, Campos et al. (2016) first demonstrated the fluorometric sensing of 4-chloro-2,6-dinitroaniline in an aqueous solution using carbon quantum dots coated with PAMAM-NH2 [208]. The presence of amine groups in the PAMAM-NH2 dendrimer results in the enhanced fluorescence intensity and quantum yield from 3.94% to 6.93%. Moreover, it is interesting that with an increase in pH starting from 2.25, the fluorescence intensity increases drastically and then become constant from pH 6 to 11. The reason for the constant fluorescence intensity may be due to the buffering effect of the surface groups (–COOH and –NH2 mainly).
Next, a selective and sensitive detection of TNT explosive residues based on CD-modified optical sensors has been developed and rapidly elevating until 2022 [99,139,196,214,248]. The research concluded that the possible mechanism for TNT detection corresponds to the Jackson–Meisenheimer (JM) complex. Particularly, functionalizing CDs with amino groups can selectively form a JM complex with typical electron-deficient groups in TNT via charge transfer during a nucleophilic aromatic substitution reaction, which consequently causes fluorescence quenching. In addition, the selectivity of the sensor against TNT was investigated by Devi’s group [214]. The proposed sensor showed high selectivity for TNT over other weaker Lewis acids, such as dinitrotoulene (DNT), 2,4-dinitrotoluic acid (TA), nitrobenzene (NB), and p-nitrophenol (NP). It can be concluded that the JM complex formed by TNT is more stabilized due to more resonance structures; more nitro groups on the TNT aromatic ring than that formed by DNT, TA, NB, and NP. All the proposed CD-based optical sensors for the detection of explosive compounds are tabulated orderly in Table 13.

5. Concluding Remarks

An enormous research effort has been devoted to the development of CDs using green and chemical precursors. The reason behind the interest in this fabrication is its fluorescent properties that make it continuously competitive to produce bright and high-quality CDs. This paper has presented the recent progress in the field of CDs, focusing on the green and chemical precursors and fluorescent properties. To the best of our knowledge, this is the first review paper that highlights each of the precursors used in the CDs with the values of the fluorescence quantum yield.
Among the available green precursors, CDs produced by plants are the most popular source because it is a more stable and straightforward process to scale up. For instance, the CDs synthesized from leaves, i.e., Calotropis procera leaves, without any addition of surface passivation have the highest quantum yield of 71.95% under the hydrothermal carbonization process, as shown in Figure 9a. It can be observed that the leaves have shown great potential as carbon sources due to their carbon-, hydrogen-, oxygen-, nitrogen-rich composition, which can eliminate the need of surface-chemical passivating agents. Other factors that contributed to the excellent photoluminescence properties include (i) heteroatom doping, such as sulfur, nitrogen, phosphorus, and boron; (ii) heating temperature and time; (iii) type and ratio of solvent; and (iv) pH of media. Even so, the quantum yields obtained from the green precursors were still low compared to the chemical precursors, as depicted in Figure 9b. It was shown that the thermal pyrolysis treatment of sodium citrate dihydrate and urea led to green fluorescence CDs with a quantum yield as high as 93%. As expected, CDs synthesized from acid reagents possess higher quantum yields due to the existence of carboxyl groups. The greater the number of carboxyl groups in the carbon source, the more amino groups of urea that will conjugate onto the surface of CDs.
Although the highest quantum yield results have been achieved using chemical precursors (acid reagents), the high toxicity of this precursor may hinder their potential for diverse applications. Therefore, future development should firstly focus on synthesizing CDs from green precursors, particularly plant parts and waste products. In addition, it is highly necessary to fabricate green precursors consisting of hydroxyl, carboxyl and amino groups to avoid the use of toxic chemical passivating agents. Furthermore, this will allow for the development of high-quality and non-toxicity CDs, finally paving the way towards their practical application in various fields.

Author Contributions

Conceptualization and writing—original draft preparation, N.A.S.O.; supervision and funding acquisition, Y.W.F. and R.I.; writing—review and editing, H.S.H. and N.S.M.R.; visualization, N.I.M.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Putra Grant Universiti Putra Malaysia (GP-IPB/2021/9700700).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Atchudan, R.; Edison, T.N.J.I.; Perumal, S.; Vinodh, R.; Sundramoorthy, A.K.; Babu, R.S.; Lee, Y.R. Leftover Kiwi Fruit Peel-Derived Carbon Dots as a Highly Selective Fluorescent Sensor for Detection of Ferric Ion. Chemosensors 2021, 9, 166. [Google Scholar] [CrossRef]
  2. Carbonaro, C.M.; Corpino, R.; Salis, M.; Mocci, F.; Thakkar, S.V.; Olla, C.; Ricci, P.C. On the Emission Properties of Carbon Dots: Reviewing Data and Discussing Models. C—J. Carbon Res. 2019, 5, 60. [Google Scholar] [CrossRef] [Green Version]
  3. Shi, W.; Song, S.; Zhang, H. Hydrothermal Synthetic Strategies of Inorganic Semiconducting Nanostructures. Chem. Soc. Rev. 2013, 42, 5714–5743. [Google Scholar] [CrossRef]
  4. Li, L.; Dong, T. Photoluminescence Tuning in Carbon Dots: Surface Passivation or/and Functionalization, Heteroatom Doping. J. Mater. Chem. C 2018, 6, 7944–7970. [Google Scholar] [CrossRef]
  5. Jorns, M.; Pappas, D. A Review of Fluorescent Carbon Dots, Their Synthesis, Physical and Chemical Characteristics, and Applications. Nanomaterials 2021, 11, 1448. [Google Scholar] [CrossRef]
  6. Long, C.; Jiang, Z.; Shangguan, J.; Qing, T.; Zhang, P.; Feng, B. Applications of Carbon Dots in Environmental Pollution Control: A Review. Chem. Eng. J. 2021, 406, 126848. [Google Scholar] [CrossRef]
  7. Sharma, V.; Tiwari, P.; Kaur, N.; Mobin, S.M. Optical Nanosensors Based on Fluorescent Carbon Dots for the Detection of Water Contaminants: A Review. Environ. Chem. Lett. 2021, 19, 3229–3241. [Google Scholar] [CrossRef]
  8. Kurian, M.; Paul, A. Recent Trends in the Use of Green Sources for Carbon Dot Synthesis—A Short Review. Carbon Trends 2021, 3, 100032. [Google Scholar] [CrossRef]
  9. Sahu, S.; Behera, B.; Maiti, T.K.; Mohapatra, S. Simple One-Step Synthesis of Highly Luminescent Carbon Dots from Orange Juice: Application as Excellent Bio-Imaging Agents. Chem. Commun. 2012, 48, 8835–8837. [Google Scholar] [CrossRef]
  10. Zhou, J.; Sheng, Z.; Han, H.; Zou, M.; Li, C. Facile Synthesis of Fluorescent Carbon Dots Using Watermelon Peel as a Carbon Source. Mater. Lett. 2012, 66, 222–224. [Google Scholar] [CrossRef]
  11. Du, F.; Zhang, M.; Li, X.; Li, J.; Jiang, X.; Li, Z.; Hua, Y.; Shao, G.; Jin, J.; Shao, Q.; et al. Economical and Green Synthesis of Bagasse-Derived Fluorescent Carbon Dots for Biomedical Applications. Nanotechnology 2014, 25, 315702. [Google Scholar] [CrossRef]
  12. Tyagi, A.; Tripathi, K.M.; Singh, N.; Choudhary, S.; Gupta, R.K. Green Synthesis of Carbon Quantum Dots from Lemon Peel Waste: Applications in Sensing and Photocatalysis. RSC Adv. 2016, 6, 72423–72432. [Google Scholar] [CrossRef]
  13. Edison, T.N.J.I.; Atchudan, R.; Shim, J.J.; Kalimuthu, S.; Ahn, B.C.; Lee, Y.R. Turn-off Fluorescence Sensor for the Detection of Ferric Ion in Water Using Green Synthesized N-Doped Carbon Dots and Its Bio-Imaging. J. Photochem. Photobiol. B Biol. 2016, 158, 235–242. [Google Scholar] [CrossRef]
  14. Shi, J.; Ni, G.; Tu, J.; Jin, X.; Peng, J. Green Synthesis of Fluorescent Carbon Dots for Sensitive Detection of Fe2+ and Hydrogen Peroxide. J. Nanopart. Res. 2017, 19, 209. [Google Scholar] [CrossRef]
  15. Zhao, J.; Huang, M.; Zhang, L.; Zou, M.; Chen, D.; Huang, Y.; Zhao, S. Unique Approach to Develop Carbon Dot-Based Nanohybrid Near-Infrared Ratiometric Fluorescent Sensor for the Detection of Mercury Ions. Anal. Chem. 2017, 89, 8044–8049. [Google Scholar] [CrossRef]
  16. Su, A.; Wang, D.; Shu, X.; Zhong, Q.; Chen, Y.; Liu, J.; Wang, Y. Synthesis of Fluorescent Carbon Quantum Dots from Dried Lemon Peel for Determination of Carmine in Drinks. Chem. Res. Chin. Univ. 2018, 34, 164–168. [Google Scholar] [CrossRef]
  17. Ding, H.; Ji, Y.; Wei, J.S.; Gao, Q.Y.; Zhou, Z.Y.; Xiong, H.M. Facile Synthesis of Red-Emitting Carbon Dots from Pulp-Free Lemon Juice for Bioimaging. J. Mater. Chem. B 2017, 5, 5272–5277. [Google Scholar] [CrossRef]
  18. Chatzimitakos, T.; Kasouni, A.; Sygellou, L.; Avgeropoulos, A.; Troganis, A.; Stalikas, C. Two of a Kind but Different: Luminescent Carbon Quantum Dots from Citrus Peels for Iron and Tartrazine Sensing and Cell Imaging. Talanta 2017, 175, 305–312. [Google Scholar] [CrossRef]
  19. Gharat, P.M.; Pal, H.; Dutta Choudhury, S. Photophysics and Luminescence Quenching of Carbon Dots Derived from Lemon Juice and Glycerol. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2019, 209, 14–21. [Google Scholar] [CrossRef]
  20. Schneider, E.M.; Bärtsch, A.; Stark, W.J.; Grass, R.N. Safe One-Pot Synthesis of Fluorescent Carbon Quantum Dots from Lemon Juice for a Hands-On Experience of Nanotechnology. J. Chem. Educ. 2019, 96, 540–545. [Google Scholar] [CrossRef]
  21. Tadesse, A.; RamaDevi, D.; Battu, G.; Basavaiah, K. Facile Green Synthesis of Fluorescent Carbon Quantum Dots from Citrus Lemon Juice for Live Cell Imaging. Asian J. Nanosci. Mater. 2018, 1, 36–46. [Google Scholar]
  22. Lu, M.; Duan, Y.; Song, Y.; Tan, J.; Zhou, L. Green Preparation of Versatile Nitrogen-Doped Carbon Quantum Dots from Watermelon Juice for Cell Imaging, Detection of Fe3+ Ions and Cysteine, and Optical Thermometry. J. Mol. Liq. 2018, 269, 766–774. [Google Scholar] [CrossRef]
  23. Rajendran, K.; Rajendiran, N. 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 2018, 5, 24008. [Google Scholar] [CrossRef]
  24. Ahmadian-Fard-Fini, S.; Salavati-Niasari, M.; Ghanbari, D. Hydrothermal Green Synthesis of Magnetic Fe3O4-Carbon Dots by Lemon and Grape Fruit Extracts and as a Photoluminescence Sensor for Detecting of E. Coli Bacteria. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2018, 203, 481–493. [Google Scholar] [CrossRef]
  25. Anindita, F.; Darmawan, N.; Mas’Ud, Z.A. Fluorescence Carbon Dots from Durian as an Eco-Friendly Inhibitor for Copper Corrosion. AIP Conf. Proc. 2018, 2014, 020008. [Google Scholar] [CrossRef]
  26. Carvalho, J.; Santos, L.R.; Germino, J.C.; Terezo, A.J.; Moreto, J.A.; Quites, F.J.; Freitas, R.G. Hydrothermal Synthesis to Water-Stable Luminescent Carbon Dots from Acerola Fruit for Photoluminescent Composites Preparation and Its Application as Sensors. Mater. Res. 2019, 22, 1–8. [Google Scholar] [CrossRef] [Green Version]
  27. Fatahi, Z.; Esfandiari, N.; Ehtesabi, H.; Bagheri, Z.; Ranjbar, Z.; Latifi, H. Physicochemical and Cytotoxicity Analysis of Green Synthesis Carbon Dots for Cell Imaging. EXCLI J. 2019, 18, 454–466. [Google Scholar]
  28. Tadesse, A.; Hagos, M.; Ramadevi, D.; Basavaiah, K.; Belachew, N. Fluorescent-Nitrogen-Doped Carbon Quantum Dots Derived from Citrus Lemon Juice: Green Synthesis, Mercury(II) Ion Sensing, and Live Cell Imaging. ACS Omega 2020, 5, 3889–3898. [Google Scholar] [CrossRef] [Green Version]
