Abstract
Environmental contamination by pharmaceuticals has become a matter of concern as they are released in sewage systems at trace levels, thus impacting biological systems. Increasing concerns about the low-level occurrence of pharmaceuticals in the environment demands sensitive and selective monitoring. Owing to their high sensitivity and specificity carbon dots (CDs) have emerged as suitable fluorescent sensors. This review discusses the current scenario of the status of pharmaceuticals in the environment, limitations associated with traditional techniques employed for their detection, and benefits offered by CDs like easy surface modification and tunable optical properties for sensing applications. Several representative means by which CDs interact with other molecules such as inner filter effect (IFE), dynamic quenching (DQ), static quenching (SQ), Förster resonance energy transfer (FRET), among others, are also discussed along with co-referencing fluorophores to design sensors. Based on developments described herein, CDs-based sensors can be expected to sense pharmaceuticals ranging from nanogram to picogram, target real-time industrial and spiked sample analysis, etc., which provides direction for future research.
Keywords: Carbon dots, Fluorescent probe, Forster resonance energy transfer mechanism, Quenching, Pharmaceutical detection
Graphical abstract
Highlights
-
•
Carbon dots (CDs) are unique fluorescent labels with photoluminescence (PL) properties that make them perfect for sensing.
-
•
Comprehensive comparison of traditional and novel CD-based analytical methods shows CDs' pharmaceutical detection benefits.
-
•
Dopants and surface modifiers affect CD PL characteristics, improving selectivity and sensitivity, according to the study.
-
•
Carbon dots (CDs) detect pharmaceuticals effortlessly, selectively, sensitively, and economically.
1. Introduction
Pharmaceuticals are essential in healthcare for treating various conditions, but they must be administered carefully to avoid adverse effects. Concerns about environmental contamination have risen as pharmaceuticals and their metabolites, classified as “emerging contaminants” for over 15 years, are frequently detected in wastewater, eventually entering natural water sources through industrial discharge and human excretion [1,2]. Developed nations often face issues with lifestyle drug misuse, while developing countries struggle with counterfeit life-saving drugs [3]. Pharmaceuticals in wastewater, with concentrations from micrograms to nanograms per liter, pose risks as current treatment systems fail to remove them entirely, allowing these compounds and their by-products to reach drinking water sources [4]. Moreover, organic solvents in pharmaceutical analysis contribute significantly to environmental and health issues, especially low-boiling-point solvents like n-hexane, toluene, methylene chloride, benzene, chloroform, methanol, ethanol, and acetonitrile, which have known toxic effects [5]. For instance, benzene can cause anemia, hexane is a neurotoxin, and prolonged exposure to chloroform can damage the liver and kidneys. Recognizing the severe impact of these solvents, some are classified as “red solvents” by Pfizer due to their high toxicity [6].
Fluorimetric techniques offer an efficient solution to reduce medication toxicity and solvent use. They require minimal sample preparation and solvent consumption. Simple instruments can detect fluorescent compounds at concentrations up to a thousand times lower than absorption spectrophotometry. By dissolving samples in a suitable solvent, chemical changes can make non-fluorescent compounds detectable, and applicable to both organic and inorganic substances.
Medicines are essential for treating and preventing diseases, and for maintaining public trust in healthcare systems. However, all medications carry a risk of side effects, making it crucial to monitor both intended and unintended outcomes to balance risk and efficacy. Detecting impurities in both active ingredients and formulations is therefore vital [7]. Early identification of unexpected substances, especially in new drugs, is important to reduce health risks. Pharmaceuticals, even at low doses, can be highly active, with contraceptives at nanograms/Liter (ng/L) levels potentially disrupting endocrine systems in municipal water. Excretion and improper disposal release pharmaceuticals into the environment, highlighting the need for effective detection to minimize their environmental and animal toxicity [[7], [8], [9]].
As pharmaceutical research advances, the development of novel and highly selective analytical techniques is crucial for faster, more efficient methods that offer cost savings and reduce solvent consumption. This work evaluates recent quantitative analytical methods and their applications in pharmaceutical analysis, which are essential for quality control and require quick, reliable, and clear results. Key techniques used for the quantitative analysis of pharmaceutical compounds include capillary electrophoresis (CE), high-performance liquid chromatography (HPLC), Ultraviolet–Visible (UV/Vis) spectrophotometry, fluorimetry, titrimetry, voltammetry (in electroanalytical methods), thin-layer chromatography (TLC), gas chromatography (GC), and vibrational spectroscopies (Fig. 1).
Fig. 1.
Traditional analytical techniques.
UV/Vis spectrophotometry is a widely used method in pharmaceutical analysis, measuring how much ultraviolet (UV) (190–380 nm) or visible (380–800 nm) radiation a chemical absorbs. It works by assessing the relationship between two beams of UV light, where absorption occurs when the energy matches the electronic transitions in the molecule. Combining UV and Vis spectrophotometry offers convenience due to its rapid analysis and ease of use [10,11]. Fluorimeters and spectrofluorometers, which measure fluorescence (FL), offer higher sensitivity compared to other absorption methods. The limitations of FL spectroscopy arise from the sensitivity of detectors to different wavelengths and variations in energy intensity. Infrared (IR) spectroscopy, covering near IR (0.8–2 μm), mid IR (2–15 μm), and far IR (15–1000 μm) ranges, provides detailed structural information of organic and inorganic compounds, with the fundamental 2–15 μm range being most informative for chemical structure analysis [12]. Nuclear Magnetic Resonance (NMR) spectroscopy examines molecular structure at the atomic level, particularly using the 1H and 13C isotopes, providing detailed structural data through molecular vibrations [13]. High-Performance Liquid Chromatography (HPLC) is an advanced technique that improves upon traditional column chromatography by enhancing separation speed, resolution, accuracy, and sensitivity. HPLC offers benefits like small sample size, customizable tests, and precise data generation, making it a valuable tool in pharmaceutical analysis. Spectrofluorimetry, the most sensitive method, can detect substances at the femtogram level, unlike other methods that require microgram-level sample preparation [14].
This review differs from existing ones by providing a detailed exploration of the synthesis, surface functionalization, and photoluminescent properties of carbon dots (CDs), with a particular emphasis on their emerging applications in photocatalysis, energy, and sensing. Unlike other reviews, we focus on the potential of CDs as safer alternatives to conventional fluorescent compounds, particularly in in-vivo analysis. The specific aims of this review are to offer a comprehensive overview of recent advancements in CDs, highlighting their diverse applications and exploring their potential as substitutes for toxic compounds, as well as their role in electron storage and transport when exposed to light. Furthermore, this review addresses key gaps in the current literature, including the need for alternative materials that can efficiently remove unfiltered contaminants in wastewater treatment and the underutilized potential of CDs in nanomaterial-based solutions. By examining these aspects, we aim to uncover new avenues for the application of CDs in both biotechnological and environmental fields.
Based on the recent advancements in the field of CDs, this review hypothesizes that CDs, due to their unique photoluminescent properties, surface functionalization, and environmental compatibility, can serve as safer and more efficient alternatives to traditional fluorescent compounds in various applications, including sensing, photocatalysis, and environmental monitoring. Furthermore, it is hypothesized that CDs have the potential to overcome the limitations of conventional wastewater treatment systems by offering innovative solutions for contaminant removal, thus providing significant benefits in both biotechnological and environmental fields.
2. Nanomaterials in sensing of pharmaceuticals
Nanomaterials are materials that have at least one dimension in the nanometer (nm) range and range in size from approximately 1 to 100 nm [15]. According to the literature, most drug identification methods are based on chromatographic techniques. These techniques are well-established and accepted by regulatory authorities. However, it has several drawbacks related to relatively high cost, analysis time, and pretreatment steps [16]. The incorporation of nanomaterials has enabled the development of novel and effective sensor platforms [17]. The use of nanomaterials for biosensor development has attracted a lot of interest, and carbon nanomaterials (CNMs) are at the forefront. CNMs are composed of sp2-bonded graphitic carbon and are mainly classified into fullerenes (0-dimensional), carbon nanotubes (CNTs) (1-dimensional), and graphene (2-dimensional) [18]. CNMs offer superior electrical conductivity, chemical stability, biocompatibility, and strong mechanical strength because of special characteristics including surface-to-volume ratio [19].
As anticipated, these characteristics can affect the stability and selectivity of nanomaterials, as can the capacity to form hydrogen bonds, stacking, dispersion forces, dative bonds, and hydrophobic interactions [20]. A quick search of the Web of Science database over the past five years (2013–2018) for the terms “fullerenes,” “CNTs,” and “graphene” yields over 12,000, 32,000, and 98,000 papers, respectively [21]. The facile functionalization and modification of such nanomaterial-based biosensors facilitated improved efficiency in the detection of antibiotic residues and narrow therapeutic index (NTI) drugs [22]. Additionally, nanoparticles (NPs) can be used for targeted drug delivery [23] and high drug-loading capacities [24]. They also show great potential in cancer therapy by improving drug performance, reducing systemic side effects, and increasing therapeutic efficacy [25]. However, there are complexities, difficult and lengthy synthetic processes, and the toxicity of heavy metal quantum dots (QDs) may hinder their application in biosensing [26]. Therefore, nanomaterials can be used to determine drugs [27]. However, some of them are toxic to humans. A common mechanism by which metal oxide NPs cause toxicity is a combination of the NPs properties and their propensity to generate reactive oxygen species [ROS] and cause toxicity in cells, genes, and neurons.
2.1. Carbon dots (CDs)
Quasi-spherical particles having a diameter of less than 10 nm, known as fluorescent CDs were first identified in 2004 [28]. CDs are a new class of nanostructures made of carbon (C) that have intriguing characteristics and small sizes. Surface-functionalized carbonaceous NPs known as CDs have extraordinary properties, including adjustable FL [29]. CDs were reported to have good biocompatibility, less toxicity [30], high photoluminescence (PL) intensity [31], high chemical stability [32], and possess excellent biological, physical, and chemical properties, thus having great potential in various applications [33,34]. Only a few of the benefits associated with their luminescence are their outstanding water solubility, biocompatibility, non-toxicity, high sensitivity to the environment, and apparent electron-donating and receiving capacities [35]. CDs exhibit appealing optical characteristics such as size-dependent PL [36], photo-induced electron transfer (PET), up-conversion luminescence, chemiluminescence, and electrochemiluminescence (ECL) [37]. These dots possess an inner sp2 and outer sp3 hybridized structure that frequently contains oxygen-containing functional groups. The surface of CDs imparts several traits to them, including effortless electron transfer, which can confer anti- or pro-oxidant behavior [38].
In addition, CDs can be easily functionalized with hydroxyl, carboxyl, carbonyl, amino, and epoxy groups on their surfaces. This additional advantage enables them to bind readily with both inorganic and organic moieties. The functional groups present on the surfaces of CDs allow them to adopt either hydrophilic or hydrophobic properties, providing the necessary thermodynamic stability in different solvents, particularly in water [39]. Moreover, CDs have exceptional sensing properties like multiplex, selective, and specific detectability. Abundant functional groups (such as amine, carboxyl, hydroxyl, etc.) or polymer chains on the surface of CDs make them highly soluble in aqueous solutions and easily functionalized with other nanomaterials [40]. Because of their extremely sensitive responses to target molecules and tunable surface functional groups, these characteristics make CDs particularly appealing in sensing applications. Another fascinating property of CDs is their tunable emission, characterised by multiple FL colors under various excitation wavelengths [41].
2.2. Properties of CDs
Without a crystal structure, CDs are always spherical and split into carbon NPs. All CDs have linked or altered chemical groups, like chains made of oxygen, amino acids, polymer etc., on their surfaces. X-ray diffraction (XRD), Raman spectroscopy, and high-resolution transmission electron microscopy (HRTEM) are the direct characterization techniques for the carbon core XRD. NMR, X-ray photoelectron spectroscopy (XPS), Fourier transforms infrared (FTIR), and matrix-assisted laser desorption ionisation time-of-flight (MALDI-TOF) are used to assess the grafting of chemical groups. These luminous CDs are therefore not made of “pure” C compounds. Various properties of CDs are depicted in Fig. 2. The PL behavior of these CDs is mostly determined by the hybridization and coupling between the C core and surrounding chemical groups [42]. In the UV spectral region, CDs typically exhibit considerable absorption with an extension into the visible spectrum. Sometimes, at a wavelength that is significantly longer than the UV absorption peak, a shoulder or a weak peak is also seen. Longer wavelength UV absorption, which typically takes the form of a shoulder or a weaker peak between 300 nm and 400 nm, is attributed to n → π∗ transitions of C=O. Shorter wavelength UV absorption, which is roughly between 200 nm and 350 nm, is attributed to π → π∗ electronic transitions of C=C and C=N [43]. The UV–Vis spectra of Yellow-Green CDs (YG-CDs) give several absorption bands at 270 nm and 382 nm, just like Blue-CDs (B-CDs) do. However, YG-CDs have a higher absorption intensity than B-CDs, which may be due to the latter's higher degree of carbonization in a more acidic hydrothermal environment [44]. The PL emission, comprising excitation-dependent and excitation-independent PL, which is caused by core-related and surface state-related emissions, is one of the CDs most fascinating characteristics [45]. Further remarkable FL characteristics of CDs include tunable PL emission, excitation wavelength-dependent PL emission, strong FL stability, and effective photobleaching resistance. Some CDs can have up-conversion PL (UCPL) emission characteristics, which means that the emission wavelengths of carbon quantum dots (CQDs) are shorter than their excitation wavelengths, in contrast to standard PL emission. The most common explanations for the up-conversion FL phenomenon are often two-photon excitation and anti-Stokes PL emissions [46]. CDs exhibit ECL capabilities and are useful for metal ions detection, optical devices, and biosensors [47]. CDs have a protracted, intense FL emission (up to a year). As CDs can withstand a wide pH range (from 3 to 12), they exhibit excellent photobleaching impedance. Excitation following direct oxidation, amplification, or inhibition of luminescence are all ways that CDs might produce chemiluminescence [40].
Fig. 2.
Properties of CDs.
2.3. Synthesis of CDs
There are currently many different synthetic approaches that can be used to make CDs. The majority of studies focus on easy, affordable, size-controllable, or large-scale synthetic ways to produce CDs of superior quality. Chemical and physical processes are used to create synthetic materials, respectively, depending on the characteristics of the transformations of C sources to final products. Moreover, two classes of top-down and bottom-up may characterise existing synthetic techniques when taking into account the link between the sources and products (Fig. 3). The synthesis process may also require additional purification of the end products using techniques like centrifugation, electrophoresis, dialysis, etc. One study by Zhu et al., for instance, used three cycles of concentration/dilution to separate CDs from other reactants [48]. For this class, CDs are produced through oxidation, laser ablation, arc discharge, and electrochemical release on somewhat macroscopic C structures including CNTs, graphite columns, graphene, suspended C powders, etc [49]. In bottom-up processes, CDs are made from a variety of small molecule precursors; most of them enable the manufacture of functionalized CDs in a single step. It consists of pyrolysis, microwave synthesis, ultrasound-assisted synthesis, and hydrothermal synthesis [50]. Several C sources are utilized, including amino acids, citric acid, sugar, and even food waste.
Fig. 3.
Schematic diagram showing various sources and methods for synthesis of CDs.