  29. Monte-Filho, S.S.; Andrade, S.I.E.; Lima, M.B.; Araujo, M.C.U. Synthesis of Highly Fluorescent Carbon Dots from Lemon and Onion Juices for Determination of Riboflavin in Multivitamin/Mineral Supplements. J. Pharm. Anal. 2019, 9, 209–216. [Google Scholar] [CrossRef]
  30. Hoan, B.T.; Tam, P.D.; Pham, V.H. Green Synthesis of Highly Luminescent Carbon Quantum Dots from Lemon Juice. J. Nanotechnol. 2019, 2019, 2852816. [Google Scholar] [CrossRef] [Green Version]
  31. Jayaweera, S.; Yin, K.; Ng, W.J. Nitrogen-Doped Durian Shell Derived Carbon Dots for Inner Filter Effect Mediated Sensing of Tetracycline and Fluorescent Ink. J. Fluoresc. 2019, 29, 221–229. [Google Scholar] [CrossRef]
  32. Hashemi, F.; Heidari, F.; Mohajeri, N.; Mahmoodzadeh, F.; Zarghami, N. Fluorescence Intensity Enhancement of Green Carbon Dots: Synthesis, Characterization and Cell Imaging. Photochem. Photobiol. 2020, 96, 1032–1040. [Google Scholar] [CrossRef]
  33. Muktha, H.; Sharath, R.; Kottam, N.; Smrithi, S.P.; Samrat, K.; Ankitha, P. Green Synthesis of Carbon Dots and Evaluation of Its Pharmacological Activities. Bionanoscience 2020, 10, 731–744. [Google Scholar] [CrossRef]
  34. Zhang, Q.; Liang, J.; Zhao, L.; Wang, Y.; Zheng, Y.; Wu, Y.; Jiang, L. Synthesis of Novel Fluorescent Carbon Quantum Dots From Rosa roxburghii for Rapid and Highly Selective Detection of O-Nitrophenol and Cellular Imaging. Front. Chem. 2020, 8, 1–9. [Google Scholar] [CrossRef]
  35. Gudimella, K.K.; Appidi, T.; Wu, H.F.; Battula, V.; Jogdand, A.; Rengan, A.K.; Gedda, G. Sand Bath Assisted Green Synthesis of Carbon Dots from Citrus Fruit Peels for Free Radical Scavenging and Cell Imaging. Colloids Surf. B Biointerfaces 2021, 197, 111362. [Google Scholar] [CrossRef]
  36. Atchudan, R.; Jebakumar Immanuel Edison, T.N.; Shanmugam, M.; Perumal, S.; Somanathan, T.; Lee, Y.R. Sustainable Synthesis of Carbon Quantum Dots from Banana Peel Waste Using Hydrothermal Process for in Vivo Bioimaging. Phys. E Low-Dimens. Syst. Nanostruct. 2021, 126, 114417. [Google Scholar] [CrossRef]
  37. Ghereghlou, M.; Esmaeili, A.A.; Darroudi, M. Green Synthesis of Fluorescent Carbon Dots from Elaeagnus angustifolia and Its Application as Tartrazine Sensor. J. Fluoresc. 2021, 31, 185–193. [Google Scholar] [CrossRef]
  38. Ashok Varman, G.; Kalanidhi, K.; Nagaraaj, P. Green Synthesis of Fluorescent Carbon Dots from Canon Ball Fruit for Sensitive Detection of Fe3+ and Catalytic Reduction of Textile Dyes. Dyes Pigment. 2022, 199, 110101. [Google Scholar] [CrossRef]
  39. Vijeata, A.; Chaudhary, S.; Chaudhary, G.R. Fluorescent Carbon Dots from Indian Bael Patra as Effective Sensing Tool to Detect Perilous Food Colorant. Food Chem. 2022, 373, 131492. [Google Scholar] [CrossRef]
  40. Atchudan, R.; Edison, T.N.J.I.; Perumal, S.; Vinodh, R.; Sundramoorthy, A.K.; Babu, R.S.; Lee, Y.R. Morus Nigra-Derived Hydrophilic Carbon Dots for the Highly Selective and Sensitive Detection of Ferric Ion in Aqueous Media and Human Colon Cancer Cell Imaging. Colloids Surf. A Physicochem. Eng. Asp. 2022, 635, 128073. [Google Scholar] [CrossRef]
  41. Chandra, S.; Bano, D.; Sahoo, K.; Kumar, D.; Kumar, V.; Kumar Yadav, P.; Hadi Hasan, S. Synthesis of Fluorescent Carbon Quantum Dots from Jatropha Fruits and Their Application in Fluorometric Sensor for the Detection of Chlorpyrifos. Microchem. J. 2022, 172, 106953. [Google Scholar] [CrossRef]
  42. Qu, Y.; Yu, L.; Zhu, B.; Chai, F.; Su, Z. Green Synthesis of Carbon Dots by Celery Leaves for Use as Fluorescent Paper Sensors for the Detection of Nitrophenols. New J. Chem. 2020, 44, 1500–1507. [Google Scholar] [CrossRef]
  43. Yin, B.; Deng, J.; Peng, X.; Long, Q.; Zhao, J.; Lu, Q.; Chen, Q.; Li, H.; Tang, H.; Zhang, Y.; et al. Green Synthesis of Carbon Dots with Down- and up-Conversion Fluorescent Properties for Sensitive Detection of Hypochlorite with a Dual-Readout Assay. Analyst 2013, 138, 6551–6557. [Google Scholar] [CrossRef]
  44. Thota, S.P.; Thota, S.M.; Bhagavatham, S.S.; Manoj, K.S.; Muthukumar, V.S.S.; Venketesh, S.; Vadlani, P.V.; Belliraj, S.K. Facile One-Pot Hydrothermal Synthesis of Stable and Biocompatible Fluorescent Carbon Dots from Lemon Grass Herb. IET Nanobiotechnol. 2018, 12, 127–132. [Google Scholar] [CrossRef]
  45. Miao, H.; Wang, L.; Zhuo, Y.; Zhou, Z.; Yang, X. Label-Free Fluorimetric Detection of CEA Using Carbon Dots Derived from Tomato Juice. Biosens. Bioelectron. 2016, 86, 83–89. [Google Scholar] [CrossRef]
  46. Liu, Y.; Liu, Y.; Park, M.; Park, S.J.; Zhang, Y.; Akanda, M.R.; Park, B.Y.; Kim, H.Y. Green Synthesis of Fluorescent Carbon Dots from Carrot Juice for in Vitro Cellular Imaging. Carbon Lett. 2017, 21, 61–67. [Google Scholar] [CrossRef] [Green Version]
  47. Liu, W.; Diao, H.; Chang, H.; Wang, H.; Li, T.; Wei, W. Green Synthesis of Carbon Dots from Rose-Heart Radish and Application for Fe3+ Detection and Cell Imaging. Sens. Actuators B Chem. 2017, 241, 190–198. [Google Scholar] [CrossRef]
  48. Vasimalai, N.; Vilas-Boas, V.; Gallo, J.; Cerqueira, M.d.F.; Menéndez-Miranda, M.; Costa-Fernández, J.M.; Diéguez, L.; Espiña, B.; Fernández-Argüelles, M.T. Green Synthesis of Fluorescent Carbon Dots from Spices for in Vitro Imaging and Tumour Cell Growth Inhibition. Beilstein J. Nanotechnol. 2018, 9, 530–544. [Google Scholar] [CrossRef]
  49. Li, L.S.; Jiao, X.Y.; Zhang, Y.; Cheng, C.; Huang, K.; Xu, L. Green Synthesis of Fluorescent Carbon Dots from Hongcaitai for Selective Detection of Hypochlorite and Mercuric Ions and Cell Imaging. Sens. Actuators B Chem. 2018, 263, 426–435. [Google Scholar] [CrossRef]
  50. Tafreshi, F.A.; Fatahi, Z.; Ghasemi, S.F.; Taherian, A.; Esfandiari, N. Ultrasensitive Fluorescent Detection of Pesticides in Real Sample by Using Green Carbon Dots. PLoS ONE 2020, 15, 1–17. [Google Scholar] [CrossRef] [Green Version]
  51. Zhao, C.; Li, X.; Cheng, C.; Yang, Y. Green and Microwave-Assisted Synthesis of Carbon Dots and Application for Visual Detection of Cobalt(II) Ions and PH Sensing. Microchem. J. 2019, 147, 183–190. [Google Scholar] [CrossRef]
  52. Kailasa, S.K.; Ha, S.; Baek, S.H.; Phan, L.M.T.; Kim, S.; Kwak, K.; Park, T.J. Tuning of Carbon Dots Emission Color for Sensing of Fe3+ Ion and Bioimaging Applications. Mater. Sci. Eng. C 2019, 98, 834–842. [Google Scholar] [CrossRef] [PubMed]
  53. Xiao-Yan, W.; Xue-Yan, H.; Tian-Qi, W.; Xu-Cheng, F. Crown Daisy Leaf Waste–Derived Carbon Dots: A Simple and Green Fluorescent Probe for Copper Ion. Surf. Interface Anal. 2020, 52, 148–155. [Google Scholar] [CrossRef]
  54. Long, R.; Tang, C.; Li, T.; Tong, X.; Tong, C.; Guo, Y.; Gao, Q.; Wu, L.; Shi, S. Dual-Emissive Carbon Dots for Dual-Channel Ratiometric Fluorometric Determination of PH and Mercury Ion and Intracellular Imaging. Microchim. Acta 2020, 187, 307. [Google Scholar] [CrossRef]
  55. Lai, Z.; Guo, X.; Cheng, Z.; Ruan, G.; Du, F. Green Synthesis of Fluorescent Carbon Dots from Cherry Tomatoes for Highly Effective Detection of Trifluralin Herbicide in Soil Samples. ChemistrySelect 2020, 5, 1956–1960. [Google Scholar] [CrossRef]
  56. Zhang, Z.; Hu, B.; Zhuang, Q.; Wang, Y.; Luo, X.; Xie, Y.; Zhou, D. Green Synthesis of Fluorescent Nitrogen–Sulfur Co-Doped Carbon Dots from Scallion Leaves for Hemin Sensing. Anal. Lett. 2020, 53, 1704–1718. [Google Scholar] [CrossRef]
  57. Rodríguez-Varillas, S.; Fontanil, T.; Obaya, Á.J.; Fernández-González, A.; Murru, C.; Badía-Laíño, R. Biocompatibility and Antioxidant Capabilities of Carbon Dots Obtained from Tomato (Solanum lycopersicum). Appl. Sci. 2022, 12, 773. [Google Scholar] [CrossRef]
  58. Cao, Y.; Wang, X.; Bai, H.; Jia, P.; Zhao, Y.; Liu, Y.; Wang, L.; Zhuang, Y.; Yue, T. Fluorescent Detection of Tetracycline in Foods Based on Carbon Dots Derived from Natural Red Beet Pigment. LWT 2022, 157, 113100. [Google Scholar] [CrossRef]
  59. Sun, R.; Liu, S. Synthesis of Photoluminescent Carbon Dots and Its Effect on Chondrocytes for Knee Joint Therapy Applications. Artif. Cells Nanomed. Biotechnol. 2019, 47, 1321–1325. [Google Scholar] [CrossRef]
  60. Deka, M.J.; Dutta, P.; Sarma, S.; Medhi, O.K.; Talukdar, N.C.; Chowdhury, D. Carbon Dots Derived from Water Hyacinth and Their Application as a Sensor for Pretilachlor. Heliyon 2019, 5, e01985. [Google Scholar] [CrossRef] [Green Version]
  61. Wang, M.; Wan, Y.; Zhang, K.; Fu, Q.; Wang, L.; Zeng, J.; Xia, Z.; Gao, D. Green Synthesis of Carbon Dots Using the Flowers of Osmanthus Fragrans (Thunb.) Lour. as Precursors: Application in Fe3+ and Ascorbic Acid Determination and Cell Imaging. Anal. Bioanal. Chem. 2019, 411, 2715–2727. [Google Scholar] [CrossRef] [PubMed]
  62. Shekarbeygi, Z.; Farhadian, N.; Khani, S.; Moradi, S.; Shahlaei, M. The Effects of Rose Pigments Extracted by Different Methods on the Optical Properties of Carbon Quantum Dots and Its Efficacy in the Determination of Diazinon. Microchem. J. 2020, 158, 105232. [Google Scholar] [CrossRef]
  63. Ghosh, S.; Gul, A.R.; Park, C.Y.; Kim, M.W.; Xu, P.; Baek, S.H.; Bhamore, J.R.; Kailasa, S.K.; Park, T.J. Facile Synthesis of Carbon Dots from Tagetes erecta as a Precursor for Determination of Chlorpyrifos via Fluorescence Turn-off and Quinalphos via Fluorescence Turn-on Mechanisms. Chemosphere 2021, 279, 130515. [Google Scholar] [CrossRef] [PubMed]
  64. Gu, D.; Shang, S.; Yu, Q.; Shen, J. Green Synthesis of Nitrogen-Doped Carbon Dots from Lotus Root for Hg(II) Ions Detection and Cell Imaging. Appl. Surf. Sci. 2016, 390, 38–42. [Google Scholar] [CrossRef]
  65. Kumar, A.; Chowdhuri, A.R.; Laha, D.; Mahto, T.K.; Karmakar, P.; Sahu, S.K. Green Synthesis of Carbon Dots from Ocimum sanctum for Effective Fluorescent Sensing of Pb2+ Ions and Live Cell Imaging. Sens. Actuators B Chem. 2017, 242, 679–686. [Google Scholar] [CrossRef]
  66. Bhamore, J.R.; Jha, S.; Park, T.J.; Kailasa, S.K. Fluorescence Sensing of Cu2+ Ion and Imaging of Fungal Cell by Ultra-Small Fluorescent Carbon Dots Derived from Acacia concinna Seeds. Sens. Actuators B Chem. 2018, 277, 47–54. [Google Scholar] [CrossRef]
  67. Fahmi, M.Z.; Haris, A.; Permana, A.J.; Nor Wibowo, D.L.; Purwanto, B.; Nikmah, Y.L.; Idris, A. Bamboo Leaf-Based Carbon Dots for Efficient Tumor Imaging and Therapy. RSC Adv. 2018, 8, 38376–38383. [Google Scholar] [CrossRef] [Green Version]
  68. Li, W.; Liu, Y.; Wu, M.; Feng, X.; Redfern, S.A.T.; Shang, Y.; Yong, X.; Feng, T.; Wu, K.; Liu, Z.; et al. Carbon-Quantum-Dots-Loaded Ruthenium Nanoparticles as an Efficient Electrocatalyst for Hydrogen Production in Alkaline Media. Adv. Mater. 2018, 30, 1800676. [Google Scholar] [CrossRef]
  69. Wei, X.; Li, L.; Liu, J.; Yu, L.; Li, H.; Cheng, F.; Yi, X.; He, J.; Li, B. Green Synthesis of Fluorescent Carbon Dots from Gynostemma for Bioimaging and Antioxidant in Zebrafish. ACS Appl. Mater. Interfaces 2019, 11, 9832–9840. [Google Scholar] [CrossRef]
  70. Dager, A.; Uchida, T.; Maekawa, T.; Tachibana, M. Synthesis and Characterization of Mono-Disperse Carbon Quantum Dots from Fennel Seeds: Photoluminescence Analysis Using Machine Learning. Sci. Rep. 2019, 9, 14004. [Google Scholar] [CrossRef]
  71. Yang, X.; Wang, D.; Luo, N.; Feng, M.; Peng, X.; Liao, X. Green Synthesis of Fluorescent N,S-Carbon Dots from Bamboo Leaf and the Interaction with Nitrophenol Compounds. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2020, 239, 118462. [Google Scholar] [CrossRef]
  72. Kalanidhi, K.; Nagaraaj, P. Facile and Green Synthesis of Fluorescent N-Doped Carbon Dots from Betel Leaves for Sensitive Detection of Picric Acid and Iron Ion. J. Photochem. Photobiol. A Chem. 2021, 418, 113369. [Google Scholar] [CrossRef]
  73. Ghosh, S.; Gul, A.R.; Park, C.Y.; Xu, P.; Baek, S.H.; Bhamore, J.R.; Kim, M.W.; Lee, M.; Kailasa, S.K.; Park, T.J. Green Synthesis of Carbon Dots from Calotropis procera Leaves for Trace Level Identification of Isoprothiolane. Microchem. J. 2021, 167, 106272. [Google Scholar] [CrossRef]
  74. Zaib, M.; Akhtar, A.; Maqsood, F.; Shahzadi, T. Green Synthesis of Carbon Dots and Their Application as Photocatalyst in Dye Degradation Studies. Arab. J. Sci. Eng. 2021, 46, 437–446. [Google Scholar] [CrossRef]
  75. Chauhan, P.; Chaudhary, S.; Kumar, R. Biogenic Approach for Fabricating Biocompatible Carbon Dots and Their Application in Colorimetric and Fluorometric Sensing of Lead Ion. J. Clean. Prod. 2021, 279, 123639. [Google Scholar] [CrossRef]
  76. Wang, C.; Xu, J.; Zhang, R.; Zhao, W. Facile and Low-Energy-Consumption Synthesis of Dual-Functional Carbon Dots from Cornus Walteri Leaves for Detection of p-Nitrophenol and Photocatalytic Degradation of Dyes. Colloids Surf. A Physicochem. Eng. Asp. 2022, 640, 128351. [Google Scholar] [CrossRef]
  77. Ge, G.; Li, L.; Chen, M.; Wu, X.; Yang, Y.; Wang, D.; Zuo, S.; Zeng, Z.; Xiong, W.; Guo, C. Green Synthesis of Nitrogen–Doped Carbon Dots from Fresh Tea Leaves for Selective Fe3+ Ions Detection and Cellular Imaging. Nanomaterials 2022, 12, 986. [Google Scholar] [CrossRef]
  78. Krishnaiah, P.; Atchudan, R.; Perumal, S.; Salama, E.S.; Lee, Y.R.; Jeon, B.H. Utilization of Waste Biomass of Poa Pratensis for Green Synthesis of N-Doped Carbon Dots and Its Application in Detection of Mn2+ and Fe3+. Chemosphere 2022, 286, 131764. [Google Scholar] [CrossRef]
  79. Tan, X.W.; Romainor, A.N.B.; Chin, S.F.; Ng, S.M. Carbon Dots Production via Pyrolysis of Sago Waste as Potential Probe for Metal Ions Sensing. J. Anal. Appl. Pyrolysis 2014, 105, 157–165. [Google Scholar] [CrossRef]
  80. Ang, W.L.; Boon Mee, C.A.L.; Sambudi, N.S.; Mohammad, A.W.; Leo, C.P.; Mahmoudi, E.; Ba-Abbad, M.; Benamor, A. Microwave-Assisted Conversion of Palm Kernel Shell Biomass Waste to Photoluminescent Carbon Dots. Sci. Rep. 2020, 10, 21199. [Google Scholar] [CrossRef]
  81. Monday, Y.N.; Abdullah, J.; Yusof, N.A.; Rashid, S.A.; Shueb, R.H. Facile Hydrothermal and Solvothermal Synthesis and Characterization of Nitrogen-Doped Carbon Dots from Palm Kernel Shell Precursor. Appl. Sci. 2021, 11, 1630. [Google Scholar] [CrossRef]
  82. Liu, S.; Liu, Z.; Li, Q.; Xia, H.; Yang, W.; Wang, R.; Li, Y.; Zhao, H.; Tian, B. Facile Synthesis of Carbon Dots from Wheat Straw for Colorimetric and Fluorescent Detection of Fluoride and Cellular Imaging. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2021, 246, 118964. [Google Scholar] [CrossRef] [PubMed]
  83. Ramanan, V.; Thiyagarajan, S.K.; Raji, K.; Suresh, R.; Sekar, R.; Ramamurthy, P. Outright Green Synthesis of Fluorescent Carbon Dots from Eutrophic Algal Blooms for in Vitro Imaging. ACS Sustain. Chem. Eng. 2016, 4, 4724–4731. [Google Scholar] [CrossRef]
  84. Moonrinta, S.; Jamnongsong, S.; Sampattavanich, S.; Kladsomboon, S.; Sajomsang, W.; Paoprasert, P. Synthesis of Biocompatible Carbon Dots from Yogurt and Gas Vapor Sensing. IOP Conf. Ser. Mater. Sci. Eng. 2018, 378, 012005. [Google Scholar] [CrossRef]
  85. Pacquiao, M.R.; de Luna, M.D.G.; Thongsai, N.; Kladsomboon, S.; Paoprasert, P. Highly Fluorescent Carbon Dots from Enokitake Mushroom as Multi-Faceted Optical Nanomaterials for Cr6+ and VOC Detection and Imaging Applications. Appl. Surf. Sci. 2018, 453, 192–203. [Google Scholar] [CrossRef]
  86. Plácido, J.; Bustamante-López, S.; Meissner, K.E.; Kelly, D.E.; Kelly, S.L. Microalgae Biochar-Derived Carbon Dots and Their Application in Heavy Metal Sensing in Aqueous Systems. Sci. Total Environ. 2019, 656, 531–539. [Google Scholar] [CrossRef] [Green Version]
  87. Zulfajri, M.; Liu, K.C.; Pu, Y.H.; Rasool, A.; Dayalan, S.; Huang, G.G. Utilization of Carbon Dots Derived from Volvariella Volvacea Mushroom for a Highly Sensitive Detection of Fe3+ and Pb2+ Ions in Aqueous Solutions. Chemosensors 2020, 8, 47. [Google Scholar] [CrossRef]
  88. Chauhan, P.; Saini, J.; Chaudhary, S.; Bhasin, K.K. Sustainable Synthesis of Carbon Dots from Agarose Waste and Prospective Application in Sensing of L-Aspartic Acid. Mater. Res. Bull. 2021, 134, 111113. [Google Scholar] [CrossRef]
  89. Shen, C.; Dong, C.; Cheng, L.; Shi, X.; Bi, H. Fluorescent Carbon Dots from Shewanella oneidensis MR–1 for Hg2+ and Tetracycline Detection and Selective Fluorescence Imaging of Gram–Positive Bacteria. J. Environ. Chem. Eng. 2022, 10, 107020. [Google Scholar] [CrossRef]
  90. Hu, Y.; Yang, J.; Tian, J.; Jia, L.; Yu, J.S. Waste Frying Oil as a Precursor for One-Step Synthesis of Sulfur-Doped Carbon Dots with PH-Sensitive Photoluminescence. Carbon 2014, 77, 775–782. [Google Scholar] [CrossRef]
  91. Zhang, C.; Xiao, Y.; Ma, Y.; Li, B.; Liu, Z.; Lu, C.; Liu, X.; Wei, Y.; Zhu, Z.; Zhang, Y. Algae Biomass as a Precursor for Synthesis of Nitrogen-and Sulfur-Co-Doped Carbon Dots: A Better Probe in Arabidopsis Guard Cells and Root Tissues. J. Photochem. Photobiol. B Biol. 2017, 174, 315–322. [Google Scholar] [CrossRef] [PubMed]
  92. Ramanan, V.; Siddaiah, B.; Raji, K.; Ramamurthy, P. Green Synthesis of Multifunctionalized, Nitrogen-Doped, Highly Fluorescent Carbon Dots from Waste Expanded Polystyrene and Its Application in the Fluorimetric Detection of Au3+ Ions in Aqueous Media. ACS Sustain. Chem. Eng. 2018, 6, 1627–1638. [Google Scholar] [CrossRef]
  93. Konwar, A.; Baruah, U.; Deka, M.J.; Hussain, A.A.; Haque, S.R.; Pal, A.R.; Chowdhury, D. Tea-Carbon Dots-Reduced Graphene Oxide: An Efficient Conducting Coating Material for Fabrication of an E-Textile. ACS Sustain. Chem. Eng. 2017, 5, 11645–11651. [Google Scholar] [CrossRef]
  94. Su, R.; Wang, D.; Liu, M.; Yan, J.; Wang, J.X.; Zhan, Q.; Pu, Y.; Foster, N.R.; Chen, J.F. Subgram-Scale Synthesis of Biomass Waste-Derived Fluorescent Carbon Dots in Subcritical Water for Bioimaging, Sensing, and Solid-State Patterning. ACS Omega 2018, 3, 13211–13218. [Google Scholar] [CrossRef]
  95. Xue, H.; Yan, Y.; Hou, Y.; Li, G.; Hao, C. Novel Carbon Quantum Dots for Fluorescent Detection of Phenol and Insights into the Mechanism. New J. Chem. 2018, 42, 11485–11492. [Google Scholar] [CrossRef]
  96. Lin, B.; Yan, Y.; Guo, M.; Cao, Y.; Yu, Y.; Zhang, T.; Huang, Y.; Wu, D. Modification-Free Carbon Dots as Turn-on Fluorescence Probe for Detection of Organophosphorus Pesticides. Food Chem. 2018, 245, 1176–1182. [Google Scholar] [CrossRef]
  97. Kumari, A.; Kumar, A.; Sahu, S.K.; Kumar, S. Synthesis of Green Fluorescent Carbon Quantum Dots Using Waste Polyolefins Residue for Cu2+ Ion Sensing and Live Cell Imaging. Sens. Actuators B Chem. 2018, 254, 197–205. [Google Scholar] [CrossRef]
  98. Aji, M.P.; Wati, A.L.; Priyanto, A.; Karunawan, J.; Nuryadin, B.W.; Wibowo, E.; Marwoto, P. Sulhadi Polymer Carbon Dots from Plastics Waste Upcycling. Environ. Nanotechnol. Monit. Manag. 2018, 9, 136–140. [Google Scholar] [CrossRef]
  99. Devi, S.; Gupta, R.K.; Paul, A.K.; Kumar, V.; Sachdev, A.; Gopinath, P.; Tyagi, S. Ethylenediamine Mediated Luminescence Enhancement of Pollutant Derivatized Carbon Quantum Dots for Intracellular Trinitrotoluene Detection: Soot to Shine. RSC Adv. 2018, 8, 32684–32694. [Google Scholar] [CrossRef] [Green Version]
  100. Feng, X.; Zhang, Y. A Simple and Green Synthesis of Carbon Quantum Dots from Coke for White Light-Emitting Devices. RSC Adv. 2019, 9, 33789–33793. [Google Scholar] [CrossRef] [Green Version]
  101. Hu, Y.; Gao, Z.; Yang, J.; Chen, H.; Han, L. Environmentally Benign Conversion of Waste Polyethylene Terephthalate to Fluorescent Carbon Dots for “on-off-on” Sensing of Ferric and Pyrophosphate Ions. J. Colloid Interface Sci. 2019, 538, 481–488. [Google Scholar] [CrossRef] [PubMed]
  102. Gunjal, D.B.; Gurav, Y.M.; Gore, A.H.; Naik, V.M.; Waghmare, R.D.; Patil, C.S.; Sohn, D.; Anbhule, P.V.; Shejwal, R.V.; Kolekar, G.B. Nitrogen Doped Waste Tea Residue Derived Carbon Dots for Selective Quantification of Tetracycline in Urine and Pharmaceutical Samples and Yeast Cell Imaging Application. Opt. Mater. 2019, 98, 109484. [Google Scholar] [CrossRef]
  103. Gunjal, D.B.; Naik, V.M.; Waghmare, R.D.; Patil, C.S.; Shejwal, R.V.; Gore, A.H.; Kolekar, G.B. Sustainable Carbon Nanodots Synthesised from Kitchen Derived Waste Tea Residue for Highly Selective Fluorimetric Recognition of Free Chlorine in Acidic Water: A Waste Utilization Approach. J. Taiwan Inst. Chem. Eng. 2019, 95, 147–154. [Google Scholar] [CrossRef]
  104. Irmania, N.; Dehvari, K.; Gedda, G.; Tseng, P.J.; Chang, J.Y. Manganese-Doped Green Tea-Derived Carbon Quantum Dots as a Targeted Dual Imaging and Photodynamic Therapy Platform. J. Biomed. Mater. Res. Part B Appl. Biomater. 2020, 108, 1616–1625. [Google Scholar] [CrossRef] [PubMed]
  105. Zhu, J.; Zhu, F.; Yue, X.; Chen, P.; Sun, Y.; Zhang, L.; Mu, D.; Ke, F. Waste Utilization of Synthetic Carbon Quantum Dots Based on Tea and Peanut Shell. J. Nanomater. 2019, 2019, 7965756. [Google Scholar] [CrossRef] [Green Version]
  106. Venkatesan, S.; Mariadoss, A.J.; Kathiravan, A.; Muthupandian, A. Fuel Waste to Fluorescent Carbon Dots and Its Multifarious Applications. Sens. Actuators B Chem. 2019, 282, 972–983. [Google Scholar] [CrossRef]
  107. Park, S.J.; Park, J.Y.; Chung, J.W.; Yang, H.K.; Moon, B.K.; Yi, S.S. Color Tunable Carbon Quantum Dots from Wasted Paper by Different Solvents for Anti-Counterfeiting and Fluorescent Flexible Film. Chem. Eng. J. 2020, 383, 123200. [Google Scholar] [CrossRef]
  108. Chaudhary, S.; Kumari, M.; Chauhan, P.; Ram Chaudhary, G. Upcycling of Plastic Waste into Fluorescent Carbon Dots: An Environmentally Viable Transformation to Biocompatible C-Dots with Potential Prospective in Analytical Applications. Waste Manag. 2021, 120, 675–686. [Google Scholar] [CrossRef]
  109. Lauria, A.; Lizundia, E. Luminescent Carbon Dots Obtained from Polymeric Waste. J. Clean. Prod. 2020, 262, 121288. [Google Scholar] [CrossRef]
  110. Ma, J.; Zhang, L.; Chen, X.; Su, R.; Shi, Q.; Zhao, S.; Xu, Q.; Xu, C. Mass Production of Highly Fluorescent Full Color Carbon Dots from the Petroleum Coke. Chin. Chem. Lett. 2020, 32, 1532–1536. [Google Scholar] [CrossRef]
  111. Liang, Y.M.; Yang, H.; Zhou, B.; Chen, Y.; Yang, M.; Wei, K.S.; Yan, X.F.; Kang, C. Waste Tobacco Leaves Derived Carbon Dots for Tetracycline Detection: Improving Quantitative Accuracy with the Aid of Chemometric Model. Anal. Chim. Acta 2022, 1191, 339269. [Google Scholar] [CrossRef]
  112. Chen, B.B.; Liu, Z.X.; Zou, H.Y.; Huang, C.Z. Highly Selective Detection of 2,4,6-Trinitrophenol by Using Newly Developed Terbium-Doped Blue Carbon Dots. Analyst 2013, 141, 2676–2681. [Google Scholar] [CrossRef]
  113. Dong, Y.; Pang, H.; Yang, H.B.; Guo, C.; Shao, J.; Chi, Y.; Li, C.M.; Yu, T. Carbon-Based Dots Co-Doped with Nitrogen and Sulfur for High Quantum Yield and Excitation-Independent Emission. Angew. Chem. Int. Ed. 2013, 52, 7800–7804. [Google Scholar] [CrossRef] [PubMed]
  114. Niu, Q.; Gao, K.; Lin, Z.; Wu, W. Amine-Capped Carbon Dots as a Nanosensor for Sensitive and Selective Detection of Picric Acid in Aqueous Solution via Electrostatic Interaction. Anal. Methods 2013, 5, 6228–6233. [Google Scholar] [CrossRef]
  115. Ahmed, G.H.G.; Laíño, R.B.; Calzón, J.A.G.; García, M.E.D. Highly Fluorescent Carbon Dots as Nanoprobes for Sensitive and Selective Determination of 4-Nitrophenol in Surface Waters. Microchim. Acta 2015, 182, 51–59. [Google Scholar] [CrossRef]