2.3.1. Top-down methods
Chemical oxidation frequently involves oxidising the substrate with an oxidative reagent. Chemical oxidation is also a quick and efficient approach to large-scale production. Coal, wood, and coconut activated carbons-all readily available in the marketplace-were the C sources. Nitric acid made it simple to etch CDs out of the amorphous structure of activated carbons. Chemicals with amine ends were then used to carry out the passivation process. The three C sources' end products had a constrained size distribution and spectacular luminosity [[51], [52], [53]]. Because a huge volume of fluorescent CDs can be created in a short period, the laser ablation including solid targets in liquid (LASL) approach for producing fluorescent CDs can be very rapid and efficient. For the synthesis of high-quality or smaller-sized nanostructures, other synthesis procedures, that include the ablation of suspended micro particles or powders, need a longer ablation period. In this technology, greater control of the size distribution can be achieved by carefully controlling heat dissipation via wavelength and intensity change. In comparison to microscale-ablated particles, the size of the emitting NPs generated in the study is in the nanoscale range [51]. Top-down electric arc or arc-discharge synthesis was used to create the first luminous CDs. In order to extract fluorescent C from raw small wall nanotubes (SWNTs), soot, and other materials using a preparative electrophoretic method, it was shown that the fluorescent materials PL quantum yield (PLQY) could reach 1.6 percent at 366 nm excitation wavelength. The benefit of this approach is that the CDs may emit multiple fluorescent colors under UV light without undergoing any surface alteration, and it just needs a simple purification procedure. It is also portable and environmentally friendly. The formation of mixes, low yield, difficulties in surface and compositional modification, and difficulty in these alterations are also drawbacks [52]. A C material, such as graphite, CNTs, or C fibre electrodes, is chemically cut using the electrochemical method, which is influenced by an electric field. The method's low cost and simplicity make it useful. Interestingly, by altering process variables like the temperature, electric field, CNT diameter, and concentration of the supporting electrolyte, CDs made electrochemically can have their sizes adjusted [53].
2.3.2. Bottom-up methods
A simple bottom-up hydrothermal approach has been reported to make CDs. The colorless solution first needs to close in a Teflon-coated stainless-steel autoclave and then be kept at a definite temperature for a particular time. The autoclave was left to naturally cool to room temperature following the reaction. The solution is either centrifuged or passed through dialysis tubing of a certain molecular weight to remove the precipitation and to eliminate leftover small molecules, the CDs were dialyzed against ultrapure water at room temperature using a membrane (MCWO 1000 Da) [54]. Depending on the frequency and strength of the applied field, ultrasound is a typical laboratory instrument that can be used to emulsify mixtures, drive chemical processes, and nebulize liquids into fine mists [55]. Microwaves are electromagnetic wave types with a broad wavelength range of 1 mm (mm) to 1 m (m) that are frequently employed in daily life and science. The microwave can also deliver high energy to break the chemical bonds in the substrate, just like a laser can. The synthesis of CDs is thought to be more energy-efficient when done in a microwave, and the reaction time can also be significantly reduced. The substrate is typically pyrolyzed and surface functionalized during microwave-assisted synthesis [56].
2.4. CDs as a sensor
The sensing mechanism of the CDs depends upon the chemical bonding of the fluorescent probe. Various mechanisms of CDs have been available (Fig. 4) for sensing the analyte, these include Förster resonance energy transfer (FRET) [57,58], inner filter effect (IFE) [59,60], photoinduced electron transfer (PET) [61,62], intramolecular charge transfer (ICT) [63,64], static quenching (SQ), dynamic quenching (DQ) [65,66], aggregation caused quenching (ACQ), aggregation-induced emission (AIE) [67].
Fig. 4.
The sensing mechanism of CDs.
2.4.1. Different mechanisms of CD
The FL behavior of CDs is governed by various mechanisms that influence their emission properties and sensing capabilities. These mechanisms include FRET, IFE, SQ and DQ, PET, ICT, as well as phenomena like ACQ and AIE.
2.4.2. Förster resonance energy transfer (FRET)
In FRET, energy is transferred non-radiatively from an excited donor molecule (D), typically a fluorophore, to a close-by ground-state acceptor molecule (A) via long-range dipole-dipole interactions. For efficient FRET, the acceptor must absorb energy at the donor's emission wavelength (s), though it does not need to emit this energy fluorescently. The efficiency of FRET is significantly influenced by three main factors: the degree of spectral overlap between the donor's emission and the acceptor's absorption, the relative orientation of the transition dipoles, and most importantly, the distance between D and A, usually in the range of 10–100 Å (Å), which is comparable to the size of many biological macromolecules [57,58].
2.4.3. Inner filter effect (IFE)
The IFE involves an absorber in a sensor system that modulates the excitation or emission of light from a fluorescent molecule. In an IFE-based sensing system, both an absorber and a fluorescent molecule must coexist. The absorber's absorption spectrum overlaps with the excitation and/or emission spectra of the fluorophore. This overlap enables the absorber to control and influence the FL emission. The system requires the absorber to exhibit a selective response to the analyte concentration, while the FL intensity of the fluorophore remains unaffected by the analyte, ensuring its role as an indicator in the sensor system [59,60].
2.4.4. Photoinduced electron transfer (PET)
PET is a mechanism in which a fluorophore is linked to a recognition receptor via a spacer. FL quenching in PET occurs due to intramolecular electron transfer between the receptor and the fluorophore. Two types of PET processes are recognized: (1) Acceptor-PET (A-PET): Electron transfer from the receptor to the fluorophore occurs when the receptor's Highest Occupied Molecular Orbital (HOMO) energy level is higher than that of the fluorophore. (2) Donor-PET (D-PET): Electron transfer from the excited fluorophore to the receptor's Lowest Unoccupied Molecular Orbital (LUMO) level occurs during D-PET. The PET process is inhibited when the receptor binds to a target analyte, leading to restored FL emission [61,62].
2.4.5. Intramolecular charge transfer (ICT)
ICT involves electron transfer from a donor (D) to an acceptor (A) within a fluorescent probe upon light stimulation. When the donor interacts with an analyte, its electron-donating ability decreases, increasing the HOMO-LUMO energy gap and causing a blue shift in FL emission. Conversely, the interaction of the acceptor with the analyte decreases the energy gap, resulting in a red shift. Thus, both FL intensity and emission wavelength can be used for analyte detection. The varying dipole moments between the ground and excited states of ICT probes make them sensitive to different solvent environments, making them ideal for detecting changes in solvent conditions [63,64].
2.4.6. Quenching mechanisms
2.4.6.1. Static quenching (SQ)
SQ occurs when a non-fluorescent ground-state complex forms between the quencher and the CDs. This complex reduces the FL intensity of the CDs and alters their absorption spectrum. Temperature increases can destabilize this ground-state complex, reducing the quenching effect.
2.4.6.2. Dynamic quenching (DQ)
DQ involves the interaction between an excited-state fluorophore and a quencher, resulting in energy or charge exchange that returns the fluorophore to its ground state. Unlike SQ, DQ affects only the excited state and does not alter the absorption spectrum of the fluorophore. The FL lifetime of the fluorophore is reduced in the presence of a quencher. Higher temperatures typically enhance the DQ effect, increasing the rate of quenching interactions [65,66].
The reduction in FL intensity due to quenching can be quantified using the Stern-Volmer equation:
| F0/F = 1 + K[Q] = 1 + kqτ0[Q] |
Where K is the Stern-Volmer quenching constant, [Q] is the quencher concentration, τ0 is unquenched lifespan, and kq is the bimolecular quenching constant, and F and F0 are FL intensities in the presence and absence of the quencher, respectively.
2.4.6.3. Aggregation-caused quenching (ACQ) and aggregation-induced emission (AIE)
In ACQ high concentrations or solid-state fluorophores often exhibit diminished FL due to self-quenching. This effect arises from interactions such as hydrophobic effects, stacking, and hydrogen bonding, leading to non-radiative decay pathways. In contrast to ACQ, AIE materials show enhanced FL in high concentrations or solid forms. In dilute solutions, non-radiative energy dissipation occurs via intramolecular rotation. However, in the aggregated state, this rotation is restricted, leading to suppressed non-radiative decay, and enhanced FL emission organization groups similar principles together, improving the clarity and logical flow of the discussion on various FL mechanisms associated with CDs [67].
2.5. Modification of CDs
Doped-CDs are carbogenic NPs with an average size of less than 10 nm that have atomic impurities added like nitrogen (N), sulfur (S), phosphorus (P), boron (B), etc. (Fig. 5) during the production process to enhance their optical, electrical, and chemical capabilities. One can distinguish between single- and multiple-doped CDs (Fig. 6) based on how many atomic impurities have been added to the structure of the CDs. One of the most well-known options is N since it has five valence electrons, a similar atomic size to C, and can form bonds with C atoms. When the N content in the CDs rises, the FL peak has shown that it will move to a longer wavelength under the same stimulation. The observation's justification is based on the N-doped material forming new FL origins. In other research, N is also seen to improve the efficacy of emissions as opposed to displacing them. The mechanism is thought to involve electrons in the conduction band and the occurrence of an upward shift in the Fermi level. Hence, various polymers used as a precursor material for C and N sources are dodecyl-grafted-poly(isobutylene-alt-maleic-anhydride) [68], hydrosoluble chitosan [69], dried shrimp shells [70], citric acid, and urea [71], dried prawn shell [72], oolong tea [73], monkey grass [74], folic acid (FA) and phosphoric acid (H3PO4) [75], lysine and ortho-phosphoric acid [76], p-phenylenediamines and ammonia water [77], gum ghatti and ethylenediamine [78], Borassus flabellifer (B. flabellifer) and aq. Ammonia [79], m-phenylenediamine [80], Ru(bpy)2(phen-NH2) [81], 4-aminophenol [82], ethylenediamine, and Kentucky bluegrass [83], β-resorcylic acid and ethylenediamine [84], black soya beans [85], glucosamine and ethylenediamine [86], citric acid and diethylenetriamine [87], highland barley as C source and ethanediamine as N resource [88], biomass bacterial cellulose [89], tartaric acid and urea [90], adenosine [91], dried chrysanthemum buds, and ethylenediamine [92]. The band-gap energy of photoexcited electrons may be modified by the S atom's density of states or emissive trap states, giving functionalized CDs additional favorable properties [[93], [94], [95], [96], [97], [98], [99], [100], [101], [102]], sodium citrate solution, and sodium thiosulfate [101], blackstrap molasses, H2O2, and garlic powder [103], mercaptosuccinic acid [104], sulphuric acid (H2SO4) [105].
Fig. 5.
Doped CDs sources.
Fig. 6.
Single, multi, and co-doped CDs.
Thus various polymers used as precursor material for C and S sources are sodium thiosulfate and sodium citrate [106]. With only three valence electrons, B (one less than C). As a result, adding a B atom to a C cluster causes the generation of p-type carriers inside CDs, altering their electrical structures and optical characteristics [[107], [108], [109], [110], [111]]. Thus various polymers used as precursor material for C and B sources are phenylboronic acid [109], boric acid, urea and citric acid [110], glucose and boric acid [112], L-ascorbic acid and boric acid [113], citric acid, H2SO4, ammonia water, boric acid, and isopropyl alcohol [114], boric acid and ethylenediamine [115], citric acid monohydrate, thiourea, and boric acid [116]. Moreover, citric acid, rhodamine B, methylene blue,1, 2-diboranyethane, N, N-dimethylformamide (DMF, anhydrous, 99.8 %) [117], etc. The P atom can have a significant impact on the chemical and electrical structures of semiconductors since it is a major electron donor in semiconductors, which are endowed with promising uses in photovoltaic devices and catalysts. The bigger size of the P atom compared to the C atom directly contributes to an increase in disorders inside the P-doped graphite carbon backbone [[118], [119], [120], [121]].
The various polymers used as precursor materials for C and P sources are L-threonine and phosphorous pentoxide [122], sucrose and H3PO4 [123], dextrose and disodium phosphate (Na2HPO4) [124], beta-cyclodextrin and sodium pyrophosphate [125], lactose, and concentrated H3PO4 [126], orthophosphoric acid and glucose [127]. Nitrogen, Sulfur co-doped carbon dots (N, S-CDs), one of the heteroatom-containing CDs, display their remarkable FL behaviors [128]. Example: thiourea and glutathione [129], p-aminobenzenesulfonic acid [130], DL-malic acid, ethanolamine and ethane-sulfonic acid [131], citric acid and L-cysteine [132], ammonium persulfate, glucose, and ethylenediamine [133], citric acid and thiourea [134], sodium citrate and thiourea [135], methionine and citric acid [136], basic fuchsin and sulfosalicylic acid [137], 4-aminophenylboronic acid hydrochloride and 2, 5-diaminobenzenesulfonic acid [138], ethylenediamine, L-cysteine and malic acid [139], citric acid and dithiooxamide [140], garlic [141], citric acid monohydrate and cysteamine [142], urea and H2SO4 [143], citric acid and thiosemicarbazide [144], trisodium citrate dehydrate and L-cysteine [145], H2SO4 [146], methionine and acrylic acid [147], polyvinylpyrrolidone and poly(4-styrenesulphonic acid) sodium salt [148], citric acid and arginine (A), histidine (H), lysine (L), cysteine (C), and methionine (M) [149]. In addition, other combination includes heating egg white, egg yolk, pigeon feathers or pigeon manure [150], garlic and alfalfa [151], thiomalic acid and diethylenetriamine [152], rose petals, ethylenediamine, L-cysteine [153], para-benzoquinone and L-cysteine [154], cigarette and H2SO4 [155], thioctic acid and triethylene tetramine [156].
Heteroatoms like N and B have been created to enhance the optical performance of CDs. The luminous characteristics of these CDs are fascinating, including temperature- and pH-dependent FL responses and excellent quantum yield [157]. Solanum betaceum (S. betaceum) fruit extract [158], trisodium citrate, urea, and boric acid [159], adenine and aminobenzene boronic acid monohydrate [160], citric acid, boric acid and ethylenediamine [161], 2,5-diaminobenzenesulfonic acid and 4-aminophenyl boronic acid [138], petroleum coke [162], 3-aminophenylboronic acid [163], citric acid, boric acid, and tris base [164], citric acid, urea, and boric acid, etc. are the various precursors used for the synthesis of N, B -doped CDs [165]. On the periodic table, N and P sit next to C and serve as important tracking elements in the field of biomedical imaging. These two components change the optical and electronic characteristics of CDs and advance our fundamental knowledge of their PLQY. Moreover, this may result in multifunctional applications in photothermal therapy (PTT) and photoimaging. Some precursors used for the synthesis of N, P-doped CDs are ethylenediamine and N-phosphonomethylaminodiacetic acid [166], 1,4-naphthalenedicarboxylic acid and urea [167], glucose, ammonia, and H3PO4 [168], adenosine 5′-mono phosphoric acid disodium salt dehydrate [169], diethylenetriaminepenta and m-phenylenediamine [170], glucose, polyethyleneimine and H3PO4 [171], phytic acid and L-arginine [172], 1,2-ethylenediamine, phthalic acid and H3PO4 [173], lily bulb [174], citric acid, tris (2-aminoethyl) amine and orthophosphoric acid [175], sodium citrate and diammonium phosphate [176], maize and urea [177], alendronate sodium [178], polyethylene glycol (PEG) (C source), and H3PO4 [179]. Various sources used for the synthesis of N, Cl-doped CDs are o-phenylenediamine, hydrochloric acid (HCl) and selenourea [180], cetylpyridinium chloride [181], 4-chloro-1,2-diaminobenzene and dopamine [182], p-phenylenediamine, melem hydrazine, and aluminum chloride hexahydrate [183], glycerine, choline chloride and urea [184], chloranil and triethylenetetramine [185], glucose, concentrated HCl and 1,2-ethylenediamine [186]. B, N, and S-doped were synthesized by 2, 5- diaminobenzene sulfonic acid, and 4- aminophenyl boronic acid hydrochloride [187]. The N, S, P, and Cl-doped were synthesized by glucose and ethylenediamine, and then H2SO4, concentrated H3PO4, and concentrated HCl were added orderly into the beaker to get the CDs [188].
3. Applications of fluorescent CDs based sensors for pharmaceuticals
As CDs are reported best for sensing pharmaceuticals. Fig. 13 illustrates a variety of CDs applications along with drug identification in pharmaceuticals. So far, some of the drugs listed in Table 1 are detected by carbon nanodots (CNDs) through a different mechanism of CDs. The study investigated the impact of deoxygenation and radical scavengers on the FL intensity of the C-dots-Ni(IV) system. It was proposed that when Ni(IV) reacts with dissolved oxygen in an alkaline solution, it produces the superoxide radical (•O2−). Both Ni(IV) and •O2− can interact with C-dots, generating C-dot•+ and C-dot•−, which then undergo electron transfer annihilation, forming an excited-state C-dot∗ as the final emitter. The target analyte was found to inhibit chemiluminescence by competitively reacting with Ni(IV). This led to the development of a flow-injection chemiluminescence method, which proved to be a fast, simple, sensitive, and robust technique for detecting trace levels of reducing compounds [189]. In a separate experiment, CDs were synthesized using sodium citrate and ammonium bicarbonate through a hydrothermal process. The FL intensity of the CDs was found to be pH-dependent, increasing with pH up to 6.0, before stabilizing between pH 7.3 and 7.7. The study also tested the effects of various interfering substances, revealing that D-penicillamine (D-PA) caused significant quenching of the FL, while increasing concentrations of Hg2+ also reduced the FL intensity, with near-total quenching occurring at 4 × 10⁻⁵ mol L⁻1 of Hg2+ [190].