  116. Zhang, H.; Chen, Y.; Liang, M.; Xu, L.; Qi, S.; Chen, H.; Chen, X. Solid-Phase Synthesis of Highly Fluorescent Nitrogen-Doped Carbon Dots for Sensitive and Selective Probing Ferric Ions in Living Cells. Anal. Chem. 2014, 86, 9846–9852. [Google Scholar] [CrossRef] [PubMed]
  117. Zhao, L.; Geng, F.; Di, F.; Guo, L.H.; Wan, B.; Yang, Y.; Zhang, H.; Sun, G. Polyamine-Functionalized Carbon Nanodots: A Novel Chemiluminescence Probe for Selective Detection of Iron(Iii) Ions. RSC Adv. 2014, 4, 45768–45771. [Google Scholar] [CrossRef]
  118. Zhang, R.; Chen, W. Nitrogen-Doped Carbon Quantum Dots: Facile Synthesis and Application as a “Turn-off” Fluorescent Probe for Detection of Hg2+ Ions. Biosens. Bioelectron. 2013, 55, 83–90. [Google Scholar] [CrossRef]
  119. Zhang, P.; Xue, Z.; Luo, D.; Yu, W.; Guo, Z.; Wang, T. Dual-Peak Electrogenerated Chemiluminescence of Carbon Dots for Iron Ions Detection. Anal. Chem. 2014, 86, 5620–5623. [Google Scholar] [CrossRef]
  120. Zhang, L.; Han, Y.; Zhu, J.; Zhai, Y.; Dong, S. Simple and Sensitive Fluorescent and Electrochemical Trinitrotoluene Sensors Based on Aqueous Carbon Dots. Anal. Chem. 2015, 87, 2033–2036. [Google Scholar] [CrossRef]
  121. Hou, J.; Dong, J.; Zhu, H.; Teng, X.; Ai, S.; Mang, M. A Simple and Sensitive Fluorescent Sensor for Methyl Parathion Based on L-Tyrosine Methyl Ester Functionalized Carbon Dots. Biosens. Bioelectron. 2015, 68, 20–26. [Google Scholar] [CrossRef] [PubMed]
  122. Cui, X.; Zhu, L.; Wu, J.; Hou, Y.; Wang, P.; Wang, Z.; Yang, M. A Fluorescent Biosensor Based on Carbon Dots-Labeled Oligodeoxyribonucleotide and Graphene Oxide for Mercury (II) Detection. Biosens. Bioelectron. 2015, 63, 506–512. [Google Scholar] [CrossRef] [PubMed]
  123. Wang, B.; Song, A.; Feng, L.; Ruan, H.; Li, H.; Dong, S.; Hao, J. Tunable Amphiphilicity and Multifunctional Applications of Ionic-Liquid-Modified Carbon Quantum Dots. ACS Appl. Mater. Interfaces 2015, 7, 6919–6925. [Google Scholar] [CrossRef] [PubMed]
  124. Liu, L.; Feng, F.; Paau, M.C.; Hu, Q.; Liu, Y.; Chen, Z.; Choic, M.M.F. Carbon Dots Isolated from Chromatographic Fractions for Sensing Applications. RSC Adv. 2015, 5, 106838–106847. [Google Scholar] [CrossRef]
  125. Li, L.; Yu, B.; You, T. Nitrogen and Sulfur Co-Doped Carbon Dots for Highly Selective and Sensitive Detection of Hg (II) Ions. Biosens. Bioelectron. 2015, 74, 263–269. [Google Scholar] [CrossRef]
  126. Wang, Y.; Kim, S.H.; Feng, L. Highly Luminescent N, S- Co-Doped Carbon Dots and Their Direct Use as Mercury(II) Sensor. Anal. Chim. Acta 2015, 890, 134–142. [Google Scholar] [CrossRef]
  127. Liu, Y.; Liao, M.; He, X.; Liu, X.; Kou, X.; Xiao, D. One-Step Synthesis of Highly Luminescent Nitrogen-Doped Carbon Dots for Selective and Sensitive Detection of Mercury(Ii) Ions and Cellular Imaging. Anal. Sci. 2015, 31, 971–977. [Google Scholar] [CrossRef] [Green Version]
  128. Hou, Y.; Lu, Q.; Deng, J.; Li, H.; Zhang, Y. One-Pot Electrochemical Synthesis of Functionalized Fluorescent Carbon Dots and Their Selective Sensing for Mercury Ion. Anal. Chim. Acta 2015, 866, 69–74. [Google Scholar] [CrossRef]
  129. Wang, J.; Zhang, P.; Huang, C.; Liu, G.; Leung, K.C.F.; Wáng, Y.X.J. High Performance Photoluminescent Carbon Dots for in Vitro and in Vivo Bioimaging: Effect of Nitrogen Doping Ratios. Langmuir 2015, 31, 8063–8073. [Google Scholar] [CrossRef]
  130. Schneider, J.; Reckmeier, C.J.; Xiong, Y.; Von Seckendorff, M.; Susha, A.S.; Kasak, P.; Rogach, A.L. Molecular Fluorescence in Citric Acid-Based Carbon Dots. J. Phys. Chem. C 2016, 121, 2014–2022. [Google Scholar] [CrossRef]
  131. Xu, S.; Lu, H. Mesoporous Structured MIPs@CDs Fluorescence Sensor for Highly Sensitive Detection of TNT. Biosens. Bioelectron. 2016, 85, 950–956. [Google Scholar] [CrossRef] [PubMed]
  132. Dai, J.; Zambrana, M.; Fidalgo, M.; Dai, J.; Zambrana, M. Amino-Functionalized Fluorescent Carbon Dots for Chemical Sensing Jingjing. MRS Adv. 2016, 1, 1365–1370. [Google Scholar] [CrossRef]
  133. He, J.; Zhang, H.; Zou, J.; Liu, Y.; Zhuang, J.; Xiao, Y.; Lei, B. Carbon Dots-Based Fluorescent Probe for “off-on” Sensing of Hg(II) and I-. Biosens. Bioelectron. 2016, 79, 531–535. [Google Scholar] [CrossRef] [PubMed]
  134. Zhao, Z.; Zhang, J.; Wang, Y.; Chen, L.; Zhang, Y. Hydrothermal Synthesis of Fluorescent Nitrogen-Doped Carbon Quantum Dots from Ascorbic Acid and Valine for Selective Determination of Picric Acid in Water Samples. Int. J. Environ. Anal. Chem. 2016, 96, 1402–1413. [Google Scholar] [CrossRef]
  135. He, G.; Xu, M.; Shu, M.; Li, X.; Yang, Z.; Zhang, L.; Su, Y.; Hu, N.; Zhang, Y. Rapid Solid-Phase Microwave Synthesis of Highly Photoluminescent Nitrogen-Doped Carbon Dots for Fe3+ Detection and Cellular Bioimaging. Nanotechnology 2016, 27, 395706. [Google Scholar] [CrossRef] [Green Version]
  136. Fan, Y.Z.; Zhang, Y.; Li, N.; Liu, S.G.; Liu, T.; Li, N.B.; Luo, H.Q. A Facile Synthesis of Water-Soluble Carbon Dots as a Label-Free Fluorescent Probe for Rapid, Selective and Sensitive Detection of Picric Acid. Sens. Actuators B Chem. 2017, 240, 949–955. [Google Scholar] [CrossRef]
  137. Chang, M.M.F.; Ginjom, I.R.; Ng, S.M. Single-Shot ‘Turn-off’ Optical Probe for Rapid Detection of Paraoxon-Ethyl Pesticide on Vegetable Utilising Fluorescence Carbon Dots. Sens. Actuators B Chem. 2017, 242, 1050–1056. [Google Scholar] [CrossRef]
  138. Song, W.; Zhang, H.J.; Liu, Y.H.; Ren, C.L.; Chen, H.L. A New Fluorescence Probing Strategy for the Detection of Parathion-Methyl Based on N-Doped Carbon Dots and Methyl Parathion Hydrolase. Chin. Chem. Lett. 2017, 28, 1675–1680. [Google Scholar] [CrossRef]
  139. Tian, X.; Peng, H.; Li, Y.; Yang, C.; Zhou, Z.; Wang, Y. Highly Sensitive and Selective Paper Sensor Based on Carbon Quantum Dots for Visual Detection of TNT Residues in Groundwater. Sens. Actuators B Chem. 2017, 243, 1002–1009. [Google Scholar] [CrossRef]
  140. Liu, T.; Cui, Z.W.; Zhou, J.; Wang, Y.; Zou, Z.G. Synthesis of Pyridinic-Rich N, S Co-Doped Carbon Quantum Dots as Effective Enzyme Mimics. Nanoscale Res. Lett. 2017, 12, 375. [Google Scholar] [CrossRef] [Green Version]
  141. Liu, H.; Zhang, Y.; Liu, J.H.; Hou, P.; Zhou, J.; Huang, C.Z. Preparation of Nitrogen-Doped Carbon Dots with High Quantum Yield from: Bombyx Mori Silk for Fe(III) Ions Detection. RSC Adv. 2017, 7, 50584–50590. [Google Scholar] [CrossRef] [Green Version]
  142. Liu, Y.; Liu, Y.; Lee, J.; Lee, J.H.; Park, M.; Kim, H.Y. Rational Designed Strategy to Dispel Mutual Interference of Mercuric and Ferric Ions towards Robust, PH-Stable Fluorescent Carbon Nanodots. Analyst 2017, 142, 1149–1156. [Google Scholar] [CrossRef] [PubMed]
  143. Huang, H.; Weng, Y.; Zheng, L.; Yao, B.; Weng, W.; Lin, X. Nitrogen-Doped Carbon Quantum Dots as Fluorescent Probe for “off-on” Detection of Mercury Ions, L-Cysteine and Iodide Ions. J. Colloid Interface Sci. 2017, 506, 373–378. [Google Scholar] [CrossRef] [PubMed]
  144. Tian, Z.; Zhang, X.; Li, D.; Zhou, D.; Jing, P.; Shen, D.; Qu, S.; Zboril, R.; Rogach, A.L. Full-Color Inorganic Carbon Dot Phosphors for White-Light-Emitting Diodes. Adv. Opt. Mater. 2017, 5, 1700416. [Google Scholar] [CrossRef]
  145. Thongsai, N.; Nagae, Y.; Hirai, T.; Takahara, A.; Uchiyama, T.; Kamitani, K.; Paoprasert, P. Multifunctional Nitrogen-Doped Carbon Dots from Maleic Anhydride and Tetraethylenepentamine via Pyrolysis for Sensing, Adsorbance, and Imaging Applications. Sens. Actuators B Chem. 2017, 253, 1026–1033. [Google Scholar] [CrossRef]
  146. Huang, Y.; Cheng, Z. Simple and Green Synthesis of Boron-, Sulfur-, and Nitrogen-Co-Doped Carbon Dots as Fluorescent Probe for Selective and Sensitive Detection of Sunset Yellow. Nano 2017, 12, 1750123. [Google Scholar] [CrossRef]
  147. Du, Q.; Zheng, J.; Wang, J.; Yang, Y.; Liu, X. The Synthesis of Green Fluorescent Carbon Dots for Warm White LEDs. RSC Adv. 2018, 8, 19585–19595. [Google Scholar] [CrossRef] [Green Version]
  148. Chen, L.; Zhang, J. Carbon Dots with Tunable Photoluminescence Properties. In Proceedings of the 2nd International Conference of Theoretical and Applied Nanoscience and Nanotechnology, Niagara Falls, ON, Canada, 10–12 June 2018; pp. 1–6. [Google Scholar] [CrossRef]
  149. Yu, T.; Wang, H.; Guo, C.; Zhai, Y.; Yang, J.; Yuan, J. A Rapid Microwave Synthesis of Green-Emissive Carbon Dots with Solid-State Fluorescence and PH-Sensitive Properties. R. Soc. Open Sci. 2018, 5, 180245. [Google Scholar] [CrossRef] [Green Version]
  150. Ren, G.; Yu, L.; Zhu, B.; Tang, M.; Chai, F.; Wang, C.; Su, Z. Orange Emissive Carbon Dots for Colorimetric and Fluorescent Sensing of 2,4,6-Trinitrophenol by Fluorescence Conversion. RSC Adv. 2018, 8, 16095–16102. [Google Scholar] [CrossRef] [Green Version]
  151. Zhang, L.; Chang, J.; Su, R.; Huang, R.; Qi, W.; He, Z. Carbon Dots and Quantum Dots-Based Nanohybrid as a Ratiometric Fluorescent Probe for Fe3+ and Phytic Acid Sensing. Chem. Eng. Trans. 2018, 70, 2065–2070. [Google Scholar] [CrossRef]
  152. Ma, Y.; Mei, J.; Bai, J.; Chen, X.; Ren, L. Ratiometric Fluorescent Nanosensor Based on Carbon Dots for the Detection of Mercury Ion. Mater. Res. Express 2018, 5, 055605. [Google Scholar] [CrossRef]
  153. Tabaraki, R.; Abdi, O. Green and Simple Turn off/on Fluorescence Sensor for Mercury (II), Cysteine and Histidine. J. Mol. Liq. 2018, 251, 77–82. [Google Scholar] [CrossRef]
  154. Li, H.; Yan, X.; Lu, G.; Su, X. Carbon Dot-Based Bioplatform for Dual Colorimetric and Fluorometric Sensing of Organophosphate Pesticides. Sens. Actuators B Chem. 2018, 260, 563–570. [Google Scholar] [CrossRef]
  155. Yan, X.; Song, Y.; Zhu, C.; Li, H.; Du, D.; Su, X.; Lin, Y. MnO2 Nanosheet-Carbon Dots Sensing Platform for Sensitive Detection of Organophosphorus Pesticides. Anal. Chem. 2018, 90, 2618–2624. [Google Scholar] [CrossRef]
  156. Prathumsuwan, T.; Jamnongsong, S.; Sampattavanich, S.; Paoprasert, P. Preparation of Carbon Dots from Succinic Acid and Glycerol as Ferrous Ion and Hydrogen Peroxide Dual-Mode Sensors and for Cell Imaging. Opt. Mater. 2018, 86, 517–529. [Google Scholar] [CrossRef]
  157. Jiang, K.; Wang, Y.; Cai, C.; Lin, H. Conversion of Carbon Dots from Fluorescence to Ultralong Room-Temperature Phosphorescence by Heating for Security Applications. Adv. Mater. 2018, 30, e1800783. [Google Scholar] [CrossRef]
  158. Jiang, K.; Wang, Y.; Gao, X.; Cai, C.; Lin, H. Facile, Quick, and Gram-Scale Synthesis of Ultralong-Lifetime Room-Temperature-Phosphorescent Carbon Dots by Microwave Irradiation. Angew. Chem. Int. Ed. 2018, 57, 6216–6220. [Google Scholar] [CrossRef]
  159. Shamsipur, M.; Molaei, K.; Molaabasi, F.; Alipour, M.; Alizadeh, N.; Hosseinkhani, S.; Hosseini, M. Facile Preparation and Characterization of New Green Emitting Carbon Dots for Sensitive and Selective off/on Detection of Fe3+ Ion and Ascorbic Acid in Water and Urine Samples and Intracellular Imaging in Living Cells. Talanta 2018, 183, 122–130. [Google Scholar] [CrossRef]
  160. Ren, G.; Meng, Y.; Zhang, Q.; Tang, M.; Zhu, B.; Chai, F.; Wang, C.; Su, Z. Nitrogen-Doped Carbon Dots for the Detection of Mercury Ions in Living Cells and Visualization of Latent Fingerprints. New J. Chem. 2018, 42, 6824–6830. [Google Scholar] [CrossRef]
  161. Ludmerczki, R.; Mura, S.; Carbonaro, C.M.; Mandity, I.M.; Carraro, M.; Senes, N.; Garroni, S.; Granozzi, G.; Calvillo, L.; Marras, S.; et al. Carbon Dots from Citric Acid and Its Intermediates Formed by Thermal Decomposition. Chem.-A Eur. J. 2019, 25, 11963–11974. [Google Scholar] [CrossRef]
  162. Hu, G.; Lei, B.; Jiao, X.; Wu, S.; Zhang, X.; Zhuang, J.; Liu, X.; Hu, C.; Liu, Y. Synthesis of Modified Carbon Dots with Performance of Ultraviolet Absorption Used in Sunscreen. Opt. Express 2019, 27, 7629. [Google Scholar] [CrossRef] [PubMed]
  163. Zhao, Y.Y.; Qu, S.N.; Feng, X.Y.; Xu, J.C.; Yang, Y.; Su, S.C.; Wang, S.P.; Ng, K.W. Tailoring the Photoluminescence Excitation Dependence of the Carbon Dots via an Alkali Treatment. J. Phys. Chem. Lett. 2019, 10, 4596–4602. [Google Scholar] [CrossRef] [PubMed]
  164. Zhao, X.; Qi, T.; Yang, M.; Zhang, W.; Kong, C.; Hao, M.; Wang, Y.; Zhang, H.; Yang, B.; Yang, J.; et al. Synthesis of Dual Functional Procaine-Derived Carbon Dots for Bioimaging and Anticancer Therapy. Nanomedicine 2020, 15, 677–689. [Google Scholar] [CrossRef] [PubMed]
  165. Wang, M.; Gao, M.; Deng, L.; Kang, X.; Yang, L.; Quan, T.; Xia, Z.; Gao, D. Composite Material Based on Carbon Dots and Molecularly Imprinted Polymers: A Facile Probe for Fluorescent Detection of 4-Nitrophenol. Nano 2020, 15, 2050105. [Google Scholar] [CrossRef]
  166. Chan, K.K.; Yap, S.H.K.; Yong, K.T. Solid State Carbon Dots-Based Sensor Using Optical Microfiber for Ferric Ion Detection. In Proceedings of the 2019 IEEE International Conference on Sensors and Nanotechnology, Penang, Malaysia, 24–25 July 2019; pp. 1–4. [Google Scholar] [CrossRef]
  167. Yang, H.; He, L.; Pan, S.; Liu, H.; Hu, X. Nitrogen-Doped Fluorescent Carbon Dots for Highly Sensitive and Selective Detection of Tannic Acid. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2019, 210, 111–119. [Google Scholar] [CrossRef]
  168. Sahiner, N.; Suner, S.S.; Sahiner, M.; Silan, C. Nitrogen and Sulfur Doped Carbon Dots from Amino Acids for Potential Biomedical Applications. J. Fluoresc. 2019, 29, 1191–1200. [Google Scholar] [CrossRef]
  169. Pu, Z.F.; Wen, Q.L.; Yang, Y.J.; Cui, X.M.; Ling, J.; Liu, P.; Cao, Q.E. Fluorescent Carbon Quantum Dots Synthesized Using Phenylalanine and Citric Acid for Selective Detection of Fe3+ Ions. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2020, 229, 117944. [Google Scholar] [CrossRef]
  170. Wang, X.; Liu, Y.; Zhou, Q.; Sheng, X.; Sun, Y.; Zhou, B.; Zhao, J.; Guo, J. A Reliable and Facile Fluorescent Sensor from Carbon Dots for Sensing 2,4,6-Trinitrophenol Based on Inner Filter Effect. Sci. Total Environ. 2020, 720, 137680. [Google Scholar] [CrossRef]
  171. Gu, S.; Hsieh, C.T.; Yuan, C.Y.; Gandomi, Y.A.; Chang, J.K.; Fu, C.C.; Yang, J.W.; Juang, R.S. Fluorescence of Functionalized Graphene Quantum Dots Prepared from Infrared-Assisted Pyrolysis of Citric Acid and Urea. J. Lumin. 2020, 217, 116774. [Google Scholar] [CrossRef]
  172. Strauss, V.; Wang, H.; Delacroix, S.; Ledendecker, M.; Wessig, P. Carbon Nanodots Revised: The Thermal Citric Acid/Urea Reaction. Chem. Sci. 2020, 11, 8256–8266. [Google Scholar] [CrossRef]
  173. Li, Y.; Yang, Y.; Jiang, Y.; Han, S. Detection of Tannic Acid Exploiting Carbon Dots Enhanced Hydrogen Peroxide/Potassium Ferricyanide Chemiluminescence. Microchem. J. 2020, 157, 105113. [Google Scholar] [CrossRef]
  174. Sutanto, H.; Alkian, I.; Romanda, N.; Lewa, I.W.L.; Marhaendrajaya, I.; Triadyaksa, P. High Green-Emission Carbon Dots and Its Optical Properties: Microwave Power Effect. AIP Adv. 2020, 10, 055008. [Google Scholar] [CrossRef]
  175. Chahal, S.; Yousefi, N.; Tufenkji, N. Green Synthesis of High Quantum Yield Carbon Dots from Phenylalanine and Citric Acid: Role of Stoichiometry and Nitrogen Doping. ACS Sustain. Chem. Eng. 2020, 8, 5566–5575. [Google Scholar] [CrossRef]
  176. Arab, S.; Masoum, S.; Seyed Hosseini, E. Engineering the Synthesis of Luminescent Carbon Dots with Ultra High-Quantum Yield Using Experimental Design Approach in Order to Develop Sensing Applications. IEEE Sens. J. 2020, 20, 1705–1711. [Google Scholar] [CrossRef]
  177. He, X.; Li, Y.; Yang, C.; Lu, L.; Nie, Y.; Tian, X. Carbon Dots–MnO2 Nanocomposites for As( Iii ) Detection in Groundwater with High Sensitivity and Selectivity. Anal. Methods 2020, 12, 5572–5580. [Google Scholar] [CrossRef]
  178. Bhatt, M.; Bhatt, S.; Vyas, G.; Raval, I.H.; Haldar, S.; Paul, P. Water-Dispersible Fluorescent Carbon Dots as Bioimaging Agents and Probes for Hg2+ and Cu2+ Ions. ACS Appl. Nano Mater. 2020, 3, 7096–7104. [Google Scholar] [CrossRef]
  179. Sun, Y.; Zheng, S.; Liu, L.; Kong, Y.; Zhang, A.; Xu, K.; Han, C. The Cost-Effective Preparation of Green Fluorescent Carbon Dots for Bioimaging and Enhanced Intracellular Drug Delivery. Nanoscale Res. Lett. 2020, 15, 55. [Google Scholar] [CrossRef] [Green Version]
  180. Luo, M.; Wei, J.; Zhao, Y.; Sun, Y.; Liang, H.; Wang, S.; Li, P. Fluorescent and Visual Detection of Methyl-Paraoxon by Using Boron-and Nitrogen-Doped Carbon Dots. Microchem. J. 2020, 154, 104547. [Google Scholar] [CrossRef]
  181. Li, D.; Chen, C.; Guo, X.; Liu, C.; Yang, W. A Simple Electrochemiluminesecence Aptasenor Using a GCE/NCQDs/Aptamers for Detection of Pb. Environ. Technol. 2021, 43, 2270–2277. [Google Scholar] [CrossRef]
  182. Supchocksoonthorn, P.; Hanchaina, R.; Sinoy, M.C.A.; de Luna, M.D.G.; Kangsamaksin, T.; Paoprasert, P. Novel Solution- and Paper-Based Sensors Based on Label-Free Fluorescent Carbon Dots for the Selective Detection of Pyrimethanil. Appl. Surf. Sci. 2021, 564, 150372. [Google Scholar] [CrossRef]
  183. Wang, S.; Chen, H.; Xie, H.; Wei, L.; Xu, L.; Zhang, L.; Lan, W.; Zhou, C.; She, Y.; Fu, H. A Novel Thioctic Acid-Carbon Dots Fluorescence Sensor for the Detection of Hg2+ and Thiophanate Methyl via S-Hg Affinity. Food Chem. 2021, 346, 128923. [Google Scholar] [CrossRef] [PubMed]
  184. Liu, Q.; Dong, Z.; Hao, A.; Guo, X.; Dong, W. Synthesis of Highly Fluorescent Carbon Dots as a Dual-Excitation Rationmetric Fluorescent Probe for the Fast Detection of Chlorogenic Acid. Talanta 2021, 221, 121372. [Google Scholar] [CrossRef] [PubMed]
  185. Tang, X.; Yu, H.; Bui, B.; Wang, L.; Xing, C.; Wang, S.; Chen, M.; Hu, Z.; Chen, W. Nitrogen-Doped Fluorescence Carbon Dots as Multi-Mechanism Detection for Iodide and Curcumin in Biological and Food Samples. Bioact. Mater. 2021, 6, 1541–1554. [Google Scholar] [CrossRef] [PubMed]
  186. Vidya, T.; Anupama, M.; Muhammed, S.; Joseph, J.; Anappara, A.A. Multi-Functional Carbon Dots for Visual Detection of Picric Acid and White-Light Emission. Mater. Res. Bull. 2021, 138, 111223. [Google Scholar] [CrossRef]
  187. Mu, X.; Wu, M.; Zhang, B.; Liu, X.; Xu, S.; Huang, Y.; Wang, X.; Song, D.; Ma, P.; Sun, Y. A Sensitive “off-on” Carbon Dots-Ag Nanoparticles Fluorescent Probe for Cysteamine Detection via the Inner Filter Effect. Talanta 2021, 221, 121463. [Google Scholar] [CrossRef]
  188. Zairov, R.R.; Dovzhenko, A.P.; Sarkanich, K.A.; Nizameev, I.R.; Luzhetskiy, A.V.; Sudakova, S.N.; Podyachev, S.N.; Burilov, V.A.; Vatsouro, I.M.; Vomiero, A.; et al. Single Excited Dual Band Luminescent Hybrid Carbon Dots-Terbium Chelate Nanothermometer. Nanomaterials 2021, 11, 3080. [Google Scholar] [CrossRef]
  189. Yi, Z.; Li, X.; Zhang, H.; Ji, X.; Sun, W.; Yu, Y.; Liu, Y.; Huang, J.; Sarshar, Z.; Sain, M. High Quantum Yield Photoluminescent N-Doped Carbon Dots for Switch Sensing and Imaging. Talanta 2021, 222, 121663. [Google Scholar] [CrossRef]
  190. Ge, M.; Huang, X.; Ni, J.; Han, Y.; Zhang, C.; Li, S.; Cao, J.; Li, J.; Chen, Z.; Han, S. One-Step Synthesis of Self-Quenching-Resistant Biomass-Based Solid-State Fluorescent Carbon Dots with High Yield for White Lighting Emitting Diodes. Dye. Pigment. 2021, 185, 108953. [Google Scholar] [CrossRef]
  191. Ma, W.; Wang, B.; Yang, Y.; Li, J. Photoluminescent Chiral Carbon Dots Derived from Glutamine. Chin. Chem. Lett. 2021, 32, 3916–3920. [Google Scholar] [CrossRef]
  192. Xu, X.; Mo, L.; Li, W.; Li, Y.; Lei, B.; Zhang, X.; Zhuang, J.; Hu, C.; Liu, Y. Red, Green and Blue Aggregation-Induced Emissive Carbon Dots. Chin. Chem. Lett. 2021, 32, 3927–3930. [Google Scholar] [CrossRef]
  193. Ji, Y.; Zou, X.; Wang, W.; Wang, T.; Zhang, S.; Gong, Z. Co-Doped S, N-Carbon Dots and Its Fluorescent Film Sensors for Rapid Detection of Cr (VI) and Ascorbic Acid. Microchem. J. 2021, 167, 106284. [Google Scholar] [CrossRef]
  194. Liu, Y.; Chen, P.; Yu, K.; Wu, Z.; Wang, N.; Yu, X. Nitrogen and Sulfur Co-Doped Carbon Dots: Facile Synthesis and Multifunctional Applications for PH Sensing, Temperature Sensing and RNA-Selective Imaging. Microchem. J. 2021, 168, 106248. [Google Scholar] [CrossRef]
  195. Yang, Z.; Liu, Y.; Lu, C.; Yue, G.; Wang, Y.; Rao, H.; Zhang, W.; Lu, Z.; Wang, X. One-Pot Synthesis of CeO2-Carbon Dots with Enhanced Peroxidase-like Activity and Carbon Dots for Ratiometric Fluorescence Detection of H2O2 and Cholesterol. J. Alloys Compd. 2021, 862, 158323. [Google Scholar] [CrossRef]
  196. Şen, F.B.; Beğiç, N.; Bener, M.; Apak, R. Fluorescence Turn-off Sensing of TNT by Polyethylenimine Capped Carbon Quantum Dots. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2022, 271, 120884. [Google Scholar] [CrossRef] [PubMed]
  197. Huo, X.; Shen, H.; Xu, Y.; Shao, J.; Liu, R.; Zhang, Z. Fluorescence Properties of Carbon Dots Synthesized by Different Solvents for PH Detector. Opt. Mater. 2022, 123, 111889. [Google Scholar] [CrossRef]
  198. Wang, Q.; Zheng, H.; Long, Y.; Zhang, L.; Gao, M.; Bai, W. Microwave-Hydrothermal Synthesis of Fluorescent Carbon Dots from Graphite Oxide. Carbon 2011, 49, 3134–3140. [Google Scholar] [CrossRef]
  199. Liu, R.; Li, H.; Kong, W.; Liu, J.; Liu, Y.; Tong, C.; Zhang, X.; Kang, Z. Ultra-Sensitive and Selective Hg2+ Detection Based on Fluorescent Carbon Dots. Mater. Res. Bull. 2013, 48, 2529–2534. [Google Scholar] [CrossRef]
  200. Shi, J.; Lu, C.; Yan, D.; Ma, L. High Selectivity Sensing of Cobalt in HepG2 Cells Based on Necklace Model Microenvironment-Modulated Carbon Dot-Improved Chemiluminescence in Fenton-like System. Biosens. Bioelectron. 2013, 45, 58–64. [Google Scholar] [CrossRef]
  201. Xu, Y.; Wu, M.; Feng, X.Z.; Yin, X.B.; He, X.W.; Zhang, Y.K. Reduced Carbon Dots versus Oxidized Carbon Dots: Photo- and Electrochemiluminescence Investigations for Selected Applications. Chem.-A Eur. J. 2013, 19, 6282–6288. [Google Scholar] [CrossRef]
  202. Dong, X.; Su, Y.; Geng, H.; Li, Z.; Yang, C.; Li, X.; Zhang, Y. Fast One-Step Synthesis of N-Doped Carbon Dots by Pyrolyzing Ethanolamine. J. Mater. Chem. C 2014, 2, 7477–7481. [Google Scholar] [CrossRef]
  203. Huang, L.; Yang, S.; Chen, L.; Chen, S. Hydrothermal Synthesis of Fluorescent Carbon Dots towards Ion Response and Silk Screen Patterns. Chem. Lett. 2015, 44, 1251–1253. [Google Scholar] [CrossRef]
  204. Ahmed, G.H.G.; Laíño, R.B.; Calzón, J.A.G.; García, M.E.D. Fluorescent Carbon Nanodots for Sensitive and Selective Detection of Tannic Acid in Wines. Talanta 2015, 132, 252–257. [Google Scholar] [CrossRef] [Green Version]
  205. Hu, Y.; Yang, J.; Jia, L.; Yu, J.S. Ethanol in Aqueous Hydrogen Peroxide Solution: Hydrothermal Synthesis of Highly Photoluminescent Carbon Dots as Multifunctional Nanosensors. Carbon 2015, 93, 999–1007. [Google Scholar] [CrossRef]
  206. Zou, W.S.; Ji, Y.J.; Wang, X.F.; Zhao, Q.C.; Zhang, J.; Shao, Q.; Liu, J.; Wang, F.; Wang, Y.Q. Insecticide as a Precursor to Prepare Highly Bright Carbon Dots for Patterns Printing and Bioimaging: A New Pathway for Making Poison Profitable. Chem. Eng. J. 2016, 294, 323–332. [Google Scholar] [CrossRef]
  207. Wu, Z.; Feng, M.; Chen, X.; Tang, X. N-Dots as a Photoluminescent Probe for the Rapid and Selective Detection of Hg2+ and Ag+ in Aqueous Solution. J. Mater. Chem. B 2016, 4, 2086–2089. [Google Scholar] [CrossRef]