Fig. 13.
Various applications of CDs.
Table 1.
Pharmaceutical applications of fluorescent CDs-based sensors.
| Sr. No. | Drug name | Wavelength excitation(λex) | Wavelength emission(λem) | Linearity range | Mechanism | Ref |
|---|---|---|---|---|---|---|
| 1. | Paracetamol | 360 nm | 445 nm | 4.48x10−7-8.96 x10−5 μM | Ni(IV) reacted with CDs produced could form an excited state C-dot∗ which interacts with the target analyte due to its competitive reaction with Ni(IV). | [189] |
| 2. | D-penicillamine | 358 nm | 442 nm | 2–24 μmol/L | Fluorescent switch sensor | [190] |
| 3. | Kaempferol | 370 nm | 490 nm | 3.5–4.9 μM | Non-radiative energy transfer (FRET) | [191] |
| 4. | Amantadine | 335 nm | 446 nm | – | FRET | [192] |
| 5. | Chloramphenicol | 400 nm | 525 nm | – | FRET | [192] |
| 6. | Hyperin | 300 nm | 540 nm | 0.22–55 μM | Non-radiative energy transfer | [193] |
| 7. | Sumatriptan | 380 nm | 446 nm | 1–20 μM | The fluorescent quenching observed is caused by the formation of hydrogen bonds between the nitro groups of DNB-CDs and the aliphatic -NH group on the phenyl aromatic ring of SUM. | [194] |
| 8. | 6-Mercaptopurine | 370 nm | 445 nm and 565 nm | 1.0–100.0 μmol L−1 | FRET | [195] |
| 9. | Gentamicin | 340 nm | 385 nm | 0 to 2.9 × 10−4 mol/L | FRET | [196] |
| 10. | Gentamicin and kanamycin | 390 nm | 482 nm | 200–2000 nM | FRET | [197] |
| 11. | α-glucosidase | 410 nm | 510 nm | 100 μL | IFE | [198] |
| 12. | Iron | 370 nm | 432 nm | 0–1.5 mM | IFE | [199] |
| 13. | Hematin | 314 nm | 362 nm | 0–75 μM | IFE | [200] |
| 14. | Cefixime | 350 nm | 455 nm | 0.2 × 10−6 M to 8 × 10−6 M | IFE | [201] |
| 15. | Tetracycline | 345 nm | 435 nm | 0.5–25 μM | IFE | [202] |
| 16. | Chlortetracycline | 370 nm | 447 nm | 0.4–20 μg/mL | IFE | [203] |
| 17. | Tetracycline | 350 nm | 420 nm | 0–50 μM | IFE | [204] |
| 18. | Tetracycline | 370 nm | 440 nm | 0.5–60 μM | Internal filtering effect | [205] |
| 19. | Tetracycline | 360 nm | 440 nm | 0.1–4.0 μg mL−1 | IFE | [206] |
| 20. | Tetracycline and oxytetracycline and chlortetracycline | 381 nm | 456 nm | TC (1–60 μM), OTC (1–40 μM) and CTC (1–80 μM) |
IFE | [207] |
| 21. | Nimesulide | 330 nm | 408 nm | 0–100 μM | IFE | [208] |
| 22. | Curcumin | 380 nm | 467 nm | 0.1–35 μM | IFE | [209] |
| 23. | Clioquinol | 360 nm | 440 nm | 40.0–400.0 μmol L−1 | IFE | [211] |
| 24. | Berberine hydrochloride | 370 nm | 452 nm | 0.5–30.0 μmol/L | IFE | [212] |
| 25. | Diacerein | 325 nm | 412 nm | 2.5–17.5 μg/mL | IFE | [213] |
| 26. | Captopril | 520–550 nm | – | 1–50 μM | IFE | [214] |
| 27. | Daunorubicin and doxorubicin | – | 450 nm | DAN (54.37–13594.34 nmolL−1) DOX (86.2–17242 nmolL−1) |
IFE | [215] |
| 28. | Mitoxantrone | 560 nm | 655 nm | 0.096–30 μM | IFE | [216] |
| 29. | Gefitinib | 320 nm | 410 nm | 0.1–20 μg/mL | IFE | [217] |
| 30. | Doxorubicin | 491 nm | 591 nm | 1–30 μM | IFE | [218] |
| 31. | Nitazoxanide | 340 nm | 418 nm | 0.25–50.0 μM | IFE | [219] |
| 32. | Myricetin | 360 nm | 436 nm | 0.2–112 μM | IFE | [220] |
| 33. | Levocetirizine and niflumic acid | 319 nm | 392 nm | (levocetirizine − 1.0–100 μM and niflumic acid − 0.5–100 μM) | IFE | [221] |
| 34. | Vitamin B12 | 545 nm | 595 nm | 1–65 μM/70–140 μM | Internal filtration effect | [222] |
| 35. | Metronidazole | 350 nm | 425 nm | 0.5–22 μM | PET | [223] |
| 36. | Thiamine | 367 nm | 430 nm | 10–50 μM | Quenching due to Cu2+ ion and regain after addition of thiamine. (Turn on) | [224] |
| 37. | Cephalosporin | 300 nm | 689 nm | 0.2–80 μmol/L | IFE | [225] |
| 38. | Prilocaine | 400 nm | 485 nm | 2.3–400 nmol L−1 | Quenching assay/Turn off the sensor | [226] |
| 39. | Enrofloxacin | 368 nm | 452 nm | 1–15 μg/mL | The FL was suppressed by Cu2+ ions and then restored by the addition of ENR, resulting in an off-on switch of the FL. | [227] |
| 40. | Chlortetracycline | 360 nm | 438 nm | 1–70 μM | The FL of CDs is quenched by the addition of QDs, while increases linearly with the concentration of CTC resulting in an turn off–on switch of the FL | [228] |
| 41. | Glutathione | 360 nm | 430 nm | 20–400 μM | The FL of CDs was quenched with the addition of Cu2+, by the formation of CDs-Cu2+ and recovered rapidly with the addition of GSH, due to the stronger interaction between GSH and CDs-Cu2+, resulting in an turn off–on switch of the FL. | [229] |
| 42. | Pentachlorophenol (PCP) | 360 nm | 440 nm | 10–300 μM | The FL of NS-CDs was significantly quenched by the addition of H2O2 and HRP and induced FL quenching. While after the addition of PCP provided a reducing environment that can protect the active group of NS-CDs from oxidation. Thus, resulting in an turn off–on switch of the FL. | [230] |
| 43. | Ampicillin | 340 nm | 427 nm | 6.6–200 ppm | Fluorescent quenching | [231] |
| 44. | Nitrofurantoin | 340 nm | 470 nm | 0–500 μM | IFE and SQ | [232] |
| 45. | Phenobarbital | 340 nm | 410 nm | 0.4–34.5 nmol L−1 | Fluorescent quenching | [233] |
| 46. | Clonazepam | 340 nm | 430 nm | 5x10−8x10−6 M | Fluorescent quenching | [234] |
| 47. | Acetaminophen | 350 nm | – | 1–80 μM | Fluorescent quenching | [235] |
| 48. | Atorvastatin | 290 nm | 423 nm | 0.025–50 μM | SQ | [236] |
| 49. | Doxorubicin | 360 nm | 441 nm | 0.5–6.5 μM | SQ | [237] |
| 50. | Imatinib | 345 nm | 415 nm | 1.0–15.0 mg/mL | SQ | [238] |
| 51. | Curcumin | 360 nm | 440 nm | 0.339–136.0 μM | SQ | [239] |
| 52. | Tetracycline | 330 nm | 520 nm | 0–27.27 μM | SQ | [240] |
| 53. | Folic acid | 360 nm | 450 nm | 10–100 μg/mL | DQ | [241] |
| 54. | Isoniazid | 270 nm | 447 nm | 4–140 μM | IFE and SQ | [242] |
| 55. | Quercetin | 350 nm | 450 nm | 0.003–80 μmol/L | IFE and SQ | [243] |
| 56. | Tigecycline | 365 nm | 500 nm | 0.005–20 μg/mL | IFE and SQ | [244] |
| 57. | Tetracycline | 360 nm | 520 nm | 0.5–40 μM | IFE and DQ | [245] |
| 58. | Nitrotyrosine | 420 nm | 679 nm | 20–105 μM | SQ and DQ | [246] |
| 59. | p-benzoquinone | 350 nm | – | 0–25 mmol/L | DQ | [247] |
| 60. | Methotrexate | 355 nm | 450 nm | 0–50 μg/mL | Fluorescent quenching | [247] |
| 61. | Doxorubicin | 350 nm | – | 0.5–25 μg/mL | Fluorescent quenching | [247] |
| 62. | Chondroitin sulfate | 488 nm | 520 nm | 0.05–2 μg/mL | Electrostatic interaction | [248] |
| 63. | Flutamide | – | – | 0.05–590 μM | – | [249] |
| 64. | Cytochrome C | 475 nm | 530 nm | 0.5–25 μM | IFE | [210] |
| 65. | Zoledronic acid | 340 nm | 430 nm | 0.1–10 μM | ZA could not quench the FL intensity of the N-CDs without Fe3+, once ZA was added to the N-CDs-Fe3+ system, the formation of a complex between ZA and Fe3+ ions occurred and the FL of N-CDs becomes on. (Turn on) | [250] |
| 66. | Glucose | 520 nm | 582 nm | 0.5–1500 μM | SQ | [251] |
3.1. FRET-based detection mechanism
The CDs were synthesized by adding distilled deionized water to diphosphorus pentoxide (P2O5) in glacial acetic acid, conducted in a fume hood. After cooling, a dark brown solid was obtained. The FL intensity of the CDs was pH-dependent, peaking at pH 6.0 before decreasing with higher pH levels. The CDs showed high selectivity for kaempferol (KAE), which significantly quenched their FL in the presence of other substances, indicating strong sensitivity towards KAE [191]. A method was developed to produce polyethyleneimine (PEI)-functionalized blue emissive CDs (b-CDs) using citric acid and PEI via a one-step hydrothermal process. These b-CDs were used to create a fluorescent immunoassay for simultaneously detecting chloramphenicol (CAP) and amantadine (AMD) in skinless chicken breasts. The b-CDs, along with green emissive CDs (g-CDs), featured numerous amino groups and distinct FL emission peaks. The study employed haptens of CAP and AMD as energy donors in a FRET system. Two-dimensional WS2 nanosheets (NSs) were used as energy acceptors, modified with antibodies specific to CAP and AMD. This selective antigen-antibody interaction facilitated the attachment of hapten-functionalized CDs to the WS2 NSs, resulting in FL quenching due to FRET. The immunoassay demonstrated potential for rapid detection of drug residues in food [192]. The CDs were synthesized using a self-catalysis method, with FL intensity decreasing rapidly in the first minute and stabilizing after 10 min, which was chosen for further experiments. The study assessed the impact of interfering substances like metal ions, biomolecules, and co-existing compounds, testing the method's effectiveness for real sample analysis. The results showed that the CDs were highly selective for hyperin (Hyp), with only slight changes in FL intensity at low concentrations and a significant decrease as Hyp concentration increased. The method was successfully applied to quantify Hyp in real samples, including fufangmuji granules and human serum [193]. The researchers developed a selective and sensitive fluorescent probe using zein biopolymer, functionalized with 3,5-dinitrobenzoyl chloride (DNB) to create DNB-CDs for detecting sumatriptan (SUM). The DNB-CDs were synthesized via direct pyrolysis of zein without additional reagents. Using the standard addition method for analysis, the DNB-CDs showed a strong selective response to SUM, unaffected by potential interfering substances. The study suggests that these CDs can be effectively used for determining SUM levels in human samples [194]. A one-step aqueous synthesis method was used to create Zn-doped CdTe quantum dots (ZnCdTe QDs), which were combined with B-CDs to develop a ratiometric fluorescent probe for detecting 6-mercaptopurine (6-MP). The probe exhibited different FL responses from the yellow emission of ZnCdTe QDs and the blue emission of CDs when exposed to 6-MP. After adding 6-MP, the FL intensity ratio remained stable for up to 20 min, with the probe showing the strongest response at pH 8.7. Using FRET, the FL of ZnCdTe QDs was selectively quenched, while the FL of CDs remained unaffected. The probe successfully detected 6-MP in human serum, offering a rapid method for 6-MP analysis in biological samples. Fig. 7 presents a schematic illustration of ratiometric FL sensing for 6-MP [195].
Fig. 7.
Schematic illustration of ratiometric fluorescence sensing for 6-MP.
The researchers developed a sensor using N-C quantum dots (QDs) by combining a CQDs solution with sodium hydroxide, dried wheat straw powder, 2,2,6,6-tetramethylpiperidine amine, and 4,5-imidazole dicarboxylic acid. The FL intensity of the N-CQDs increased gradually at pH 7 as the gentamicin (GEN) concentration increased. The sensor's selectivity for GEN was evaluated using GEN analogs and metal ions, with metal ions causing only a slight decrease in FL intensity (around 5 %). This detection method was proven effective for accurately detecting GEN in water samples, offering a fast and simple approach with significant potential for detecting GEN residues in aqueous solutions [196]. A unique two-in-one sensor based on Au@CQDs nanocomposites (NCs) was developed for detecting gentamicin (GENTA) and kanamycin (KMC). In the synthesis, CQDs acted as reducing agents to create Au@CQDs NCs. Initially, the FL intensity of CQDs decreased due to energy transfer after the formation of Au@CQDs NCs. However, the FL intensity recovered in the presence of the antibiotics, indicating that the sensor could effectively detect both GENTA and KMC through qualitative/colorimetric and quantitative/fluorimetric methods. When tested with spiked milk and eggs, the sensor showed excellent recovery, demonstrating its potential for food safety applications [197].
3.2. IFE-based detection mechanism
In this study, CDs were synthesized using 4-hydroxybenzoic acid and ethylenediamine, with 4-nitrophenyl-α-D-glucopyranoside (NGP) serving as the substrate for the enzyme α-glucosidase. The enzyme catalyzed the release of 4-nitrophenol from NGP. Acarbose, a common anti-diabetes drug, was used as an inhibitor in the assays. The addition of acarbose inhibited α-glucosidase activity, leading to a recovery of FL intensity, as shown by an increase in FL when inhibitors were added. The sensor exhibited minimal interference from other substances, demonstrating its high selectivity and applicability for detecting α-glucosidase activity [198]. CDs were synthesized using diammonium hydrogen citrate and PEG-400 via a microwave-assisted method. The resulting CDs were evaluated for cytotoxicity and cell imaging, with their FL intensity showing a pH-dependent behavior. The study also investigated the FL-quenching effects of various metal ions, finding that Fe3+ had the most significant impact. This quenching was attributed to energy transfer processes within the excited CDs. Additionally, the CDs were successfully used for cellular imaging in BGC-823 and CT26.WT cells, showcasing their potential as optical nanoprobes for cell imaging applications due to their high FL brightness [199]. Mg-doped carbon dots (Mg-CDs) were synthesized using a simple, one-pot microwave-assisted method by dissolving citric acid in a magnesium chloride (MgCl2) solution and mixing it with ethylenediamine. The resulting Mg-CDs nanoprobes (Mg-CDs-0.1) demonstrated high selectivity and sensitivity for detecting hematin due to the strong IFE between Mg-CDs and hematin. Specificity tests revealed that Mg-CDs-0.1 showed a significant FL intensity reduction only in the presence of hematin, with no impact from other molecules or metal ions. Similar results were observed with Mg-CDs-0.5, confirming their high reproducibility and selectivity for detecting hematin in human red cell hemolysate [200].