  208. Campos, B.B.; Contreras-Cáceres, R.; Bandosz, T.J.; Jiménez-Jiménez, J.; Rodríguez-Castellón, E.; Esteves da Silva, J.C.G.; Algarra, M. Carbon Dots as Fluorescent Sensor for Detection of Explosive Nitrocompounds. Carbon 2016, 106, 171–178. [Google Scholar] [CrossRef]
  209. Li, H.; Sun, C.; Vijayaraghavan, R.; Zhou, F.; Zhang, X.; MacFarlane, D.R. Long Lifetime Photoluminescence in N, S Co-Doped Carbon Quantum Dots from an Ionic Liquid and Their Applications in Ultrasensitive Detection of Pesticides. Carbon 2016, 104, 33–39. [Google Scholar] [CrossRef]
  210. Liu, X.; Pang, J.; Xu, F.; Zhang, X. Simple Approach to Synthesize Amino-Functionalized Carbon Dots by Carbonization of Chitosan. Sci. Rep. 2016, 6, 31100. [Google Scholar] [CrossRef] [Green Version]
  211. Wu, X.; Song, Y.; Yan, X.; Zhu, C.; Ma, Y.; Du, D.; Lin, Y. Carbon Quantum Dots as Fluorescence Resonance Energy Transfer Sensors for Organophosphate Pesticides Determination. Biosens. Bioelectron. 2017, 94, 292–297. [Google Scholar] [CrossRef]
  212. Wu, P.; Li, W.; Wu, Q.; Liu, Y.; Liu, S. Hydrothermal Synthesis of Nitrogen-Doped Carbon Quantum Dots from Microcrystalline Cellulose for the Detection of Fe3+ Ions in an Acidic Environment. RSC Adv. 2017, 7, 44144–44153. [Google Scholar] [CrossRef] [Green Version]
  213. Mohapatra, S.; Bera, M.K.; Das, R.K. Rapid “Turn-on” Detection of Atrazine Using Highly Luminescent N-Doped Carbon Quantum Dot. Sens. Actuators B Chem. 2018, 263, 459–468. [Google Scholar] [CrossRef]
  214. Devi, S.; Gupta, R.K.; Paul, A.K.; Tyagi, S. Waste Carbon Paper Derivatized Carbon Quantum Dots/(3- Aminopropyl)Triethoxysilane Based Fluorescent Probe for Trinitrotoluene Detection. Mater. Res. Express 2019, 6, 1–12. [Google Scholar] [CrossRef]
  215. Feng, Y.; Tan, L.; Tang, Q.; Zhong, W.; Yang, X. Synthesis of Carbon Dots from PEG6000 and Papain for Fluorescent and Doxycycline Sensing. Nano 2018, 13, 1–9. [Google Scholar] [CrossRef]
  216. Siddique, A.B.; Pramanick, A.K.; Chatterjee, S.; Ray, M. Amorphous Carbon Dots and Their Remarkable Ability to Detect 2,4,6-Trinitrophenol. Sci. Rep. 2018, 8, 9770. [Google Scholar] [CrossRef]
  217. Zhao, D.; Liu, X.; Wei, C.; Qu, Y.; Xiao, X.; Cheng, H. One-Step Synthesis of Red-Emitting Carbon Dots: Via a Solvothermal Method and Its Application in the Detection of Methylene Blue. RSC Adv. 2019, 9, 29533–29540. [Google Scholar] [CrossRef] [Green Version]
  218. Fang, J.; Zhuo, S.; Zhu, C. Fluorescent Sensing Platform for the Detection of P-Nitrophenol Based on Cu-Doped Carbon Dots. Opt. Mater. 2019, 97, 109396. [Google Scholar] [CrossRef]
  219. Xu, Y.; Yan, L.; Yue, M.; Yu, Y. Green Hydrothermal Synthesis of Fluorescent Carbon Dots from Glucose and Taurine for Fe3+ and PH Sensing. Nano 2019, 14, 1950014. [Google Scholar] [CrossRef]
  220. Gogoi, J.; Chowdhury, D. Calcium-Modified Carbon Dots Derived from Polyethylene Glycol: Fluorescence-Based Detection of Trifluralin Herbicide. J. Mater. Sci. 2020, 55, 11597–11608. [Google Scholar] [CrossRef]
  221. Saravanan, A.; Maruthapandi, M.; Das, P.; Ganguly, S.; Margel, S.; Luong, J.H.T.; Gedanken, A. Applications of N-Doped Carbon Dots as Antimicrobial Agents, Antibiotic Carriers, and Selective Fluorescent Probes for Nitro Explosives. ACS Appl. Bio Mater. 2020, 3, 8023–8031. [Google Scholar] [CrossRef]
  222. Tammina, S.K.; Yang, Y. 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. 2020, 387, 112134. [Google Scholar] [CrossRef]
  223. Houdová, D.; Soto, J.; Castro, R.; Rodrigues, J.; Soledad Pino-González, M.; Petković, M.; Bandosz, T.J.; Algarra, M. Chemically Heterogeneous Carbon Dots Enhanced Cholesterol Detection by MALDI TOF Mass Spectrometry. J. Colloid Interface Sci. 2021, 591, 373–383. [Google Scholar] [CrossRef] [PubMed]
  224. Hu, Y.; Gao, Z.; Luo, J. Fluorescence Detection of Malachite Green in Fish Tissue Using Red Emissive Se,N,Cl-Doped Carbon Dots. Food Chem. 2021, 335, 127677. [Google Scholar] [CrossRef] [PubMed]
  225. Mohammadi, S.; Mohammadi, S.; Salimi, A. A 3D Hydrogel Based on Chitosan and Carbon Dots for Sensitive Fluorescence Detection of MicroRNA-21 in Breast Cancer Cells. Talanta 2021, 224, 121895. [Google Scholar] [CrossRef] [PubMed]
  226. Liu, Q.; Zhao, F.; Shi, B.; Changli, L. Mussel-Inspired Polydopamine-Encapsulated Carbon Dots with Dual Emission for Detection of 4-Nitrophenol and Fe3+. Luminescence 2021, 36, 431–442. [Google Scholar] [CrossRef]
  227. Bisauriya, R.; Antonaroli, S.; Ardini, M.; Angelucci, F.; Ricci, A.; Pizzoferrato, R. Tuning the Sensing Properties of N and S Co-Doped Carbon Dots for Colorimetric Detection of Copper and Cobalt in Water. Sensors 2022, 22, 2487. [Google Scholar] [CrossRef]
  228. Ezati, P.; Rhim, J.W.; Molaei, R.; Priyadarshi, R.; Roy, S.; Min, S.; Kim, Y.H.; Lee, S.G.; Han, S. Preparation and Characterization of B, S, and N-Doped Glucose Carbon Dots: Antibacterial, Antifungal, and Antioxidant Activity. Sustain. Mater. Technol. 2022, 32, e00397. [Google Scholar] [CrossRef]
  229. Kong, B.; Yang, T.; Cheng, F.; Qian, Y.; Li, C.; Zhan, L.; Li, Y.; Zou, H.; Huang, C. Carbon Dots as Nanocatalytic Medicine for Anti-Inflammation Therapy. J. Colloid Interface Sci. 2022, 611, 545–553. [Google Scholar] [CrossRef]
  230. Zhu, X.; Han, L.; Liu, H.; Sun, B. A Smartphone-Based Ratiometric Fluorescent Sensing System for on-Site Detection of Pyrethroids by Using Blue-Green Dual-Emission Carbon Dots. Food Chem. 2022, 379, 132154. [Google Scholar] [CrossRef]
  231. Fen, Y.W.; Yunus, W.M.M.; Yusof, N.A. Surface Plasmon Resonance Optical Sensor for Detection of Pb2+ Based on Immobilized P-Tert-Butylcalix[4]Arene-Tetrakis in Chitosan Thin Film as an Active Layer. Sens. Actuators B Chem. 2012, 171, 287–293. [Google Scholar] [CrossRef]
  232. Daniyal, W.M.E.M.M.; Fen, Y.W.; Abdullah, J.; Sadrolhosseini, A.R.; Saleviter, S.; Omar, N.A.S. Exploration of Surface Plasmon Resonance for Sensing Copper Ion Based on Nanocrystalline Cellulose-Modified Thin Film. Opt. Express 2018, 26, 34880. [Google Scholar] [CrossRef]
  233. Anas, N.A.A.; Fen, Y.W.; Omar, N.A.S.; Ramdzan, N.S.M.; Daniyal, W.M.E.M.M.; Saleviter, S.; Zainudin, A.A. Optical Properties of Chitosan/Hydroxyl-Functionalized Graphene Quantum Dots Thin Film for Potential Optical Detection of Ferric (III) Ion. Opt. Laser Technol. 2019, 120, 105724. [Google Scholar] [CrossRef]
  234. Daniyal, W.M.E.M.M.; Fen, Y.W.; Anas, N.A.A.; Omar, N.A.S.; Ramdzan, N.S.M.; Nakajima, H.; Mahdi, M.A. Enhancing the Sensitivity of a Surface Plasmon Resonance-Based Optical Sensor for Zinc Ion Detection by the Modification of a Gold Thin Film. RSC Adv. 2019, 9, 41729–41736. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Ramdzan, N.S.M.; Fen, Y.W.; Omar, N.A.S.; Anas, N.A.A.; Liew, J.Y.C.; Daniyal, W.M.E.M.M.; Hashim, H.S. Detection of Mercury Ion Using Surface Plasmon Resonance Spectroscopy Based on Nanocrystalline Cellulose/Poly(3,4-Ethylenedioxythiophene) Thin Film. Measurement 2021, 182, 109728. [Google Scholar] [CrossRef]
  236. Omar, N.A.S.; Irmawati, R.; Fen, Y.W.; Abdullah, J.; Daud, N.F.M.; Daniyal, W.M.E.M.M.; Mahdi, M.A. A Sensing Approach for Manganese Ion Detection by Carbon Dots Nanocomposite Thin Film-based Surface Plasmon Resonance Sensor. Optik 2021, 243, 167435. [Google Scholar] [CrossRef]
  237. Mustapha Kamil, Y.; Al-Rekabi, S.H.; Abu Bakar, M.H.; Fen, Y.W.; Mohammed, H.A.; Halip, N.H.M.; Alresheedi, M.T.; Mahdi, M.A. Arsenic Detection Using Surface Plasmon Resonance Sensor with Hydrous Ferric Oxide Layer. Photonic Sens. 2021, 12, 220306. [Google Scholar] [CrossRef]
  238. Gonçalves, H.M.R.; Duarte, A.J.; Esteves da Silva, J.C.G. Optical Fiber Sensor for Hg(II) Based on Carbon Dots. Biosens. Bioelectron. 2010, 26, 1302–1306. [Google Scholar] [CrossRef] [Green Version]
  239. Sun, X.; Li, G.; Yin, Y.; Zhang, Y.; Li, H. Carbon Quantum Dot-Based Fluorescent Vesicles and Chiral Hydrogels with Biosurfactant and Biocompatible Small Molecule. Soft Matter 2018, 14, 6983–6993. [Google Scholar] [CrossRef]
  240. Wang, X.; Guo, H.; Wu, N.; Xu, M.; Zhang, L.; Yang, W. A Dual-Emission Fluorescence Sensor Constructed by Encapsulating Double Carbon Dots in Zeolite Imidazole Frameworks for Sensing Pb2+. Colloids Surf. A Physicochem. Eng. Asp. 2021, 615, 126218. [Google Scholar] [CrossRef]
  241. Jia, M.; Peng, L.; Yang, M.; Wei, H.; Zhang, M.; Wang, Y. Carbon Dots with Dual Emission: A Versatile Sensing Platform for Rapid Assay of Cr (VI). Carbon 2021, 182, 42–50. [Google Scholar] [CrossRef]
  242. Hashim, H.S.; Fen, Y.W.; Omar, N.A.S.; Fauzi, N.I.M.; Daniyal, W.M.E.M.M. Recent Advances of Priority Phenolic Compounds Detection Using Phenol Oxidases-Based Electrochemical and Optical Sensors. Meas. J. Int. Meas. Confed. 2021, 184, 109855. [Google Scholar] [CrossRef]
  243. Babar, D.G.; Garje, S.S. Nitrogen and Phosphorus Co-Doped Carbon Dots for Selective Detection of Nitro Explosives. ACS Omega 2020, 5, 2710–2717. [Google Scholar] [CrossRef] [PubMed]
  244. Fauzi, N.I.M.; Fen, Y.W.; Omar, N.A.S.; Hashim, H.S. Recent Advances on Detection of Insecticides Using Optical Sensors. Sensors 2021, 21, 3856. [Google Scholar] [CrossRef] [PubMed]
  245. Hou, J.; Dong, G.; Tian, Z.; Lu, J.; Wang, Q.; Ai, S.; Wang, M. A Sensitive Fluorescent Sensor for Selective Determination of Dichlorvos Based on the Recovered Fluorescence of Carbon Dots-Cu(II) System. Food Chem. 2016, 202, 81–87. [Google Scholar] [CrossRef] [PubMed]
  246. Mandal, P.; Sahoo, D.; Sarkar, P.; Chakraborty, K.; Das, S. Fluorescence Turn-on and Turn-off Sensing of Pesticides by Carbon Dot-Based Sensor. New J. Chem. 2019, 43, 12137–12151. [Google Scholar] [CrossRef]
  247. Bera, M.K.; Behera, L.; Mohapatra, S. A Fluorescence Turn-down-up Detection of Cu2+ and Pesticide Quinalphos Using Carbon Quantum Dot Integrated UiO-66-NH2. Colloids Surf. A Physicochem. Eng. Asp. 2021, 624, 126792. [Google Scholar] [CrossRef]
  248. Liu, Y.; Zhou, Q.; Wu, Y.; Li, S.; Sun, Y.; Sheng, X.; Zhan, Y.; Zhao, J.; Guo, J.; Zhou, B. Sensitive Detection of 2,4,6-Trinitrotoluene Utilizing Fluorescent Sensor from Carbon Dots and Reusable Magnetic Core-Shell Nanomaterial. Talanta 2021, 233, 122498. [Google Scholar] [CrossRef]
Figure 1. Schematic of hydrothermal reaction of CDs from citrus lemon juice and ethylenediamine.
Figure 1. Schematic of hydrothermal reaction of CDs from citrus lemon juice and ethylenediamine.
Nanomaterials 12 02365 g001
Figure 2. UV–Vis spectra of CDs from Calotropis procera leaves through hydrothermal method at 200 °C for 4 h. Reproduced with copyright permission of Elsevier [73].
Figure 2. UV–Vis spectra of CDs from Calotropis procera leaves through hydrothermal method at 200 °C for 4 h. Reproduced with copyright permission of Elsevier [73].
Nanomaterials 12 02365 g002
Figure 3. Scheme illustration of tunable PL of CDs with different degrees of oxidation.
Figure 3. Scheme illustration of tunable PL of CDs with different degrees of oxidation.
Nanomaterials 12 02365 g003
Figure 4. (a) Absorption spectra of three CD species and pure citrazinic acid. Adapted with permission from ref. [130]. Copyright 2017 American Chemical Society; (b) the salt-induced phase separation using acetone. Reproduced with copyright permission of Elsevier [137].
Figure 4. (a) Absorption spectra of three CD species and pure citrazinic acid. Adapted with permission from ref. [130]. Copyright 2017 American Chemical Society; (b) the salt-induced phase separation using acetone. Reproduced with copyright permission of Elsevier [137].
Nanomaterials 12 02365 g004
Figure 5. Proposed possible mechanisms for the formation CDs using 3–bromophenol. Reproduced with copyright permission of Elsevier [206].
Figure 5. Proposed possible mechanisms for the formation CDs using 3–bromophenol. Reproduced with copyright permission of Elsevier [206].
Nanomaterials 12 02365 g005
Figure 6. Schematic diagram of target detection and possible detection mechanism.
Figure 6. Schematic diagram of target detection and possible detection mechanism.
Nanomaterials 12 02365 g006
Figure 7. Surface functionalization of (a) GCE/NCQDs, (b) GCE/NCQDs/aptamer, and (c) GCE/NCQDs/aptamer/Pb2+.
Figure 7. Surface functionalization of (a) GCE/NCQDs, (b) GCE/NCQDs/aptamer, and (c) GCE/NCQDs/aptamer/Pb2+.
Nanomaterials 12 02365 g007
Figure 8. Schematic diagram of binding reactions of the prepared CD-functionalized core-shell nanospheres for detection of λ-cyhalothrin.
Figure 8. Schematic diagram of binding reactions of the prepared CD-functionalized core-shell nanospheres for detection of λ-cyhalothrin.
Nanomaterials 12 02365 g008
Figure 9. Quantum yields of the CDs based on (a) green precursors and (b) chemical precursors.
Figure 9. Quantum yields of the CDs based on (a) green precursors and (b) chemical precursors.
Nanomaterials 12 02365 g009
Table 1. Summary of the synthesis CDs from fruits.
Table 1. Summary of the synthesis CDs from fruits.
PrecursorTechniquePropertiesYearReference
Carbon SourcePassivation/SolventParticle SizeFluorescenceQuantum Yield
Orange juice-HydrothermalTEM- 2.5 nmλem- 441 nm
λex- 360 nm
26%2012[9]
Watermelon peels-CarbonizationTEM- 2.0 nmλem- 490–580 nm7.1%2012[10]
Sugar cane bagasseSodium hydroxide solutionHydrothermal carbonizationHRTEM- 1.8 nmλem- 475 nm
λex- 370 nm
12.3%2014[11]
Lemon peelsSulfuric acidHydrothermal carbonizationTEM- 1 to 3 nmλem- 441 nm
λex- 360 nm
14%2016[12]
Prunus avium extractAmmoniaHydrothermal carbonizationHRTEM- 7 nmλem- 411 nm
λex- 310 nm
13%2016[13]
CornstalkDistilled waterHydrothermalTEM- 5.2 nmλem- 500 nm
λex- 420 nm
7.6%2017[14]
Corn bractAnhydrous ethanolSolvothermalTEM- 1.8–3.4 nmλem- 470 nm and 678 nm
λex- 406 nm
6.9%2017[15]
Dried lemon peelDeionized waterHydrothermalTEM- 9.5 nmλem- 505 nm
λex- 425 nm
11%2017[16]
Pulp-free lemon juiceEthanolSolvothermalAFM- 1.5 nm
TEM- 4.6 nm
λem- 631 nm
λex- 540 nm
28%2017[17]
Citrus lemon peels-CarbonizationTEM- 4.5 nmλem- 435 nm
λex- 330 nm
16.8%2017[18]
Citrus sinensis peelsTEM- 6.5 nmλem- 455 nm
λex- 365 nm
15.5%
Lemon juice-Thermal decomposition-λem- 400 nm
λex- 320 nm
7%2018[19]
Lemon juicePoly(ethylenimine)CarbonizationTEM- 5.7 nmλem- 540 nm
λex- 420 nm
-2018[20]
Citrus lemon juice-HydrothermalTEM- 5.8 nmλem- 450 nm
λex- 360 nm
10.20%2018[21]
Watermelon juiceEthanolHydrothermalTEM- 3–7 nmλem- 439 nm
λex- 355 nm
10.6%2018[22]
Jackfruit juiceEthanol and distilled waterHydrothermalHRTEM- <2.5 nmλem- 485 nm
λex- 395 nm
14.6%2018[23]
LemonEthylenediamineHydrothermalTEM- 20 nm-20%2018[24]
Grapefruit
Durian juiceWater and ethanolCarbonization---2018[25]
Acerola fruitWaterHydrothermal-λem- 504 nm
λex- 360 nm
-2019[26]
Bitter orange juice-HydrothermalAFM- 2–4 nm
DLS- 1–2 nm
λem- 390 nm
λex- 325 nm
19.9%2019[27]
Citrus lemon juiceEthylenediamineHydrothermalHRTEM- 3 nmλem- 452 nm
λex- 360 nm
31%2019[28]
Lemon and onion juicesAmmonium hydroxide solutionMicrowave assisted carbonizationTEM- 6.15 nmλem- 425 nm
λex- 340 nm
23.6%2019[29]
Lemon juice-HydrothermalHRTEM- 3–15 nmλem- 524 nm
λex- 420 nm
21.37%2019[30]
Durian shellTris base and deionized waterHydrothermal carbonizationTEM- 6.5 nmλem- 414 nm
λex- 340 nm
12.93%2019[31]
LemonHydroxylamineHydrothermalHRTEM- 2 nmλem- 430–470 nm
λex- 360 nm
5%2020[32]
PomegranateSodium hydroxide and polyethylene glycolMicrowaveHRTEM- 1 to 5 nmλem- 532 nm-2020[33]
Watermelon peelsλem- 515 nm
Rosa roxburghii fruitsWaterHydrothermalTEM- 2.5 nmλem- 450 nm
λex- 360 nm
24.8%2020[34]
Citrus fruit peelsDeionized waterSand bathTEM- 4.6 nmλem- 510 nm
λex- 420 nm
-2021[35]
Banana peelDeionized waterHydrothermalTEM- 5 nmλex- 355 nm20%2021[36]
Elaeagnus angustifoliaUltrapure waterHydrothermalTEM- <10 nmλem- 410 nm
λex- 330 nm
16.8%2021[37]
Kiwi (Actinidia deliciosa) fruit peels-Hydrothermal carbonizationTEM- 5.6 nmλem- 432 nm
λex- 360 nm
14%2021[1]
Ammonium hydroxideTEM- 5.1 nm19%
Canon ball fruitDistilled waterHydrothermalTEM- <15 nmλem- ~500 nm
λex- 380 nm
7.24%2022[38]
Indian Bael patra
-hard shell
-pulp
-pulp and gum
-Hydrothermal carbonizationTEM- 3 nm
TEM- 6 nm
TEM- 8 nm
59.39%
59.07%
55.25%
2022[39]
Morus nigra (black mulberry)Deionized waterHydrothermalTEM- 4.5 nmλem- 427 nm
λex- 360 nm
24%2022[40]
JatrophaDistilled waterHydrothermalTEM- 3.2 nmλem- 462 nm
λex- 370 nm
13.7%2022[41]
Table 2. Summary of the synthesis of CDs from vegetables.
Table 2. Summary of the synthesis of CDs from vegetables.
PrecursorTechniquePropertiesYearReference
Carbon SourcePassivation/SolventParticle SizeFluorescenceQuantum Yield
Celery leavesGlutathione and double distilled waterHydrothermalTEM- 2.08 nmλem- 415 nm
λex- 340 nm
53%2013[42]
Sweet pepperWaterCarbonizationTEM- 4.6 nmλem- 450 nm
λex- 360 nm
19.3%2013[43]
Lemon grassWaterHydrothermal-λem- 440 nm
λex- 320 nm
23.3%2016[44]
Tomato juice-HydrothermalHRTEM- 3 nm
DLS- 3 nm
λem- 440 nm
λex- 367 nm
13.9%2016[45]
Carrot juice-HydrothermalTEM- 5.5 nmλem- 442–565 nm
λex- 360–520 nm
5.16%2017[46]
Rose-heart radishUltrapure waterHydrothermalTEM- 3.6 nmλem- 420 nm
λex- 330 nm
13.6%2017[47]
TurmericEthylenediamineHydrothermalTEM- 20 nm-20%2018[24]
CinnamonUltrapure waterHydrothermalTEM- 3.4 nmλem- 465 nm
λex- 370 nm
35.7%2018[48]
Red chiliTEM- 3.1 nmλem- 477 nm
λex- 380 nm
26.8%
TurmericTEM- 4.3 nmλem- 460 nm
λex- 370 nm
38.3%
Black pepperTEM- 3.5 nmλem- 489 nm
λex- 390 nm
43.6%
HongcaitaiUltrapure waterHydrothermalTEM- 1.9 nmλem- 410 nm
λex- 330 nm
21.0%2018[49]
Cauliflower-HydrothermalAFM- 4 nm
DLS- 1.54 nm
λem- 380 nm
λex- 325 nm
43%2019[50]
KelpEthylenediamineMicrowave irradiationTEM- 3.7 nmλem- 450 nm
λex- 370 nm
23.5%2019[51]
TomatoSulfuric acidChemical oxidationHRTEM- 5–10 nmλem- 450
λex- 360 nm
12.70%2019[52]
Phosphoric acidλem- 520
λex- 420 nm
4.21%
Phosphoric acidλem- 560 nm
λex- 460 nm
2.76%
Crown daisy leaf wasteUltrapure water and ureaHydrothermalTEM- 5–10 nmλem- 380 nm
λex- 300 nm
-2019[53]
CabbageAnhydrous ethanolSolvothermalTEM- 3.4 nmλem- 500 nm and 678 nm
λex- 410 nm
12.4%2020[54]
Cherry tomatoes-HydrothermalTEM- 7 nmλem- 430 nm
λex- 340 nm
9.7%2020[55]
TomatoHydroxylamineHydrothermalHRTEM- 3 nmλem- 430–470 nm
λex- 360 nm
3.38%2020[32]
Scallion leavesWaterHydrothermalTEM- 3.5 nmλem- 418 nm
λex- 320 nm
3.2%2020[56]
Tomato-HydrothermalHRTEM- 9 nmλem- 430 nm
λex- 344 nm
1.24%2021[57]
Red beetWaterHydrothermalTEM- 4.66 nmλem- 438 nm
λex- 350 nm
8.17%2022[58]
Table 3. Summary of the synthesis of CDs from flowers.
Table 3. Summary of the synthesis of CDs from flowers.
PrecursorTechniquePropertiesYearReference
Carbon SourcePassivation/SolventParticle SizeFluorescenceQuantum Yield
Selenicereus grandiflorus-BoilingHRTEM- 2.5 nmλem- 440 and 365 nm3.8%2019[59]
Water hyacinthPhosphoric acidCarbonizationSEM- ≤10 nm
DLS- ≤10 nm
TEM- 5.22 nm
λem- 370 nm
λex- 300 nm
17.02%2019[60]
Osmanthus fragransUltrapure waterHydrothermalTEM- 2.23 nmλem- 410 nm
λex- 340 nm
18.53%2019[61]
Rose flowers:
Blue
Red
Yellow
WaterHydrothermalTEM- 37 nm
TEM- 39 nm
TEM- 33 nm
λex- 335 nm
λex- 330 nm
λex- 340 nm
46%
44%
48%
2020[62]
EthanolTEM- 30 nm
TEM- 27 nm
TEM- 26 nm
λex- 420 nm
λex- 410 nm
λex- 425 nm
43%
46%
47%
Tagetes erectaDeionized waterSolvo(hydro)-thermal carbonizationFESEM- 3.41 nmλem- 495 nm
λex- 420 nm
63.7%2021[63]
Table 4. Summary of the synthesis of CDs from leaves, seeds, and stems.
Table 4. Summary of the synthesis of CDs from leaves, seeds, and stems.
PrecursorTechniquePropertiesYearReference
Carbon SourcePassivation/SolventParticle SizeFluorescenceQuantum Yield
Lotus root-MicrowaveTEM- 9.41 nmλem- 435 nm
λex- 360 nm
19.0%2016[64]
Ocimum sanctum leavesDistilled waterHydrothermalTEM- 2.23 nmλem- 410 nm
λex- 340 nm
9.3%2017[65]
Acacia concinna seedsMethanolMicrowave treatmentHRTEM- 2.5 nmλem- 468 nm
λex- 390 nm
10.20%2018[66]
Acetonitrile7.20%
Acetone7.85%
Bamboo leavesSodium hydroxide and sodium hypochloritePyrolysisAFM- 2 nmλem- 425–475 nm-2018[67]
Gingko leaves-PyrolysisTEM- 4.11 nmλem- 427 nm
λex- 360 nm
21.7%2018[68]
Gynostemma-CalcinationTEM- 2.5 nmλem- 400 nm
λex- 320 nm
5.7%2019[69]
Fennel seeds (Foeniculum vulgare)-PyrolysisTEM- 3.9 nmλem- 417 nm
λex- 240 nm
9.5%2019[70]
Bamboo leaves-CalcinationTEM- 11 nmλem- 419 nm
λex- 313 nm
5.18%2020[71]
Betel leavesAmmoniaHydrothermalHRTEM- less 10 nm
PSA-3.7 nm
λem- 402 nm
λex- 320 nm
4.21%2021[72]
Calotropis procera leavesDeionized waterHydrothermal carbonizationFETEM- 4.3 nmλem- 416 nm
λex- 340 nm
71.95%2021[73]
Elettaria cardamomum leavesDistilled waterUltrasonication-λem- 520 and 850 nm
λex- 514 nm
-2021[74]
Pearl millet seedsDouble distilled waterThermal treatmentHRTEM- 4–5 nmλem- 415 nm
λex- 250 nm
52%2021[75]
Cornus walteri leavesMaleic anhydride, hydrogen peroxide and waterHydrothermalTEM- 3.53 nmλem- 550 nm
λex- 420 nm
18.34%2022[76]
Tea leavesUrea and ultrapure waterHydrothermalTEM- 2.32 nmλem- 455 nm
λex- 360 nm
-2022[77]
Kentucky bluegrassEthylenediamineHydrothermalTEM- 9 nmλem- 370–470 nm
λex- 280–400 nm
7%2022[78]
Table 5. Summary of the synthesis of CDs from crop residues.
Table 5. Summary of the synthesis of CDs from crop residues.
PrecursorTechniquePropertiesYearReference
Carbon SourcePassivation/SolventParticle SizeFluorescenceQuantum Yield
Sago waste-Thermal pyrolysisSEM- 6–17 nmλem- 390 nm
λex- 315 nm
-2014[79]
Palm kernel shellDiethylene glycolMicrowave irradiationTEM- 6.6 to 7 nmλem- 438–459 nm
λex- 370 nm
44.0%2020[80]
Palm kernel shellUltrapure water and ethylenediamineHydrothermalTEM-2 nmλem- 430–450 nm
λex- 350–400 nm
13.7%2021[81]
Ethanol and L-phenylalanine8.6%
Wheat strawDeionized waterHydrothermalTEM- 2.1 nm
DLS- 5.7 nm
λem- 470 nm
λex- 380 nm
7.5%2021[82]
Table 6. Summary of the synthesis of CDs from fungi/bacteria species.
Table 6. Summary of the synthesis of CDs from fungi/bacteria species.