Carbon dots (CDs) were synthesized from environmentally friendly pomegranate juice using a simple hydrothermal method and tested as sensitive indicators for detecting cefixime (CEF). The detection method relies on the interaction between palladium ions (Pd (II)) and CEF in an acidic buffer (pH 4). The excitation spectra of the synthesized CDs and the Pd (II)-CEF complex were found to overlap, leading to a decrease in FL intensity when both CEF and Pd (II) were present, enabling quantitative detection of CEF. The technique showed minimal interference from other ions and biomolecules. The method was successfully applied to analyze urine samples from a healthy individual and pharmaceutical formulations, demonstrating its potential for rapid CEF detection in real samples [201]. A rapid and efficient microwave-assisted method was used to synthesize glycerol-urea CDs (GUCDs). The ability of GUCDs to detect antibiotics in urine samples was tested by adding various compounds, including antibiotics, to GUCDs solutions. The FL signal of GUCDs remained mostly unaffected by the presence of other molecules, except for tetracycline (TC) antibiotics (TC, doxycycline, and oxytetracycline (OTC)), which significantly altered the responses. The FL intensity of GUCDs was pH-dependent, with the highest emission observed at pH below 4.0. This pH was chosen to maximize the FL signal and prevent TC degradation. As a result, a simple, low-cost method based on the reduction of GUCDs FL was developed to detect TC in urine samples, indicating the high selectivity of GUCDs for TC antibiotics [202]. In a study focused on the development of a fast FL sensor, CDs were synthesized using hawthorn, and the sensor was designed to detect chlortetracycline (CTC) in pork samples. The results showed that substances like CAP, sulfanilamide, and florfenicol did not interfere with the sensor's reaction system. However, TC antibiotics (such as TC, doxycycline, and OTC) did affect the FL response, with TC having the strongest impact. This highlighted that the N-CDs were particularly selective for detecting TC antibiotics, making them useful for detecting harmful substances in food [203]. The europium doped carbon dot (Eu-CDs) were synthesized using citric acid, melamine, and Europium (III) Nitrate Hexahydrate (Eu(NO₃)₃·6H₂O) in a one-pot hydrothermal method, with Eu serving as a TC-binding site and providing FL. Formaldehyde was added as a passivating agent to modify the nitrogen's chemical state in the Eu-CDs. The study found that the maximum FL intensity ratio for both TC and Al³⁺ occurred at pH 8.0, which was identified as the optimal pH. The Eu-CDs exhibited stable FL and color, unaffected by biological ions or other antibiotics. Notably, only TC caused a significant FL enhancement at 620 nm and a color shift from blue to red, demonstrating the Eu-CDs' high selectivity for TC detection [204]. In a separate study, B-CDs were synthesized with the help of salicylic acid, melamine, and distilled water, incorporating zinc nitrate hexahydrate and 2-methylimidazole. The CDs were then modified into a hierarchical mesoporous zeolitic imidazolate framework-8 (HZIF-8) using hydrogel as a template. The specificity of the CDs@HZIF-8 system for TC was evaluated by introducing various potential interfering substances, such as metal ions, amino acids, vitamins, and antibiotics. The results showed that the sensor maintained excellent selectivity and anti-interference capacity for TC, confirming its robustness for real-world applications [205]. A novel FL sensor was developed using nitrogen-doped carbon dots (N-CDs) embedded in zinc-based metal-organic frameworks and molecularly imprinted polymer (ZIF-8&N-CDs@MIP). This sensor, optimized at pH 7, demonstrated high sensitivity and selectivity for TC, with minimal interference from its structural analogs, such as CAP, OTC, and CTC. The sensor showed excellent selectivity, improved sensitivity, and fast response times for detecting TC [206]. N-CDs were synthesized using pitaya peel and 1,2-ethylenediamine via a hydrothermal method. The N-CDs exhibited a “turn-off” FL response when interacting with TC and OTC, while showing a “turn-on” FL response with CTC, making them highly sensitive for CTC detection. Selectivity tests demonstrated that the FL intensity of N-CDs was minimally affected by other potential interfering substances, including metal ions, antibiotics, amino acids, and anions. The sensor proved to be a versatile, dual-mode platform for detecting TC with high sensitivity and adaptability. For example, the schematic illustration of the synthetic process of N-CDs and their application as a multifunctional nano-sensor for sensing TC, OTC, and CTC is shown in Fig. 8 [207].
Fig. 8.
Schematic illustration of synthetic process of N-CDs and application as a multifunctional nano-sensor for sensing of TC, OTC, and CTC.
In this study, CDs were synthesized using trimesic acid and 4-amino acetanilide hydrochloride through a one-step hydrothermal process. This method involved combining glucose solutions of varying concentrations with glucose oxidase (GOx) in a Britton-Robinson buffer at 37 °C for 40 min to produce hydrogen peroxide (H2O2). When CDs were added to the mixture, their FL was quenched by H2O2, with the quenching intensity being directly proportional to the H2O2 concentration, enabling glucose detection. The method showed good sensitivity and selectivity when tested with biological samples, including human serum and urine from diabetic and healthy individuals. Additionally, B, N-doped CDs were synthesized using ammonium citrate and bis(pinacolato)diboron via a hydrothermal method. Interference studies revealed that nimesulide significantly quenched the FL, while other substances had little to no effect. This demonstrated the high sensitivity and selectivity of the sensor. Furthermore, the B, N-CDs proved useful for fluorescent staining and anti-counterfeit applications. The method was successfully applied to pharmaceutical samples for detecting Nimesulide [208].
N and Cl-functionalized CDs (N, Cl-CDs) were synthesized quickly and environmentally using glucose, ethylenediamine, and HCl. The FL intensity of the N, Cl-CDs was rapidly quenched by curcumin (Cur), indicating its effective interaction with the CDs. Selectivity tests showed that Cur was the only substance among the investigated compounds that significantly reduced the FL intensity, while other potential interfering substances had little to no impact. This demonstrates that N, Cl-CDs exhibit high selectivity for Cur and are unaffected by interference from other drugs, making them a promising tool for Cur detection [209]. A study developed a label-free probe for detecting cytochrome c (Cyt C) using nitrogen and fluorine co-doped carbon dots (N, F-CDs), which were easily synthesized through a solvothermal method using 3,4-difluorophenylhydrazine as a precursor. The results demonstrated that the N, F-CDs probe exhibited excellent anti-interference properties, making it a promising candidate for detecting Cyt C. Additionally, the probe showed potential as a temperature sensor with higher sensitivity, suggesting its broad application in biosensing [210].
In this study, CDs were synthesized using ammonium citrate and ammonium thiocyanate through a hydrothermal process. The addition of Cu2⁺ initially increased the FL intensity of the CDs. However, when clioquinol (CQ) was introduced, its high affinity for Cu2⁺ displaced Cu2⁺ from the CDs surface, leading to FL quenching due to IFE. The study examined the effects of pH, ionic strength, and UV exposure time on the FL intensity. The difference in FL intensity between the CDs + Cu2⁺ system and the CDs + Cu2⁺ + CQ system was greatest at pH 8.0. Cu2⁺ adsorbed on the CDs surface, stabilizing them and preventing non-radiative recombination, which increased FL intensity. A new UV absorption peak confirmed the formation of the Cu2⁺-CQ complex. Interference studies showed that CQ could significantly quench the FL of CDs, while other metal ions had negligible effects. This system was proposed as a fluorescent sensor for detecting CQ [211].
Silicon-doped carbon quantum dots (Si-CQDs) were synthesized through a straightforward one-pot hydrothermal method using 3-aminopropyltrimethoxysilane and H2SO4. The pH had no significant effect on the FL behavior of the Si-CQDs. The study also examined how the FL intensity of Si-CQDs was influenced by the addition of various metal ions, amino acids, saccharides, and other substances that could be found in urine samples. Among these, only berberine hydrochloride (BH) caused a significant reduction in FL intensity, while other drugs had minimal effects. These results demonstrate that the developed sensor has high sensitivity and selectivity for BH detection [212].
Chitosan was utilized as a C and N source in the carbonization process to develop a fluorescent sensor. The FL of the resulting probe was effectively quenched by diacerein (DIA), producing rapid and stable FL responses. The sensing system demonstrated high precision and accuracy, allowing for the selective detection of DIA in its tablet dosage form, even in the presence of co-formulated medications [213]. Rhodamine B was dissolved in ultrapure water and used in a solvothermal process to synthesize CDs. Various chemicals, including sugars (glucose, sucrose), inorganic ions (K+, Na+, Ca2+, Cl−, NO3−), amino acids (histidine, alanine, threonine, phenylalanine, arginine), and medications (paracetamol, artemisinin), were tested for their impact on the sensor. When these substances were added individually without captopril (CP), there was no significant change in the FL and absorbance signals. However, the addition of CP led to a substantial enhancement in both FL and absorbance signals. These results demonstrated the high selectivity and robustness of the CP sensor, which was able to tolerate interference from coexisting chemicals [214].
Using a one-pot green synthesis method, a green bell was chosen to produce CDs due to its high-quality contents, including carotenoids, ascorbic acid, carbohydrates, and other carbonaceous organic compounds. To create MSA-CdTe quantum dots (QDs), cadmium chloride (CdCl2), mercaptosuccinic acid (MSA), sodium tellurite (Na2TeO3), and a borate-acetic acid buffer solution were combined. Then, Si@CdTe QDs were synthesized using MSA-CdTe QDs, 3-Aminopropyl triethoxysilane (APTES), and ethanol. The ideal pH for detecting doxorubicin (DOX) and daunorubicin (DAN) was found to be pH 7. Sodium chloride (NaCl) did not cause significant changes in the signal intensity, which was used to assess the impact of potential interferents like glutamine, fructose, and glucose in biological samples. These results demonstrated that the ratiometric fluorescence (RF) sensors were highly effective in detecting both DAN and DOX [215].
Red-emission carbon dots (R-CDs) were synthesized using a solvothermal process involving formamide, N, N-dimethylformamide, and citric acid. The optimal pH for the reaction system was found to be 6.0. When used as a fluorescent probe for detecting mitoxantrone (MITX), R-CDs exhibited excellent selectivity and resistance to interference, as the presence of interference compounds had little to no effect on either the R-CDs or the R-CDs-MITX system's FL. The developed label-free fluorescent nanoprobe proved to be highly effective for the rapid and accurate detection of MITX in human serum samples [216]. The green tea leaf residue was used as the carbon source for synthesizing carbon dots (named as T-CDs) employing a combination of high-temperature pyrolysis and oxidation with concentrated H2SO4. The FL intensity of the T-CDs was found to decrease as the concentration of gefitinib increased, with the FL quenching effect being proportional to the drug concentration. Selectivity tests revealed that no other substances at a concentration of 20 μg/mL significantly affected the FL of T-CDs, except for gefitinib, indicating the high specificity of T-CDs for gefitinib detection. The method utilized the IFE of gefitinib to quench the FL of T-CDs, and it was successfully applied to detect gefitinib in urine samples [217]. A study developed a FL probe for detecting DOX using plum-based carbon quantum dots (PCQDs) synthesized via a simple bottom-up method. This ratiometric FL probe was tested on urine and serum samples, which may contain interfering substances. The results showed that, apart from DOX, the absorption peaks of interfering chemicals did not overlap with the emission peak of the PCQDs. Only DOX significantly increased the I591/I491 ratio of the PCQDs, indicating a strong response. The probe exhibited high selectivity for DOX over other drugs, such as cytarabine, cytoxan, 5-fluorouracil, and methotrexate, which had no effect on the I591/I491 readings. This demonstrated that the presence of other substances did not interfere with the detection of DOX, making the probe a reliable and selective tool for DOX detection in complex samples [218].
Plant-based sulfur and nitrogen self-co-doped carbon quantum dots (S, N-CQDs) were synthesized in an environmentally friendly, cost-effective, and rapid one-pot process using onion and cabbage juices and distilled water. The optimal pH for the procedure was determined to be 8.0. The applicability of the synthesized S, N-CQDs for FL sensing of nitazoxanide and hemoglobin (Hb) was evaluated by testing their response to various potential interfering substances, including glycine, glucose, sucrose, citric acid, sodium benzoate, sodium citrate, urea, lysine, nicotinamide, glutathione, and various ions (Ca2⁺, K⁺, Na⁺, Mg2⁺, Cl⁻, SO₄2⁻). The results demonstrated that these green, safe, affordable, and sustainable S, N-CQDs have significant potential for use in pharmaceutical and biological applications [219].
N-CDs were synthesized using diethylenetriamine and citric acid through a hydrothermal process. The FL intensity of the sensor decreased linearly with an increase in myricetin concentration. The sensor showed high selectivity for myricetin, as its signal intensity was unaffected by various structural analogs such as warfarin, KAE, and luteolin, among others. Additionally, the sensor exhibited strong anti-interference properties, with minimal impact from potential interfering ions or compounds, ensuring reliable detection of myricetin. The schematic diagram of the preparation process of N-CDs and the detection principle for myricetin is shown in Fig. 9 [220].
Fig. 9.
Schematic diagram about the preparation process of N-CDs and detection principle for myricetin.
Water-soluble carbon dots (SD-CDs) were synthesized using a microwave-assisted approach with 5-sulpha anthranilic acid (SAA) and 1,5-diphenylycarbazide (DPC) as precursors. The selectivity of SD-CDs towards levocetirizine and niflumic acid was investigated in the presence of various metal ions (Na+, Ni2+, Mn2+, Cd2+, K+, Ca2+), anions (Br−, SO42−, Cl−, NO3−), and drugs. The results showed that the FL intensity of SD-CDs was enhanced exclusively by levocetirizine, even in the presence of other interfering substances. On the other hand, niflumic acid caused a significant FL quenching when mixed with other chemicals, demonstrating a high selectivity of SD-CDs for these two compounds. The addition of metal ions, anions, and other drugs did not result in FL enhancement or quenching, confirming the robustness and specificity of SD-CDs. These findings indicate that SD-CDs could serve as a reliable analytical platform for detecting levocetirizine and niflumic acid in real-world samples [221].
Orange-emission fluorescent multifunctional carbon dots (O-CDs) were synthesized using safranine T and ethanol through a one-step hydrothermal process. These O-CDs exhibit excitation-independent FL properties, making them suitable for applications in biosensing, cellular labeling, and photovoltaic materials. When tested for selectivity, O-CDs showed significant FL quenching only in the presence of vitamin B12 (VB12), indicating their high selectivity for VB12 analysis. The O-CDs also demonstrated a longer FL emission duration compared to other fluorescent techniques for VB12 analysis, making them highly effective for detecting VB12 in biological samples. The mechanism of FL quenching was attributed to IFE, confirmed through UV–Vis absorption spectra and lifetime measurements, showing that VB12 absorbed the excitation spectrum of O-CDs, leading to FL quenching. This specificity was further validated as the FL did not change in the presence of amino acids or other vitamins [222].
3.3. PET-based detection mechanism
In a study using a high-pressure microwave method, biomass CDs were synthesized from longan peels. To enhance the selectivity and sensitivity of the fluorescent probes, CDs were integrated with molecularly imprinted polymers (MIPs) and restricted access materials (RAM-MIPs), resulting in the creation of CDs@RAM-MIPs. The results showed that CDs@RAM-MIPs exhibited excellent selectivity for metronidazole (MNZ), with no significant interference from other drugs. The unique recognition cavities of the CDs@RAM-MIPs selectively bind MNZ, allowing for precise detection of the target molecule. The method's applicability was also evaluated by measuring MNZ levels in serum samples, demonstrating its potential for accurate detection in complex biological samples [223].
3.4. On, off, and off-on based detection mechanism
Fluorescent CDs were synthesized by a microwave-assisted method using a mixture of coconut water and ethanol (1:1 v/v). The resulting CDs displayed luminescent properties, which were influenced by the reaction temperature and microwave exposure time during the synthesis. The study showed that pure coconut water itself did not exhibit any FL, but the CDs produced during the microwave reaction displayed strong FL. These CDs were then used as a sensing platform for thiamine detection. The FL of the CDs was quenched by the addition of Cu2+ ions. However, when thiamine was added to the mixture of CDs and Cu2+, the FL intensity was restored to its original value. This restoration of FL intensity was directly proportional to the concentration of thiamine, allowing for the quantification of thiamine in real samples such as blood and urine. The method proved highly selective for thiamine, with minimal interference from other potentially coexisting molecules and ions. This makes the CDs a useful tool for thiamine analysis in biological samples, demonstrating a practical and sensitive approach to thiamine detection [224].
In a separate study, N-CDs were synthesized using a hydrothermal method to detect zoledronic acid (ZA) in blood serum. The N-CDs exhibited a FL intensity change when ZA was added. Specifically, the FL was quenched when Fe3+ ions were present, but it was restored upon the introduction of ZA, forming a complex with Fe3+. The optimal pH for this FL sensing system was found to be 4.0, using an acetate buffer. This method showed high selectivity for ZA, as there were no significant FL changes when other molecules or ions, even in excess, were added. The study also compared the FL response of ZA to other substances like fludarabine (FLD), a chemotherapy drug, and found that FLD did not interfere with the sensing process. This demonstrates that the N-CDs-Fe3+ complex can be effectively used for selective and sensitive detection of ZA in blood serum, highlighting the potential of N-CDs for chemical sensing and medical applications [225].
Saffron was used as a precursor to synthesize CDs through a hydrothermal method, offering a novel and cost-effective approach. The fluorescent properties of these CDs enabled the sensitive detection of prilocaine in human plasma. Additionally, the synthesized CDs were applied for bioimaging, specifically for imaging olfactory mucosa cells and bone marrow cells. The results demonstrated excellent imaging capabilities, suggesting that these saffron-derived CDs are promising candidates for bioimaging applications, particularly in cancer cell imaging. This highlights their potential for medical and diagnostic use [226].
The researchers synthesized N-CDs using DL-malic acid and glycine. Initially, adding enrofloxacin (ENR) to the N-CDs caused no change in the FL intensity or shape, indicating no interaction between the two. However, the introduction of Cu2+ restored the FL of the N-CDs, suggesting that Cu2+ interacts with both ENR and the N-CDs. Additionally, the concentration of N-CDs in the solution increased with the addition of Cu2+, likely due to the aggregation of some N-CDs or the formation of larger complexes between Cu2+ and ENR. This approach was shown to selectively detect ENR in tap and river water samples, demonstrating its potential for environmental monitoring [227].
In a study, the FL intensity of CDs increased with rising concentrations of CTC, with a blue shift in the emission peak. This sensor showed excellent selectivity for CTC over other TC antibiotics, structurally similar drugs, various ions, and naturally occurring amino acids. The sensor was also effective in detecting CTC in milk and tap water, indicating its application in food safety and environmental testing [228].
Moreover, N, S-CDs were synthesized using a one-pot ionothermal method with deep eutectic solvent and microcrystalline cellulose (MCC) as solvents and dopants. These N, S-CDs were used to detect glutathione (GSH), where the FL quenching caused by the formation of a CD-Cu2+ complex was reversed by adding GSH. The interaction between GSH and Cu2+ restored the FL, and as GSH concentration increased, the FL became more intense. This work demonstrated that GSH is selective for the recovery of CD FL, and the study also evaluated the impact of interfering substances like amino acids and small reducing molecules. The findings suggest broad applications for biomass-derived carbon dots in selective FL sensing. The schematic illustration of the preparation of N, S-CDs and their detection capabilities for Cu2⁺, and GSH is shown in Fig. 10 [229].
Fig. 10.
Schematic preparation of N, S-CDs and its detection for Cu2+ and GSH.
Food-derived crawfish shells were utilized as green precursors to synthesize N, S-CDs via a hydrothermal method. The FL of the N, S-CDs was significantly quenched upon the addition of H2O2 and horseradish peroxidase (HRP). This quenching occurred because the hydroxyl radicals generated by the reaction between H2O2 and HRP oxidized the surface groups (N, sulfur, and oxygen) of the N, S-CDs, altering their surface states and causing FL suppression. However, when pentachlorophenol (PCP) was introduced to the NSCDs@HRP/H2O2 system, the FL was restored. PCP provided a reducing environment that prevented the oxidation of the active groups on the N, S-CDs, thus reactivating the FL. This fluorescent sensing method proved highly effective for detecting PCP in real samples, highlighting the potential of this approach for environmental and analytical applications [230].
3.5. Quenching-based detection mechanism
In one study, CDs were synthesized through hydrothermal treatment using glucose, ampicillin, 2-phenyl glycine, and (+)-6-amino penicillanic acid. The β-lactam subunit within ampicillin, with its N and S atoms, was identified as the key factor contributing to the unique absorption and emission properties of the resulting CDs. Unlike a glucose solution without hydrothermal treatment, the product solutions containing varying amounts of ampicillin exhibited distinct FL, showcasing the significant role of ampicillin in the FL properties of the CDs [231]. In a study involving co-doped CNDs containing N and phosphorus (P), the synthesis was achieved using FA through a single-step solvothermal method. The interference study showed that the FL intensity of the CNDs remained almost unchanged when exposed to common excipients like lactose, L-ascorbic acid, talc, magnesium tartrate, and starch, suggesting that these substances did not interfere with the analytical method. This indicates the reliability of the CND-based sensor. Furthermore, in vivo studies revealed that the CNDs had no significant cytotoxic, hematological, or biochemical effects on animals, supporting their biocompatibility [232].
In another study, Cedrus was used as a green source for producing CDs through a hydrothermal method. These CDs were then modified with MIPs using reverse micro-emulsion technology, resulting in MIPs-Green Source CDs (MIPs-GSCDs). The sensor was optimized in terms of pH, temperature, and response time. The MIPs-GSCDs exhibited high sensitivity and selectivity for phenobarbital in human blood plasma, demonstrating the practical application of this sensor for real sample analysis [233].
A separate study used citric acid and ammonia to synthesize CDs through pyrolysis at high temperatures. These CDs were tested for their selectivity in plasma and pharmaceutical formulations, particularly in the presence of potential interfering substances. While most compounds had little effect on the FL of CDs, clonazepam caused a significant quenching of FL, likely due to interactions between its nitro group and the surface functional groups of the CDs. This high selectivity allows for the precise detection of clonazepam in low concentrations, making the sensor valuable for monitoring trace amounts in pharmaceutical and plasma samples [234].
Additionally, red-emitting copper nanoclusters (CuNCs) were synthesized using DNA as a template and combined with blue-emitting CDs to form a self-assembled complex, DNA-CuNC/CDs. This complex was tested as a ratiometric FL nanoplatform for detecting arginine (Arg) and acetaminophen (AP) in human serum samples. Upon the addition of AP, the FL of the complex shifted from purplish-red to blue, and the FL of the CDs was restored, demonstrating the ability to detect AP. The system showed excellent selectivity and sensitivity, with minimal interference from other substances. This ratiometric FL nanoplatform also highlighted the potential for building “INHIBIT” logic gates at the molecular level, offering exciting prospects for early disease diagnosis and real-time clinical monitoring [235].
A microwave-assisted method was used to synthesize CDs from glucose and deep eutectic solvents (DESs), composed of choline chloride and urea mixed with deionized water. The CDs were tested for FL properties by mixing them with various ions and molecules at a concentration of 3 μM. The FL intensity remained mostly unaffected by most species, except for atorvastatin, which caused significant quenching, indicating the sensor's high selectivity for atorvastatin. FL quenching occurred quickly at room temperature after adding atorvastatin. The FL intensity was pH-dependent, increasing between pH 3–8, with a peak at pH 7.4, which was chosen as the optimum pH for the sensor. The FL intensity decreased significantly with increasing atorvastatin concentration, demonstrating the CDs sensitivity for atorvastatin detection. Nitrogen and chloride-doped carbon dots (N/Cl-CDs) were synthesized using the same method to enhance the sensor's selectivity and sensitivity. These N/Cl-CDs were successfully used as a fluorescent nanosensor to detect atorvastatin in blood, showcasing their potential for monitoring atorvastatin in biological samples [236].
Hydrothermal synthesis was used to create nitrogen and phosphorus co-doped carbon dots (N, P-CDs) with intense blue FL using alanine and diammonium phosphate. The FL intensity remained stable with an increase in pH until DOX was introduced, which caused a significant decrease in the FL intensity, indicating that the N, P-CDs were sensitive to DOX. The FL intensity reached its peak at pH 7.0, and with increasing concentrations of DOX, the FL intensity gradually reduced. When tested against various common metal cations and biomolecules, only DOX caused a marked “turn-off” effect, demonstrating the high selectivity of the N, P-CDs for DOX detection [237]. Thiosemicarbazide and citric acid were utilized in a one-step hydrothermal method to synthesize nitrogen and sulfur-doped carbon quantum dots (N, S-CQDs). These N, S-CQDs demonstrated strong FL and were used for the detection of imatinib (IMA) in pharmaceutical and biological samples. As the concentration of IMA increased, the FL intensity of the N, S-CQDs decreased, providing a simple and sensitive approach for IMA detection, with the added benefits of low cost and ease of use. The systematic detection process of imatinib is depicted in Fig. 11 [238].
Fig. 11.
Outline of the synthesis process and applications of N,S-CQDs.
Similarly, a fluorescent sensor for Cur detection was developed based on N, S-CQDs and MIPs. The CQDs with blue FL were synthesized using citric acid and o-phenylenediamine. The sensor, composed of N, S-CQD@MIP, showed higher quenching efficiency for Cur compared to the N, S-CQD@NIP composite, highlighting its superior sensitivity. Even when the concentration of interfering drugs was ten times higher than Cur, the FL quenching efficiency remained largely unaffected, demonstrating excellent selectivity for Cur. Further tests with various ions and proteins confirmed that the sensor was highly selective and specific for Cur detection [239]. Highly luminous CDs were synthesized from wild lemon leaves using one-step microwave pyrolysis. The selectivity of the CDs was tested by introducing a variety of foreign substances, including biomolecules, metal ions, and antibiotics. The CDs demonstrated exceptional selectivity for TC detection, as they remained highly fluorescent even in the presence of various interferents. This biocompatible, label-free nanoprobe was successfully used to detect TC in environmental water samples, showing good recovery rates and promising potential for environmental monitoring [240].
Researchers synthesized Aconitic acid-based carbon dots (AA-CDs) using aconitic acid (AA) and 1,2-ethylenediamine via a hydrothermal method. They found that the addition of FA could effectively quench the intrinsic FL of AA-CDs without requiring additional surface modifications or passivation. The AA-CDs were successfully used for detecting FA in food and pharmaceutical samples. Additionally, the researchers demonstrated the potential of FA-AA-CDs for targeted imaging, as they were able to distinguish cancer cells (HeLa, SMMC-7721, and A549) based on varying levels of folate receptor's (FRs) expression. This showed the feasibility of using turn-on FL for imaging cancer cells with overexpressed FRs indicating a promising application for cancer diagnostics and imaging. Fig. 12 presents a schematic illustration of ratiometric FL sensing for FA [241].
Fig. 12.
Schematic illustration for the detection of folic acid.
A new type of fluorescent CDs was synthesized by pyrolyzing folic acid. These fluorescent CDs exhibit excitation-independent FL, allowing for rapid and sensitive detection of isoniazid through a combination of IFE and SQ effects. This makes them promising for drug testing applications, potentially offering new methods and technologies for detecting pharmaceutical compounds [242]. Another study focused on the synthesis of fluorescent nitrogen-doped carbon quantum dots (N-CQDs) using a hydrothermal process involving natural osmanthus fragrans, without the use of harmful substances or surface chemical modifications. In tests for potential interference, the FL intensity of N-CQDs remained stable when mixed with various metal cations, indicating that they are highly selective for quercetin (QT) sensing. These N-CQDs can serve as a reversible, sensitive platform for detecting QT, making them suitable for clinical diagnostics and other biomedical applications [243].
Additionally, Fluorine, Nitrogen-doped carbon dots (F, N-CDs) were synthesized using tetrafluoroterephthalic acid and tetraethylenepentamine. These F, N-CDs exhibited a rapid decrease in FL intensity as the concentration of tigecycline (TIGE) increased. The FL quenching effect was highly specific, as other potential interferences showed minimal impact. The method demonstrated strong anti-interference properties, making it suitable for real-time drug monitoring. The approach was successfully applied for detecting labeled TIGE and TIGE injections in human plasma, highlighting its potential for precise and reliable drug monitoring in clinical settings [244].
A sensor was developed by coupling ovalbumin (OVA) and 3-aminophenyl boronic acid (3-APBA) with CDs through the specific interaction of cis-diol bonds in glycoproteins and borate groups on the CDs surface. This CDs-functionalized OVA acted as both a stabilizing and reducing agent during the synthesis of gold nanoclusters (AuNCs), forming a CDs-AuNCs nanocomposite. The sensor demonstrated the highest ratiometric FL efficiency at pH 8.5, producing a bright yellow color. The sensor was tested for TC detection, showing high selectivity and minimal interference from other antibiotics, amino acids, and potential contaminants. This triple-emission sensor is highly selective for tetracyclines (TC's) and has potential for practical applications in antibiotic detection [245]. In a single step, CDs were synthesized using GSH and formamide via a microwave-mediated process. Interference studies demonstrated that the CDs were highly selective for detecting nitrotyrosine (nTyr), even in the presence of other biologically significant compounds like tyrosine (Tyr) and phosphotyrosine (pTyr). Without nTyr, the FL emission of the CDs was only slightly reduced (by 10 %) in the presence of these interferents, indicating that the CDs were specifically designed for nTyr detection with minimal interference [246].
The CDs were synthesized using a hydrothermal method with citric acid as the carbon source and ethylenediamine as the co-reactant. These CDs were employed as probes to detect dopamine (DA) and DOX in human serum solutions. In the serum, DA effectively quenched the FL of the CDs, establishing a strong linear correlation. Similarly, DOX was also detected with a good linear relationship in the serum solution. The results indicate that CDs are highly suitable for detecting DA and quinone-based drugs, such as DOX, in human serum due to their excellent stability, eco-friendliness, and anti-interference properties. The method was also successfully applied to quantitatively identify DOX and MITX, both of which share a typical quinone structure. This approach provides a solid foundation for further biological applications, offering reliable and efficient detection capabilities in clinical and pharmaceutical settings [247].
Fluorescent CDs were synthesized using citric acid and L-glutathione to create a platform for detecting chondroitin sulfate (CS). The positively charged N-doped CDs (P-NCDs) were combined with FAM-labeled random-sequence single-stranded DNA (F-ssDNA), creating a sensitive and simple ratiometric homogeneous assay through competitive electrostatic interactions. The process of CS detection was confirmed by measuring the zeta potential, which decreased due to the electrostatic interaction between negatively charged CS and P-NCDs/F-ssDNA. The zeta potential decreased significantly when CS was added, showing a stronger interaction between P-NCDs and CS (−12.5 mV) compared to P-NCDs and F-ssDNA. Further validation was achieved by comparing the zeta potential of P-NCDs with CS (−4.5 mV) to that of their mixture (−19.75 mV). The system was successfully applied to detect CS in actual joint fluid, demonstrating its potential for clinical use. This study highlights new insights into the development of fluorescent nanomaterials and DNA for biosensing applications [248].
CDs were synthesized using CO(NO3)2.6H2O and cetrimonium bromide (CTAB), followed by dispersion with multi-walled carbon nanotubes (MWCNTs) via ultrasonication. The FL intensity of the CDs was found to be pH-dependent, with the highest redox peak current observed at pH 7.0 for the drug molecules flutamide (FLU) and nitrofurantoin (NF). The sensor was evaluated for practical use in human urine samples, showing its capability for the simultaneous detection of these drug molecules. This study demonstrates a straightforward method for synthesizing a hybrid nanocomposite and highlights its potential for drug detection applications [249].