PrecursorTechniquePropertiesYearReference
Carbon SourcePassivation/SolventParticle SizeFluorescenceQuantum Yield
Algal bloomsPhosphoric acidMicrowaveTEM- 8.5 nmλem- 438 nm
λex- 360 nm
13%2016[83]
YogurtHydrochloric acidPyrolysisTEM- 3.5 nmλem- 420 nm
λex- 320 nm
2.4%2018[84]
Enokitake mushroomSulfuric acidHydrothermalTEM- 4 nmλem- 470 nm
λex- 360 nm
11%2018[85]
Sulfuric acid and tetraethylenepentamine39%
Microalgae biocharPotassium permanganateOxidizing agent and autoclaveAFM- 68 nmλex- 398 nm
λex- 280 nm
-2019[86]
MushroomUltrapure waterHydrothermalTEM- 5.8 nmλem- 440 nm
λex- 360 nm
11.5%2020[87]
Agarose waste-Thermal treatmentHRTEM- 2–10 nmλem- 420 nm
λex- 300 nm
62%2021[88]
Shewanella oneidensisLuria-BertaniHydrothermal-λem- 410 nm
λex- 320 nm
7%2022[89]
Table 7. Summary of the synthesis of CDs from waste products.
Table 7. Summary of the synthesis of CDs from waste products.
PrecursorTechniquePropertiesYearReference
Carbon SourcePassivation/SolventParticle SizeFluorescenceQuantum Yield
Waste frying oilSulfuric acidCarbonizationHRTEM- 2.6 nmλem- 378 nm
λex- 300 nm
3.66%2014[90]
Nannochloropsis biocrude oilSulfuric acidHydrothermal liquefactionTEM- 4 nmλem- 280–560 nm
λex- 400–550 nm
13.71%2017[91]
PolystyreneEthylenediamineSolvothermalTEM- 4 nmλem- 456 nm
λex- 380 nm
~20%2017[92]
Assam CTC teaAcetic acidCarbonizationTEM- <10 nmλex- 380–500 nm-2017[93]
Expired milkWaterSubcritical waterTEM- <5 nmλem- 440 nm
λex- 360 nm
8.64%2018[94]
Long flame coal powderDeionized waterOzone oxidationDLS- 4.2 nmλem- 530 nm
λex- 470 nm
8.4%2018[95]
Paper wasteSodium hydroxide and distilled waterHydrothermalTEM- 2–4 nmλem- 420 nm
λex- 320 nm
20%2018[96]
Waste polyolefinSulfuric acid and nitric acidUltrasonic-assisted chemical oxidationTEM- 2.5 nmλem- 540 nm
λex- 490 nm
4.84%2018[97]
Polypropylene plastic wasteEthanolHeating processTEM- <20 nmλem- 410–465 nm
λex- 365 nm
-2018[98]
Bike sootDeionized waterHydrothermalTEM- 4.2 nmλem- 396 nm
λex- 240 nm
~5.63%2018[99]
Nitric acidTEM- 5.6 nmλem-
λex-
3.25%
Phosphoric acid λem- 560 nm
λex- 460 nm
2.76%
Coke powderHydrogen peroxideChemical oxidationTEM- 6.5 nmλem- 410 nm
λex- 330 nm
9.2%2019[100]
Waste plastic bottlesHydrogen peroxideAir oxidation and hydrothermalTEM- 3–10 nmλem- 434 nm
λex- 340 nm
5.2%2019[101]
Waste tea powderNitric acidChemical oxidationTEM- 3.2 nmλem- 430 nm
λex- 310 nm
2.47%2019[102]
Waste tea powder-CarbonizationTEM- 5 nmλem- 415 nm
λex- 315 nm
4.76%2019[103]
Waste green tea powderDeionized water and manganese chlorideHydrothermalTEM- 5 nmλem- 410–440 nm
λex- 360 nm
12%2019[104]
Tieguanyin Tea leavesAcetic acidHydrothermalTEM- 7–9 nmλem- blue
λex- 325 nm
-2019[105]
Peanut shellλem- blue
λex- 335 nm
Kerosene sootNitric acidOxidative acid treatmentHRTEM- 5 nmλem- 510 nm
λex- 300–360 nm
~3%2019[106]
Paper wasteDeionized waterHydro-/solvothermalTEM- 2.6 nmλem- blue
λex- 360 nm
12%2020[107]
EthanolTEM- 4.0 nmλem- 435 nm (cyan)
λex- 360 nm
27%
2-propanolTEM- 4.4 nmλem- 435 nm (cyan)
λex- 360 nm
10%
Polybags-Hydrothermal carbonizationHRTEM- 5–10 nmλem- 420–425 nm
λex- 310 nm
62%2021[108]
Cupsλem- 420–425 nm
λex- 310 nm
65%
Bottlesλem- 420–425 nm
λex- 310 nm
64%
Polymeric waste-HydrothermalTEM- 3 nmλem- 400 nm
λex- 310 nm
-2021[109]
4,7,10-trioxa-1,13-tridecanediamineλem- 440 nm
λex- 365 nm
Heavy oil-Hydrothermal---2021[110]
Light deasphalted oil (LDAO)--~64%
Heavy deasphalted oil (HDAO)--23.5%
AsphaltTEM- 2.39 nmλem- 610 nm
λex- 475 nm
11.5%
TEM- 1.77 nmλem- 560 nm
λex- 420 nm
17.7%
TEM- 1.21 nmλem- 510 nm
λex- 410 nm
28.3%
TEM- 1.18 nmλem- 440 nm
λex- 350 nm
64%
Waste tobacco leavesEthylenediamine and ultrapure waterHydrothermalTEM- 6.30 nmλem- 430 nm
λex- 360 nm
13.7%2022[111]
Table 11. Summary of the developed optical sensors for phenol detection.
Table 11. Summary of the developed optical sensors for phenol detection.
Phenolic CompoundMaterialOptical SensorRange of DetectionLimit of DetectionLinear Correlation CoefficientYearReference
2-4-6-TrinitrophenolAmine-capped CDsFluorescent0–50 × 10−5 M0.9996 μM-2013[114]
2-4-6-TrinitrophenolTb-CDsFluorescent500 nM–100 μM200 nM0.9912013[112]
4-NitrophenolCDsFluorescent0.1–50 µM28 nM-2014[115]
Tannic acidPEGA-CDsFluorescent0.05–0.6 μM0.01 μM-2015[204]
2-4-6-TrinitrophenolN-CQDsFluorescent0.27–34.1 µM50 nM0.9922016[134]
2-4-6-TrinitrophenolCDsFluorescent0.1–26.5 µM51 nM0.9952017[136]
2-4-6-TrinitrophenolCDsFluorescent-0.127 µM-2018[150]
PhenolCDsFluorescent0–50 µM0.076 µM0.9982018[95]
Tannic acidN-CDsFluorescent0.4–9.0 µM0.12 µM0.99902019[167]
4-NitrophenolCDs@MIPsFluorescent0–0.03594 mM35 nM-2019[165]
p-NitrophenolCu-doped carbon dotsFluorescent0.5–50 μM0.08 μM0.9982019[218]
2-4-6-TrinitrophenolCDsFluorescent0–305.54 μM0.023 μM-2020[170]
o-NitrophenolCQDsFluorescent0.08–40 µM15.2 nM0.9992020[34]
4-NitrophenolN,CDsFluorescent0.25–125 μM0.05 μM0.99192020[222]
2-4-6-TrinitrophenolN@CDsFluorescent1–75 μM2.45 μM0.9942020[221]
TrinitrophenolwsNP-CDsFluorescent100–300 μM23 μM0.98612020[243]
Tannic acidNitrogen-doped CDsChemiluminescence0.2–10 μM39.3 nM0.99712020[173]
Chlorogenic acidN,S-CDsFluorescent0.9314–83.82 μM0.3387 μM0.99702021[184]
4-NitrophenolCDs@PDAFluorescent2–34 μM7.29 μM
3.44 μM
0.992
0.993
2021[226]
2-4-6-TrinitrophenolN-CDsFluorescent0.3–3.3 μM0.11 μM0.99232021[72]
p-NitrophenolG-CDsFluorescent0–50 μM0.0175 μM0.99512022[76]
Where Tb-CDs: terbium doped carbon dots, PEGA-CDs: polyethyleneglycol bis(3-aminopropyl)-carbon dots, CDs@MIPs: carbon dots fabricated with molecularly imprinted polymers, N@CDs: nitrogen doped carbon dots, wsNP-CDs: water soluble nitrogen and phosphorous doped carbon dots, CDs@PDA: polydopamine encapsulated carbon dots, G-CDs: green-emitting carbon dots.
Table 12. Summary of the developed optical sensors for pesticides detection.
Table 12. Summary of the developed optical sensors for pesticides detection.
PesticideMaterialOptical SensorRange of DetectionLimit of DetectionLinear Correlation CoefficientYearReference
Methyl parathionTyr-CDsFluorescent1.0 × 10−10–1.0 × 10−4 M4.8 × 10−11 M0.9972015[121]
Paraoxon-ethylCDsFluorescent0–5.80 mM0.22 µM0.99742016[137]
Dichlorvos,
malathion,
ethion
CDs/Cu(II)/AChE/ATChClFluorescent6 nM–0.6 nM
6 nM–0.8 nM
8 nM–0.8 nM
3.8 nM
3.4 nM
4.2 nM
0.998
0.996
0.997
2016[245]
CarbarylN,S co-doped CQDSPhotoluminescent0.00003131–3.131 µM0.02485 µM-2016[209]
ParaoxonCQDsFluorescent0.1817–181.7 nM0.1817 nM0.9942017[211]
ParaoxonBChE-ATCh-MnO2-CDsFluorescent0.1817–18.17 nM0.05451 nM0.99412017[155]
ChlorpyrifosFe-modified CDsFluorescent0.028523–2.8523 µM0.008557 µM-2017[96]
Methyl parathionN-doped CDs-MPHFluorescent2.38–73.78 µM0.338 µM0.99342017[138]
ParaoxonCDsFluorescent0–1.817 µM0.00145 µM0.9932018[154]
AtrazineN-CQDsFluorescent0–1.0 nM3 pM0.98122018[213]
Atrazine; chlorpyrifos; imidacloprid; lindane; tetradifonCDsFluorescent-0.12 µM;
0.029 µM;
0.013 µM;
0.14 µM;
0.04 µM
-2019[246]
PretilachlorCDsFluorescent5.7 μM–61.5 μM2.9 µM0.98472019[60]
DiazinonCDsFluorescent0.8214 nM–16.43 μM0.8214 nM-2020[50]
Glyphosate1.4787 nM–29.574 μM0.01183 μM
Amicarbazone1.036 nM–20.72 μM0.002072 μM
Methyl-paraoxonB,N-CDsFluorescent0.1–15 μM0.1 μM0.99672020[180]
DiazinonCDsFluorescent0.02–10 μM0.01 μM0.97272020[62]
TrifluralinCa-modified CDsFluorescent-7.89 µM0.962020[220]
QuinalphosOPCD@UiO-66-NH2Fluorescent0–16 μM0.3 nM0.9922021[247]
ChlorpyrifosTEF-CDsFluorescent0.05–100.0 μM0.00599 μM0.99592021[63]
Quinalphos0.01–50.0 μM0.0057 μM0.9965
PyrimethanilCDsFluorescent0.5–75 μM14 nM0.99072021[182]
Thiophanate methylSCDs/Hg2+Fluorescent0.05–2.0 μM7.6 nM0.99982021[183]
2.0–5.0 μM0.9983
IsoprothiolaneCDsFluorescent1 mM-0.05 μM11.58 nM0.99212021[73]
λ-CyhalothrinCD-functionalized core-shell nanopsheresRatiometric fluorescent3.045–456.8 nM0.146 nM0.9882022[230]
ChlorpyrifosJ-CQDsFluorescent57.05–513.4 nM7.701 nM0.9932022[41]
Where Tyr-CDs: L- tyrosine methyl ester functionalized carbon dots, CDs/Cu(II)/AChE/ATChCl: carbon dots/copper ion/acetyl cholinesterase/acetylthiocholine, BChE-ATCh-MnO2-CDs: butyrylcholinesterase- acetylthiocholine- manganese dioxide- carbon dots, N-doped CDs-MPH: nitrogen-doped carbon dots and methyl parathion hydrolase, B,N-CDs: boron and nitrogen-doped carbon dots, Ca-modified CDs: calcium-modified carbon dots, OPCD@UiO-66-NH2: carbon dots derived from orthophenylenediamine incorporated UiO-66-NH2, TEF-CDs: carbon dots from Tagetes erecta flower, SCDs/Hg2+: sulfur-doped carbon dots/mercury ions, J-CQDs: Jatropha-carbon quantum dots.
Table 13. Summary of the developed optical sensors for explosive compound detection.
Table 13. Summary of the developed optical sensors for explosive compound detection.
Explosive CompoundMaterialOptical SensorRange of DetectionLimit of DetectionLinear Correlation
Coefficient
YearReference
2,4,6-TrinitrotolueneN-rich CNDsFluorescent10 nM–1.5 μM1 nM-2015[120]
2,4-DinitrotolueneAmine-functionalized CDsFluorescent1 mM–50 mM1 mM-2016[132]
TrinitrotolueneM-MIPs@CDsFluorescent-17 nM-2016[131]
4-Chloro-2,6-dinitroanilineCQDs@PAMAM-NH2Fluorescent1.0 × 10−5–6.0 × 10−5 M2 μM0.9942016[208]
2,4,6-TrinitrotolueneEthylenediamine-modified CDsFluorescent-0.213 µM0.9972017[139]
TrinitrotolueneNitrogen-doped CQDFluorescent4.4 nM–26.4 µM0.03258 µM0.99932018[214]
TrinitrotolueneCDs capped with EDAFluorescent44.03–220.14 nM57.24 nM0.954572018[99]
CDs (in the presence of nitric acid) capped with\EDA44.03–220.14 nM48.43 nM0.9752
88.06–264.17 nM21.88 nM0.99602
2,4,6-TrinitrotolueneCDs and Fe@SiO2-NH2Chemosensor44.03–8806 nM9.466 nM-2021[248]
2,4,6-TrinitrotoluenePEI-CQDsFluorescent0–38.17 µM0.4094 µM0.99792022[196]
Where N-rich CNDs: nitrogen-rich carbon nanodots, M-MIPs@CDs: mesoporous structured molecularly imprinted polymers capped carbon dots, CQDs@PAMAM-NH2: carbon quantum dots functionalized with amine groups by PAMAM dendrimer, EDA: ethylenediamine, CDs and Fe@SiO2-NH2: carbon dots and the magnetism of amino-functionalized magnetic core-shell nanomaterial, PEI-CQDs: polythylenimine capped carbon quantum dots.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Omar, N.A.S.; Fen, Y.W.; Irmawati, R.; Hashim, H.S.; Ramdzan, N.S.M.; Fauzi, N.I.M. A Review on Carbon Dots: Synthesis, Characterization and Its Application in Optical Sensor for Environmental Monitoring. Nanomaterials 2022, 12, 2365. https://doi.org/10.3390/nano12142365

AMA Style

Omar NAS, Fen YW, Irmawati R, Hashim HS, Ramdzan NSM, Fauzi NIM. A Review on Carbon Dots: Synthesis, Characterization and Its Application in Optical Sensor for Environmental Monitoring. Nanomaterials. 2022; 12(14):2365. https://doi.org/10.3390/nano12142365

Chicago/Turabian Style

Omar, Nur Alia Sheh, Yap Wing Fen, Ramli Irmawati, Hazwani Suhaila Hashim, Nur Syahira Md Ramdzan, and Nurul Illya Muhamad Fauzi. 2022. "A Review on Carbon Dots: Synthesis, Characterization and Its Application in Optical Sensor for Environmental Monitoring" Nanomaterials 12, no. 14: 2365. https://doi.org/10.3390/nano12142365

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

Article Metrics

Back to TopTop