4. Current challenges and future prospects
Many pharmaceutical drugs have been detected using HPLC, HPTLC, UV–Vis, and Fourier transform infrared spectroscopy (FT-IR) to date. Analytical experiments (sampling, sample preparation, instrumental analysis, data processing, and data interpretation) present distinct problems in achieving the goal of enhancing the existing state of pharmaceutical analysis and, more broadly, functionalization. To overcome these challenges CDs can be used for detection just by converting the non-fluorescent substances into fluorescent substances and sensing through a different mechanism. Pharmaceutical residues of antibiotics, nonsteroidal anti-inflammatory drugs (NSAIDs), lipid-regulators, hormones, β-blocker, anticonvulsants, steroids, opiate drugs, antidepressants, and anti-diabetics in the environment have recently been discovered as a hazard, with their presence in the aquatic environment being particularly critical [252]. Experiments conducted in both laboratory and natural settings have revealed that oral contraceptives are leaving traces in water bodies and causing the feminization of fish and amphibians, while psychiatric drugs are leaving residues that alter fish behavior. Furthermore, the excessive use of antibiotics has led to a rapid rise in antimicrobial resistance, which has become a major global health crisis [253]. Previous studies have found that common antidepressant medications are accumulated in the brain tissue of fish that inhabit water downstream from wastewater treatment plants (WWTPs) [254]. Numerous techniques have been proposed for eliminating antibiotics, including advanced oxidation, reverse osmosis, membrane filtration, electrochemical methods, and biological treatments. However, most of these methods are expensive, produce by-products, or are not as efficient. Removal of antibiotics by using various adsorbents (sawdust, activated carbon, NPs, sludge biochar) has been reported. Antibiotics such as gatifloxacin have been removed by more than 90 % using NPs and 68.5 % and 64 % of ciprofloxacin by using kandira stone and sawdust as an adsorbent [255]. The Fenton oxidation process combined with ultrafiltration (UF), sand filtration, reverse osmosis (RO), and nanofiltration (NF) resulted in recovery above 90 %. In another study, the electro-Fenton process was used to remove diclofenac sodium where about 80 % drug removal was observed [256]. Recently, various techniques have been developed to remove antibiotics and antibiotic resistance genes. These techniques include the use of Fe3O4/red mud NPs, 3D hierarchical porous-structured biochar aerogels, calcined layered double hydroxides, co-doped UiO-66 NPs, Cu@TiO2 hybrids, bioelectrochemical systems, and aerobic granulation process. The results of studies have shown that most of these methods are effective in removing antibiotic residues and antibiotic resistance genes, with removal rates ranging from 85 % to 95 % [257]. A recent study found that a mix of domestic and livestock sewage in rural wastewater contained five steroid hormones [androsta-1,4-diene-3,17-dione (ADD), 4-androstene-3,17-dione, 19-norethindrone (19-NTD), testosterone (T), and progesterone ] and four biocides [N, N-diethyl-3-methylbenzamide (DEET), triclosan (TCS), carbendazim (CBD), and methylparaben (MP)]. However, the results showed that an integrated constructed wetland (ICW) was able to significantly reduce the levels of the detected hormones and biocides, with a reduction rate of 97.4 ± 0.09 % and 92.4 ± 0.54 %, respectively [258].
Due to the physical and chemical complexity of the NSAIDs compounds, there is no single method that is sufficiently effective against all types of contaminants [259]. Evidence reveals that pharmaceutical chemicals, physicochemical properties, and structural complexity make them difficult to remove completely in traditional WWTPs, highlighting their unintentional persistence in the environment. There is a toxic effect of pharmaceutical compounds on the human body at low concentrations. Hence, a more sensitive method is required to detect the pharmaceutical compounds at low concentrations, and for this; CDs are the best alternative method to detect the pharmaceutical compounds at the femtogram level. Nowadays, nanomaterials can be used for the treatment of wastewater and purification. Materials such as graphene oxide-based NPs (GONPs), mesoporous MnxCo3-xO4 NPs, etc., have been used for the removal of ibuprofen, ciprofloxacin, and levofloxacin from water and aqueous solution by the mechanism of adsorption mainly due to electrostatic interactions, accelerated electron transfer and by photocatalytic degradation. The percentage removal of ibuprofen, ciprofloxacin, and levofloxacin was found to be 98.2 %, 100 %, and 95 %, respectively [[260], [261], [262]]. CDs with an optimum size of 5 nm–10 nm and modifying the surface functionalization of CDs can be used for various applications apart from bio-sensing. It includes tumor theranostics (cancer-targeted drug delivery, gene delivery, PTT, photodynamic therapy (PDT)), bio-imaging biomarkers (cell membrane, cytoplasm, for self-targeted bioimaging and diagnosis of tumor cells), catalysis, detection of small molecules, photoacoustic imaging [263], light-emitting diode (LED) device [264], FL ink [265], in agriculture [266], etc.
5. Concluding remark
This review highlights the recent advancements in CDs, focusing on their synthesis, surface functionalization, photoluminescent properties, and diverse applications across fields such as photocatalysis, energy, and sensing. CDs have emerged as promising alternatives to traditional fluorescent compounds, particularly in in-vivo analysis, due to their safer nature and exceptional FL properties. The unique physical and optical characteristics of CDs, including their efficient electron storage and transport capabilities when exposed to light, present vast untapped potential. Looking forward, the development of more reliable and efficient synthesis methods will likely drive further exploration of innovative applications. The versatility of CDs positions them as impactful materials in biotechnology and environmental remediation, offering safer substitutes for conventional analytical techniques. Given the complexity of contaminants that conventional wastewater treatment plants cannot fully eliminate, there is an increasing need for advanced materials like CDs to address these challenges. With their exceptional electron transfer abilities and light-harvesting efficiency, CDs hold significant promise in photovoltaic and photocatalytic applications. Additionally, their potential extends to areas such as bio-sensing, tumor theranostics, bio-imaging, LED technology, small molecule detection, and agriculture. The continued research and development of CDs are expected to unlock new applications, broadening their utility in various scientific and industrial fields.
CRediT authorship contribution statement
Sandesh R. Lodha: Writing – original draft, Visualization, Conceptualization. Jesika G. Merchant: Writing – original draft. Arya J. Pillai: Writing – original draft, Visualization. Anil H. Gore: Writing – review & editing. Pravin O. Patil: Writing – review & editing, Writing – original draft. Sopan N. Nangare: Writing – review & editing. Gajanan G. Kalyankar: Writing – review & editing. Shailesh A. Shah: Supervision, Resources. Dinesh R. Shah: Supervision, Resources. Shashikant P. Patole: Writing – review & editing, Supervision, Funding acquisition.
Declaration of competing interest
The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
Acknowledgments
Sandesh Lodha would like to thanks Maliba Pharmacy College, Uka Tarsadia University for providing resource facilities. Shashikant P. Patole would like to thanks Khalifa University of Science and Technology for supporting funds for high quality publications.
Contributor Information
Sandesh R. Lodha, Email: sandeshlodha@gmail.com.
Anil H. Gore, Email: anilanachem@gmail.com.
Shashikant P. Patole, Email: Shashikant.Patole@ku.ac.ae.
References
- 1.Sanvicens N., et al. Biosensors for pharmaceuticals based on novel technology. Trends Anal. Chem. 2011;30(3):541–553. [Google Scholar]
- 2.Khan H.K., Rehman M.Y.A., Malik R.N. Fate and toxicity of pharmaceuticals in water environment: an insight on their occurrence in South Asia. J. Environ. Manag. 2020;271 doi: 10.1016/j.jenvman.2020.111030. [DOI] [PubMed] [Google Scholar]
- 3.Ginebreda A., Barceló D., Rodríguez-Mozaz S. In: Removal and Degradation of Pharmaceutically Active Compounds in Wastewater Treatment. Rodriguez-Mozaz S., Blánquez Cano P., Sarrà Adroguer M., editors. Springer International Publishing; Cham: 2021. Environmental risk assessment of pharmaceuticals in wastewater treatment; pp. 1–21. [Google Scholar]
- 4.Buledi J., et al. Current perspective and developments in electrochemical sensors modified with nanomaterials for environmental and pharmaceutical analysis. Curr. Anal. Chem. 2022;18(1):102–115. [Google Scholar]
- 5.Joshi D.R., Adhikari N. An overview on common organic solvents and their toxicity. J. Pharma. Res. Int. 2019:1–18. [Google Scholar]
- 6.Alfonsi K., et al. Green chemistry tools to influence a medicinal chemistry and research chemistry based organisation. Green Chem. 2008;10(1):31–36. [Google Scholar]
- 7.Bottoni P., Caroli S. Fake pharmaceuticals: a review of current analytical approaches. Microchem. J. 2019;149 [Google Scholar]
- 8.Menard K.P. vol. 13. 2013. (Impurities Detection - Impurities Detection in Pharmaceuticals). [Google Scholar]
- 9.Wertheimer A.I., Chaney N.M., Santella T. Counterfeit pharmaceuticals: current status and future projections. J. Am. Pharmaceut. Assoc. 2003;43(6):710–717. doi: 10.1331/154434503322642642. quiz 717-717. [DOI] [PubMed] [Google Scholar]
- 10.Beckett Ah S.J. A&C Black; 1988. Practical Pharmaceutical Chemistry. Fourth Ed. Vol. Part II. [Google Scholar]
- 11.Chatwal G.R. Himalaya publishing house; 1979. A.S., Instrumental Methods of Chemical Analysis:(for Hons. And Post-graduate Students of Indian and Foreign Universities) [Google Scholar]
- 12.Jm T. Jenny Stanford Publishing; 2018. Infrared Spectroscopy. [Google Scholar]
- 13.Pavia D.L.L.G., Kriz G.S., Vyvyan J.A. Cengage learning; 2014. Introduction to Spectroscopy. [Google Scholar]
- 14.Skoog Da H.F., Crouch S.R. Cengage learning; 2017. Principles of Instrumental Analysis. [Google Scholar]
- 15.Majhi K.C., Yadav M. Green Sustainable Process for Chemical and Environmental Engineering and Science, Inamuddin, et al. Elsevier; 2021. Chapter 5 - synthesis of inorganic nanomaterials using carbohydrates; pp. 109–135. [Google Scholar]
- 16.Porto L.S., et al. Carbon nanomaterials: synthesis and applications to development of electrochemical sensors in determination of drugs and compounds of clinical interest. Rev. Anal. Chem. 2019;38(3) [Google Scholar]
- 17.Qian L., et al. Nanomaterial-based electrochemical sensors and biosensors for the detection of pharmaceutical compounds. Biosens. Bioelectron. 2021;175 doi: 10.1016/j.bios.2020.112836. [DOI] [PubMed] [Google Scholar]
- 18.Mostofizadeh A., et al. Synthesis, properties, and applications of low-dimensional carbon-related nanomaterials. J. Nanomater. 2011;2011 [Google Scholar]
- 19.Hwang H.S., et al. Carbon nanomaterials as versatile platforms for biosensing applications. Micromachines. 2020;11(9):814. doi: 10.3390/mi11090814. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Scida K., et al. Recent applications of carbon-based nanomaterials in analytical chemistry: critical review. Anal. Chim. Acta. 2011;691(1–2):6–17. doi: 10.1016/j.aca.2011.02.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Blanco-López M.C., Garcia Carlos D., Crevillén Agustin G., Escarpa Alberto, editors. Carbon-based nanomaterials in analytical chemistryAnal. Bioanal. Chem. 2019;411(15):3219–3220. doi: 10.1007/s00216-019-01785-3. [DOI] [PubMed] [Google Scholar]
- 22.Hassanpour S., et al. Carbon based nanomaterials for the detection of narrow therapeutic index pharmaceuticals. Talanta. 2021;221 doi: 10.1016/j.talanta.2020.121610. [DOI] [PubMed] [Google Scholar]
- 23.Singh R., Lillard J.W. Nanoparticle-based targeted drug delivery. Exp. Mol. Pathol. 2009;86(3):215–223. doi: 10.1016/j.yexmp.2008.12.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Liu Y., et al. Development of high‐drug‐loading nanoparticles. ChemPlusChem. 2020;85 doi: 10.1002/cplu.202000496. [DOI] [PubMed] [Google Scholar]
- 25.Qiao W., et al. Cancer therapy based on nanomaterials and nanocarrier systems. J. Nanomater. 2010;2010 [Google Scholar]
- 26.Katubi K., et al. Synthesis of manganese ferrite/graphene oxide magnetic nanocomposite for pollutants removal from water. Processes. 2021;9:589. [Google Scholar]
- 27.Sharma D., Hussain C.M. Smart nanomaterials in pharmaceutical analysis. Arab. J. Chem. 2020;13(1):3319–3343. [Google Scholar]
- 28.Xu X., et al. Electrophoretic analysis and purification of fluorescent single-walled carbon nanotube fragments. J. Am. Chem. Soc. 2004;126:12736–12737. doi: 10.1021/ja040082h. [DOI] [PubMed] [Google Scholar]
- 29.Shafi A., et al. Carbon-Based Material for Environmental Protection and Remediation. 2020. Eco-friendly fluorescent carbon nanodots: characteristics and potential applications. IntechOpen. [Google Scholar]
- 30.Wang Y., Hu A. Carbon quantum dots: synthesis, properties and applications. J. Mater. Chem. C. 2014;2:6921–6939. [Google Scholar]
- 31.Sabet M., Mahdavi K. Green synthesis of high photoluminescence nitrogen-doped carbon quantum dots from grass via a simple hydrothermal method for removing organic and inorganic water pollution. Appl. Surf. Sci. 2019;463:283–291. [Google Scholar]
- 32.Atchudan R., et al. Facile green synthesis of nitrogen-doped carbon dots using Chionanthus retusus fruit extract and investigation of their suitability for metal ion sensing and biological applications. Sensor. Actuator. B Chem. 2017;246:497–509. [Google Scholar]
- 33.Bhatt S., et al. Green route for synthesis of multifunctional fluorescent carbon dots from Tulsi leaves and its application as Cr(VI) sensors, bio-imaging and patterning agents. Colloids Surf. B Biointerfaces. 2018;167:126–133. doi: 10.1016/j.colsurfb.2018.04.008. [DOI] [PubMed] [Google Scholar]
- 34.Liu W., et al. Green synthesis of carbon dots from rose-heart radish and application for Fe3+ detection and cell imaging. Sensor. Actuator. B Chem. 2017;241:190–198. [Google Scholar]
- 35.Sciortino A., Cannizzo A., Messina F. Carbon Nanodots: A Review—From the Current Understanding of the Fundamental Photophysics to the Full Control of the Optical Response. 2018;4:67. [Google Scholar]
- 36.Sarkar S., et al. Size dependent photoluminescence property of hydrothermally synthesized crystalline carbon quantum dots. J. Lumin. 2016;178:314–323. [Google Scholar]
- 37.Dong Y., Cai J., Chi Y. 2016. Carbon Based Dots and Their Luminescent Properties and Analytical Applications. [Google Scholar]
- 38.Khan S., et al. Recent advances in carbon nanodots: a promising nanomaterial for biomedical applications. Int. J. Mol. Sci. 2021;22:6786. doi: 10.3390/ijms22136786. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Gayen B., Palchoudhury S., Chowdhury J.N. 2019. Carbon dots: a mystic star in the world of nanoscience; pp. 1–19. [Google Scholar]
- 40.Mansuriya B., Altintas Z. Carbon dots: classification, properties, synthesis, characterization, and applications in health care—an updated review (2018–2021) Nanomaterials. 2021;11:2525. doi: 10.3390/nano11102525. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Arcudi F., Đorđević L., Prato M.J.A.C. Synthesis, separation, and characterization of small and highly fluorescent nitrogen-doped carbon NanoDots. 2016;55(6):2107–2112. doi: 10.1002/anie.201510158. [DOI] [PubMed] [Google Scholar]
- 42.Zhu S., et al. The photoluminescence mechanism in carbon dots (graphene quantum dots, carbon nanodots, and polymer dots): current state and future perspective. Nano Res. 2015;8(2):355–381. [Google Scholar]
- 43.Javed N., O'Carroll D.M.J.P., Characterization P.S. vol. 38. 2021. (Carbon Dots and Stability of Their Optical Properties). [Google Scholar]
- 44.Lu S., et al. PH-dependent synthesis of novel structure-controllable polymer-carbon NanoDots with high acidophilic luminescence and super carbon dots assembly for white light-emitting diodes. ACS Appl. Mater. Interfaces. 2016;8 doi: 10.1021/acsami.5b11579. [DOI] [PubMed] [Google Scholar]
- 45.Wu Z.L., Liu Z., Yuan Y.H. Carbon dots: materials, synthesis, properties and approaches to long-wavelength and multicolor emission. 2017;5(21):3794–3809. doi: 10.1039/c7tb00363c. [DOI] [PubMed] [Google Scholar]
- 46.Cui L., et al. Carbon dots: synthesis, properties and applications. Nanomaterials. 2021;11:3419. doi: 10.3390/nano11123419. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Bhartiya P., et al. Carbon dots : chemistry, properties and applications. J. Indian Chem. Soc. 2016;93:759–766. [Google Scholar]
- 48.Zhu A., et al. A two-photon "turn-on" fluorescent probe based on carbon nanodots for imaging and selective biosensing of hydrogen sulfide in live cells and tissues. Analyst. 2014;139 doi: 10.1039/c3an02086j. [DOI] [PubMed] [Google Scholar]
- 49.Miao P., et al. Recent advances in carbon nanodots: synthesis, properties and biomedical applications. Nanoscale. 2015;7(5):1586–1595. doi: 10.1039/c4nr05712k. [DOI] [PubMed] [Google Scholar]
- 50.Reyes D., et al. Laser ablated carbon nanodots for light emission. Nanoscale Res. Lett. 2016;11(1):424. doi: 10.1186/s11671-016-1638-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Xu H., et al. One-step synthesis of nitrogen-doped carbon nanodots for ratiometric pH sensing by femtosecond laser ablation method. Appl. Surf. Sci. 2017;414:238–243. [Google Scholar]
- 52.Xu Q., et al. Heteroatom-doped carbon dots: synthesis, characterization, properties, photoluminescence mechanism and biological applications. J. Mater. Chem. B. 2016;4(45):7204–7219. doi: 10.1039/c6tb02131j. [DOI] [PubMed] [Google Scholar]
- 53.Goryacheva I.Y., Sapelkin A.V., Sukhorukov G.B. Carbon nanodots: mechanisms of photoluminescence and principles of application. TrAC, Trends Anal. Chem. 2017;90:27–37. [Google Scholar]
- 54.Han Z., et al. Hydrothermal synthesis of carbon dots and their application for detection of chlorogenic acid. Luminescence. 2020;35 doi: 10.1002/bio.3803. [DOI] [PubMed] [Google Scholar]
- 55.Skrabalak S.E. Ultrasound-assisted synthesis of carbon materials. Phys. Chem. Chem. Phys. : Phys. Chem. Chem. Phys. 2009;11 25:4930–4942. doi: 10.1039/b823408f. [DOI] [PubMed] [Google Scholar]
- 56.Liu H., et al. Microwave-assisted synthesis of wavelength-tunable photoluminescent carbon nanodots and their potential applications. ACS Appl. Mater. Interfaces. 2015;7(8):4913–4920. doi: 10.1021/am508994w. [DOI] [PubMed] [Google Scholar]
- 57.Liu Z., He W., Guo Z. Metal coordination in photoluminescent sensing. Chem. Soc. Rev. 2013;42(4):1568–1600. doi: 10.1039/c2cs35363f. [DOI] [PubMed] [Google Scholar]
- 58.Hussain S.A., et al. Fluorescence resonance energy transfer (FRET) sensor. J. Spectros. Dyn. 2014 [Google Scholar]
- 59.Molaei M. Principles, mechanisms, and application of carbon quantum dots in sensors: a review. Anal. Methods. 2020;12 [Google Scholar]
- 60.Wang T., Zeng L.-H., Li D.-L. A review on the methods for correcting the fluorescence inner-filter effect of fluorescence spectrum. Appl. Spectrosc. Rev. 2017;52(10):883–908. [Google Scholar]
- 61.Sun W., et al. Activity-based sensing and theranostic probes based on photoinduced electron transfer. Acc. Chem. Res. 2019;52(10):2818–2831. doi: 10.1021/acs.accounts.9b00340. [DOI] [PubMed] [Google Scholar]
- 62.Wang X., et al. Photoinduced electron transfers with carbon dots. Chem. Commun. 2009;(25):3774–3776. doi: 10.1039/b906252a. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Hu J., et al. Mechanisms for carbon dots-based chemosensing, biosensing, and bioimaging: a review. Anal. Chim. Acta. 2022;1209 doi: 10.1016/j.aca.2021.338885. [DOI] [PubMed] [Google Scholar]
- 64.Misra R., Bhattacharyya S. Intramolecular charge transfer: theory and applications. Front matter. 2018 [Google Scholar]
- 65.Zu F., et al. The quenching of the fluorescence of carbon dots: a review on mechanisms and applications. Microchim. Acta. 2017;184(7):1899–1914. [Google Scholar]
- 66.Liu Y., et al. Nitrogen doped graphene quantum dots as a fluorescent probe for mercury(II) ions. Microchim. Acta. 2019;186 doi: 10.1007/s00604-019-3249-4. [DOI] [PubMed] [Google Scholar]
- 67.Li X., et al. Advances and perspectives in carbon dot-based fluorescent probes: mechanism, and application. Coord. Chem. Rev. 2021;431 [Google Scholar]
- 68.Cayuela A., et al. One-step synthesis and characterization of N-doped carbon nanodots for sensing in organic media. Anal. Chem. 2016;88(6):3178–3185. doi: 10.1021/acs.analchem.5b04523. [DOI] [PubMed] [Google Scholar]
- 69.Wang L., et al. High-yield synthesis of strong photoluminescent N-doped carbon nanodots derived from hydrosoluble chitosan for mercury ion sensing via smartphone APP. Biosens. Bioelectron. 2016;79:1–8. doi: 10.1016/j.bios.2015.11.085. [DOI] [PubMed] [Google Scholar]
- 70.Liu R., et al. Shrimp-shell derived carbon nanodots as carbon and nitrogen source to fabricate three-dimensional N-doped porous carbon electrocatalyst for oxygen reduction reaction. Phys. Chem. Chem. Phys. 2016;18 doi: 10.1039/c5cp06970j. [DOI] [PubMed] [Google Scholar]
- 71.Gao Z.-h., et al. A fluorescent probe based on N-doped carbon dots for highly sensitive detection of Hg2+ in aqueous solutions. Anal. Methods. 2016;8(10):2297–2304. [Google Scholar]
- 72.Zhang H., et al. Fluorescence determination of nitrite in water using prawn-shell derived nitrogen-doped carbon nanodots as fluorophores. ACS Sens. 2016;1 [Google Scholar]
- 73.Shi L., et al. Green-fluorescent nitrogen-doped carbon nanodots for biological imaging and paper-based sensing. Anal. Methods. 2017;9(14):2197–2204. [Google Scholar]
- 74.Zhang H., et al. Nitrogen-doped carbon Nanodots@Nanospheres as an efficient electrocatalyst for oxygen reduction reaction. Electrochim. Acta. 2015;165:7–13. [Google Scholar]
- 75.Campos B.B., et al. Carbon dots on based folic acid coated with PAMAM dendrimer as platform for Pt(IV) detection. J. Colloid Interface Sci. 2016;465:165–173. doi: 10.1016/j.jcis.2015.11.059. [DOI] [PubMed] [Google Scholar]
- 76.Chin S.-F., et al. Nitrogen doped carbon nanodots as fluorescent probes for selective detection and quantification of Ferric(III) ions. Opt. Mater. 2017;73:77–82. [Google Scholar]
- 77.Chen X., et al. Multifunctional sensing applications of biocompatible N-doped carbon dots as pH and Fe3+ sensors. Microchem. J. 2019;149 [Google Scholar]
- 78.Sekar A., Yadav R., Easwaramoorthy D. N-Doped fluorescent carbon nanodots derived out of Gum ghatti for the fluorescence tracking of mercury ions Hg2+ in the aqueous phase. Mater. Today: Proc. 2021 [Google Scholar]
- 79.Arul V., Chandrasekaran P., Sethuraman M.G. Reduction of Congo red using nitrogen doped fluorescent carbon nanodots obtained from sprout extract of Borassus flabellifer. Chem. Phys. Lett. 2020;754 [Google Scholar]
- 80.Saravanan A., et al. Applications of N-doped carbon dots as antimicrobial agents, antibiotic carriers, and selective fluorescent probes for nitro explosives. ACS Appl. Bio Mater. 2020;3(11):8023–8031. doi: 10.1021/acsabm.0c01104. [DOI] [PubMed] [Google Scholar]
- 81.Li L., et al. A novel electrochemiluminescence sensor based on Ru(bpy)32+/N-doped carbon nanodots system for the detection of bisphenol A. Anal. Chim. Acta. 2015;895:104–111. doi: 10.1016/j.aca.2015.08.055. [DOI] [PubMed] [Google Scholar]
- 82.Omer K.M., Sartin M. Dual-mode colorimetric and fluorometric probe for ferric ion detection using N-doped carbon dots prepared via hydrothermal synthesis followed by microwave irradiation. Opt. Mater. 2019;94:330–336. [Google Scholar]
- 83.Krishnaiah P., et al. 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 doi: 10.1016/j.chemosphere.2021.131764. [DOI] [PubMed] [Google Scholar]
- 84.Moniruzzaman M., Kim J. N-doped carbon dots with tunable emission for multifaceted application: solvatochromism, moisture sensing, pH sensing, and solid state multicolor lighting. Sensor. Actuator. B Chem. 2019;295:12–21. [Google Scholar]
- 85.Jia J., et al. Highly luminescent N-doped carbon dots from black soya beans for free radical scavenging, Fe3+ sensing and cellular imaging. Spectrochim. Acta Mol. Biomol. Spectrosc. 2019;211:363–372. doi: 10.1016/j.saa.2018.12.034. [DOI] [PubMed] [Google Scholar]
- 86.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. Chem. 2020;387 [Google Scholar]
- 87.Liu X., et al. Hydrothermal synthesis of N-doped carbon dots for selective fluorescent sensing and cellular imaging of cobalt(II) Microchim. Acta. 2017;184 [Google Scholar]
- 88.Xie Y., et al. Green hydrothermal synthesis of N-doped carbon dots from biomass highland barley for the detection of Hg(2) Sensors. 2019;19(14) doi: 10.3390/s19143169. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Li S., et al. Fluorescent N-doped carbon dots from bacterial cellulose for highly sensitive bacterial detection. Bioresources. 2019 [Google Scholar]
- 90.Konar S., et al. N-doped carbon dot as fluorescent probe for detection of cysteamine and multicolor cell imaging. Sensor. Actuator. B Chem. 2019;286:77–85. [Google Scholar]
- 91.Zhang W.J., et al. A ratiometric fluorescent and colorimetric dual-signal sensing platform based on N-doped carbon dots for selective and sensitive detection of copper(II) and pyrophosphate ion. Sensor. Actuator. B Chem. 2019;283:215–221. [Google Scholar]
- 92.Bu L., et al. One-step synthesis of N-doped carbon dots, and their applications in curcumin sensing, fluorescent inks, and super-resolution nanoscopy. Microchim. Acta. 2019:186. doi: 10.1007/s00604-019-3762-5. [DOI] [PubMed] [Google Scholar]
- 93.Liu S., et al. Hydrothermal treatment of grass: a low-cost, green route to nitrogen-doped, carbon-rich, photoluminescent polymer nanodots as an effective fluorescent sensing platform for label-free detection of Cu(II) ions. Adv. Mater. 2012;24(15):2037–2041. doi: 10.1002/adma.201200164. [DOI] [PubMed] [Google Scholar]
- 94.Zhao A., et al. Ionic liquids as precursors for highly luminescent, surface-different nitrogen-doped carbon dots used for label-free detection of Cu2+/Fe3+ and cell imaging. Anal. Chim. Acta. 2014;809:128–133. doi: 10.1016/j.aca.2013.10.046. [DOI] [PubMed] [Google Scholar]
- 95.Wang W., et al. Facile synthesis of water-soluble and biocompatible fluorescent nitrogen-doped carbon dots for cell imaging. Analyst. 2014;139 doi: 10.1039/c3an02098c. [DOI] [PubMed] [Google Scholar]
- 96.Xu J., et al. Low-cost synthesis of carbon nanodots from natural products used as a fluorescent probe for the detection of ferrum(iii) ions in lake water. Anal. Methods. 2014;6(7):2086–2090. [Google Scholar]
- 97.Chandra S., et al. Luminescent S-doped carbon dots: an emergent architecture for multimodal applications. J. Mater. Chem. B. 2013;1:2375–2382. doi: 10.1039/c3tb00583f. [DOI] [PubMed] [Google Scholar]
- 98.Kwon W., et al. Sulfur-incorporated carbon quantum dots with a strong long-wavelength absorption band. J. Mater. Chem. C. 2013;1(10):2002–2008. [Google Scholar]
- 99.Qian Z., et al. Highly luminescent N-doped carbon quantum dots as an effective multifunctional fluorescence sensing platform. Chemistry (Weinheim an der Bergstrasse, Germany) 2014;20 doi: 10.1002/chem.201304374. [DOI] [PubMed] [Google Scholar]
- 100.Chen X., et al. Synthesis and unique photoluminescence properties of nitrogen-rich quantum dots and their applications. Angew. Chem. 2014;126 doi: 10.1002/anie.201408422. [DOI] [PubMed] [Google Scholar]
- 101.Zhang D., et al. Ionic-liquid-stabilized fluorescent probe based on S-doped carbon dot-embedded covalent-organic frameworks for determination of histamine. Mikrochim. Acta. 2019;187(1):28. doi: 10.1007/s00604-019-3833-7. [DOI] [PubMed] [Google Scholar]
- 102.Wang Y., et al. Novel visible-light-driven S-doped carbon dots/BiOI nanocomposites: improved photocatalytic activity and mechanism insight. J. Mater. Sci. 2017;52:1–12. [Google Scholar]
- 103.Mentik H., et al. Proceedings of the International Seminar of Science and Applied Technology (ISSAT 2020) Atlantis Press; 2020. Synthesis and material characterizations of S-doped carbon nanodots from blackstrap molasses. [Google Scholar]
- 104.Shamsipur M., et al. Highly selective aggregation assay for visual detection of mercury ion based on competitive binding of sulfur-doped carbon nanodots to gold nanoparticles and mercury ions. Microchim. Acta. 2016 In press. [Google Scholar]
- 105.Hu Y., et al. 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]
- 106.Xu Q., et al. Preparation of highly photoluminescent sulfur-doped carbon dots for Fe(iii) detection. J. Mater. Chem. A. 2015;3(2):542–546. [Google Scholar]
- 107.Han Y., et al. Non-metal single/dual doped carbon quantum dots: a general flame synthetic method and electro-catalytic properties. Nanoscale. 2015;7 doi: 10.1039/c4nr07116f. [DOI] [PubMed] [Google Scholar]
- 108.Mahani M., et al. A carbon dot and molecular beacon based fluorometric sensor for the cancer marker microRNA-21. Mikrochim. Acta. 2019;186(3):132. doi: 10.1007/s00604-019-3233-z. [DOI] [PubMed] [Google Scholar]
- 109.Jia Y., et al. Boron doped carbon dots as a multifunctional fluorescent probe for sorbate and vitamin B12. Mikrochim. Acta. 2019;186(2):84. doi: 10.1007/s00604-018-3196-5. [DOI] [PubMed] [Google Scholar]
- 110.Bourlinos A.B., et al. Green and simple route toward boron doped carbon dots with significantly enhanced non-linear optical properties. Carbon. 2015;83:173–179. [Google Scholar]
- 111.Wang F., et al. Fluorescence quenchometric method for determination of ferric ion using boron-doped carbon dots. Microchim. Acta. 2015;183:273–279. [Google Scholar]
- 112.Van Tam T., et al. Synthesis of B-doped graphene quantum dots as a metal-free electrocatalyst for the oxygen reduction reaction. J. Mater. Chem. A. 2017;5(21):10537–10543. [Google Scholar]
- 113.Wang Z., et al. Preparation of boron-doped carbon dots for fluorometric determination of Pb(II), Cu(II) and pyrophosphate ions. Microchim. Acta. 2017;184:4775–4783. [Google Scholar]
- 114.Yan Y., et al. B-doped graphene quantum dots implanted into bimetallic organic framework as a highly active and robust cathodic catalyst in the microbial fuel cell. Chemosphere. 2022;286 doi: 10.1016/j.chemosphere.2021.131908. [DOI] [PubMed] [Google Scholar]
- 115.Shen C., et al. Facile access to B-doped solid-state fluorescent carbon dots toward light emitting devices and cell imaging agents. J. Mater. Chem. C. 2015;3:6668–6675. [Google Scholar]
- 116.Pal A., et al. Boron doped carbon dots with unusually high photoluminescence quantum yield for ratiometric intracellular pH sensing. ChemPhysChem. 2019;20(8):1018–1027. doi: 10.1002/cphc.201900140. [DOI] [PubMed] [Google Scholar]
- 117.Peng Z., et al. Facile synthesis of “boron-doped” carbon dots and their application in visible-light-driven photocatalytic degradation of organic dyes. Nanomaterials. 2020;10:1560. doi: 10.3390/nano10081560. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Zhang Y., et al. Phosphorus-doped carbon nitride solid: enhanced electrical conductivity and photocurrent generation. J. Am. Chem. Soc. 2010;132(18):6294–6295. doi: 10.1021/ja101749y. [DOI] [PubMed] [Google Scholar]
- 119.Stegner A.R., et al. Electronic transport in phosphorus-doped silicon nanocrystal networks. Phys. Rev. Lett. 2008;100(2) doi: 10.1103/PhysRevLett.100.026803. [DOI] [PubMed] [Google Scholar]
- 120.Fang X., et al. Phosphorus-doped p-type ZnO nanorods and ZnO nanorod p−n homojunction LED fabricated by hydrothermal method. J. Phys. Chem. 2009;C:113. [Google Scholar]
- 121.Maciel I.O., et al. Synthesis, electronic structure, and Raman scattering of phosphorus-doped single-wall carbon nanotubes. Nano Lett. 2009;9(6):2267–2272. doi: 10.1021/nl9004207. [DOI] [PubMed] [Google Scholar]
- 122.Chen Z., Wang S., Yang X. Phosphorus-doped carbon dots for sensing both Au (III) and l-methionine. J. Photochem. Photobiol. Chem. 2018;365:178–184. [Google Scholar]
- 123.Shi D., et al. P-doped carbon dots act as a nanosensor for trace 2,4,6-trinitrophenol detection and a fluorescent reagent for biological imaging. RSC Adv. 2015;5(119):98492–98499. [Google Scholar]
- 124.Molkenova A., Atabaev T.S. Phosphorus-doped carbon dots (P-CDs) from dextrose for low-concentration ferric ions sensing in water. Optik. 2019 [Google Scholar]
- 125.Ju Y.J., et al. Green synthesis of blue fluorescent P-doped carbon dots for the selective determination of picric acid in an aqueous medium. Anal. Sci. 2019;35(2):147–152. doi: 10.2116/analsci.18P372. [DOI] [PubMed] [Google Scholar]
- 126.Omer K.M., Hassan A.Q. Chelation-enhanced fluorescence of phosphorus doped carbon nanodots for multi-ion detection. Microchim. Acta. 2017;184(7):2063–2071. [Google Scholar]
- 127.Mathew R.M., et al. Metal free, phosphorus doped carbon nanodot mediated photocatalytic reduction of methylene blue. React. Kinet. Mech. Catal. 2020;129(2):1131–1143. [Google Scholar]
- 128.Zhang J., et al. A fluorescence method based on N, S-doped carbon dots for detection of ammonia in aquaculture water and freshness of fish. Sustainability. 2021;13:8255. [Google Scholar]
- 129.Kaleem W., Kumar A., Panda D. Luminescent N, S-doped carbon nanodot: an effective two-fluorophore system of pyridone and thiazolopyridone. J. Phys. Chem. C. 2018;122 [Google Scholar]
- 130.Deng Y., et al. Preparation of N/S doped carbon dots and their application in nitrite detection. RSC Adv. 2021;11(18):10922–10928. doi: 10.1039/d0ra10766b. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131.Lu W., et al. Comparative study for N and S doped carbon dots: synthesis, characterization and applications for Fe3+ probe and cellular imaging. Anal. Chim. Acta. 2015;898:116–127. doi: 10.1016/j.aca.2015.09.050. [DOI] [PubMed] [Google Scholar]
- 132.Ibrayev N., et al. Optical properties of N- and S-doped carbon dots based on citric acid and L-cysteine. Fullerenes, Nanotub. Carbon Nanostruct. 2021;30:22–26. [Google Scholar]
- 133.Shen J., et al. Highly fluorescent N, S-co-doped carbon dots and their potential applications as antioxidants and sensitive probes for Cr (VI) detection. Sensor. Actuator. B Chem. 2017;248:92–100. [Google Scholar]
- 134.Pajewska-Szmyt M., Buszewski B., Gadzała-Kopciuch R. Sulphur and nitrogen doped carbon dots synthesis by microwave assisted method as quantitative analytical nano-tool for mercury ion sensing. Mater. Chem. Phys. 2020;242 [Google Scholar]
- 135.Han Z., et al. One-pot hydrothermal synthesis of nitrogen and sulfur co-doped carbon dots and their application for sensitive detection of curcumin and temperature. Microchem. J. 2019;146:300–308. [Google Scholar]
- 136.Ahmed F., et al. “ON-OFF-ON” fluorescence switches based on N,S-doped carbon dots: facile hydrothermal growth, selective detection of Hg2+, and as a reversive probe for guanine. Anal. Chim. Acta. 2021;1183 doi: 10.1016/j.aca.2021.338977. [DOI] [PubMed] [Google Scholar]
- 137.Cheng S., et al. One-step synthesis of N, S-doped carbon dots with orange emission and their application in tetracycline antibiotics, quercetin sensing, and cell imaging. Mikrochim. Acta. 2021;188(10):325. doi: 10.1007/s00604-021-04969-w. [DOI] [PubMed] [Google Scholar]
- 138.Liu Y., et al. Red emission B, N, S-Co-doped carbon dots for colorimetric and fluorescent dual mode detection of Fe3+ ions in complex biological fluids and living cells. ACS Appl. Mater. Interfaces. 2017;9 doi: 10.1021/acsami.6b15746. [DOI] [PubMed] [Google Scholar]
- 139.Li J., et al. The one-step preparation of green-emissioned carbon dots through hydrothermal route and its application. J. Nanomater. 2019;2019:1–10. [Google Scholar]
- 140.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. doi: 10.1016/j.aca.2015.07.051. [DOI] [PubMed] [Google Scholar]
- 141.Sun C., et al. Synthesis of nitrogen and sulfur Co-doped carbon dots from garlic for selective detection of Fe(3.) Nanoscale Res. Lett. 2016;11(1):110. doi: 10.1186/s11671-016-1326-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142.Yang M., et al. N, S co-doped carbon dots with high quantum yield: tunable fluorescence in liquid/solid and extensible applications. J. Nanoparticle Res. 2017;19 [Google Scholar]
- 143.Cheng C., Xing M., Wu Q. A universal facile synthesis of nitrogen and sulfur co-doped carbon dots from cellulose-based biowaste for fluorescent detection of Fe3+ ions and intracellular bioimaging. Mater. Sci. Eng. C. 2019;99:611–619. doi: 10.1016/j.msec.2019.02.003. [DOI] [PubMed] [Google Scholar]
- 144.Belal F., et al. A novel eplerenone ecofriendly fluorescent nanosensor based on nitrogen and sulfur-carbon quantum dots. J. Fluoresc. 2021;31(1):85–90. doi: 10.1007/s10895-020-02638-4. [DOI] [PubMed] [Google Scholar]
- 145.Guo Z., et al. A facile synthesis of high-efficient N,S co-doped carbon dots for temperature sensing application. Dyes Pigments. 2020;173 [Google Scholar]
- 146.Zhang C., et al. 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. doi: 10.1016/j.jphotobiol.2017.06.024. [DOI] [PubMed] [Google Scholar]
- 147.Cui X., et al. Dual functional N- and S-co-doped carbon dots as the sensor for temperature and Fe3+ ions. Sensor. Actuator. B Chem. 2017;242:1272–1280. [Google Scholar]
- 148.Bonet-San-Emeterio M., et al. Modification of electrodes with N-and S-doped carbon dots. Evaluation of the electrochemical response. Talanta. 2020;212 doi: 10.1016/j.talanta.2020.120806. [DOI] [PubMed] [Google Scholar]
- 149.Sahiner N., et al. Nitrogen and sulfur doped carbon dots from amino acids for potential biomedical applications. J. Fluorescence. 2019;29(5):1191–1200. doi: 10.1007/s10895-019-02431-y. [DOI] [PubMed] [Google Scholar]
- 150.Ye Q., et al. Formation of N, S-codoped fluorescent carbon dots from biomass and their application for the selective detection of mercury and iron ion. Spectrochim. Acta Mol. Biomol. Spectrosc. 2017;173:854–862. doi: 10.1016/j.saa.2016.10.039. [DOI] [PubMed] [Google Scholar]
- 151.Guo Y., et al. Carbon dots doped with nitrogen and sulfur and loaded with copper(II) as a “turn-on” fluorescent probe for cystein, glutathione and homocysteine. Microchim. Acta. 2016;183:1409–1416. [Google Scholar]
- 152.Chandra S., et al. Sulphur and nitrogen doped carbon dots: a facile synthetic strategy for multicolour bioimaging, tiopronin sensing, and Hg2+ ion detection. Nano-Struct. Nano. Obj. 2017;12:10–18. [Google Scholar]
- 153.Dhenadhayalan N., Lin K.C., Saleh T. Recent advances in functionalized carbon dots toward the design of efficient materials for sensing and catalysis applications. Small. 2019;16 doi: 10.1002/smll.201905767. [DOI] [PubMed] [Google Scholar]
- 154.Arshad F., Sk M. Aggregation-induced red shift in N, S-doped chiral carbon dot emission for moisture sensing. New J. Chem. 2019;43 [Google Scholar]
- 155.Bandi R., et al. Facile Conversion of Toxic Cigarette Butts to N,S-Codoped Carbon Dots and Their Application in Fluorescent Film, Security Ink, Bioimaging, Sensing and Logic Gate Operation. 2018;3:13454–13466. doi: 10.1021/acsomega.8b01743. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156.Tang M., et al. 2019. Nitrogen- and Sulfur-Doped Carbon Dots as Peroxidase Mimetics: Colorimetric Determination of Hydrogen Peroxide and Glutathione, and Fluorimetric Determination of lead(II) p. 186. [DOI] [PubMed] [Google Scholar]
- 157.Liu J., et al. Facile synthesis of N, B-doped carbon dots and their application for multisensor and cellular imaging. Ind. Eng. Chem. Res. 2017;56 [Google Scholar]
- 158.Arul V., et al. Efficient green synthesis of N,B co-doped bright fluorescent carbon nanodots and their electrocatalytic and bio-imaging applications. Diam. Relat. Mater. 2021;116 [Google Scholar]
- 159.Mohammed L., Omer K. Dual functional highly luminescence B, N Co-doped carbon nanodots as nanothermometer and Fe3+/Fe2+ sensor. Sci. Rep. 2020;10 doi: 10.1038/s41598-020-59958-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 160.Wang Z.X., et al. Fluorometric and colorimetric determination of hypochlorite using carbon nanodots doped with boron and nitrogen. Mikrochim. Acta. 2019;186(6):328. doi: 10.1007/s00604-019-3443-4. [DOI] [PubMed] [Google Scholar]
- 161.Tian B., et al. B- and N-doped carbon dots by one-step microwave hydrothermal synthesis: tracking yeast status and imaging mechanism. J. Nanobiotechnol. 2021;19 doi: 10.1186/s12951-021-01211-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 162.Liu H., et al. Boron and nitrogen co-doped carbon dots for boosting electrocatalytic oxygen reduction. N. Carbon Mater. 2021 [Google Scholar]
- 163.Tian T., et al. One-pot synthesis of boron and nitrogen co-doped carbon dots as the fluorescence probe for dopamine based on the redox reaction between Cr(VI) and dopamine. Sensor. Actuator. B Chem. 2017;240:1265–1271. [Google Scholar]
- 164.Tummala S., Lee C.H., Ho Y.P. Boron, and nitrogen co-doped carbon dots as a multiplexing probe for sensing of p-nitrophenol, Fe (III), and temperature. Nanotechnology. 2021;32(26) doi: 10.1088/1361-6528/abeeb6. [DOI] [PubMed] [Google Scholar]
- 165.Bian W., et al. Boron and nitrogen co-doped carbon dots as a sensitive fluorescent probe for the detection of curcumin. Luminescence. 2017;33 doi: 10.1002/bio.3390. [DOI] [PubMed] [Google Scholar]
- 166.Zhou J., et al. Carbon dots doped with heteroatoms for fluorescent bioimaging: a review. Microchim. Acta. 2017;184(2):343–368. [Google Scholar]
- 167.Qie X., et al. High photoluminescence nitrogen, phosphorus co-doped carbon nanodots for assessment of microbial viability. Colloids Surf. B Biointerfaces. 2020;191 doi: 10.1016/j.colsurfb.2020.110987. [DOI] [PubMed] [Google Scholar]
- 168.Shi B., et al. Nitrogen and phosphorus Co-doped carbon nanodots as a novel fluorescent probe for highly sensitive detection of Fe(3+) in human serum and living cells. ACS Appl. Mater. Interfaces. 2016;8(17):10717–10725. doi: 10.1021/acsami.6b01325. [DOI] [PubMed] [Google Scholar]
- 169.Choi Y., Zheng X.T., Tan Y.N. Bioinspired carbon dots (biodots): emerging fluorophores with tailored multiple functionalities for biomedical, agricultural and environmental applications. Mol. Sys. Des. Eng. 2020;5(1):67–90. [Google Scholar]
- 170.Li J., et al. Highly N,P-doped carbon dots: rational design, photoluminescence and cellular imaging. Microchim. Acta. 2017;184(8):2933–2940. [Google Scholar]
- 171.Zhao D., et al. Rapid and low-temperature synthesis of N, P co-doped yellow emitting carbon dots and their applications as antibacterial agent and detection probe to Sudan Red I. Mater Sci Eng C Mater Biol Appl. 2021;119 doi: 10.1016/j.msec.2020.111468. [DOI] [PubMed] [Google Scholar]
- 172.Tall A., et al. Green emitting N, P-doped carbon dots as efficient fluorescent nanoprobes for determination of Cr(VI) in water and soil samples. Microchem. J. 2021;166 [Google Scholar]
- 173.Du F., et al. Facile, rapid synthesis of N,P-dual-doped carbon dots as a label-free multifunctional nanosensor for Mn(VII) detection, temperature sensing and cellular imaging. Sensor. Actuator. B Chem. 2018;277 [Google Scholar]
- 174.Gu D., et al. Nitrogen and phosphorus co-doped carbon dots derived from lily bulbs for copper ion sensing and cell imaging. Opt. Mater. 2018;83:272–278. [Google Scholar]
- 175.Singh R., et al. Study of the fluorescence based applications of water soluble (N, P) doped carbon dots synthesized via microwave assisted green pyrolysis. Nanosci. Nanotechnol. - Asia. 2020;10:827–839. [Google Scholar]
- 176.Xu Q., et al. Synthesis, mechanical investigation, and application of nitrogen and phosphorus co-doped carbon dots with a high photoluminescent quantum yield. Nano Res. 2018;11(7):3691–3701. [Google Scholar]
- 177.Dong G., et al. Facile synthesis of N, P-doped carbon dots from maize starch via a solvothermal approach for the highly sensitive detection of Fe3+ RSC Adv. 2020;10(55):33483–33489. doi: 10.1039/d0ra06209j. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 178.Shamsipur M., et al. One-step synthesis and characterization of highly luminescent nitrogen and phosphorus co-doped carbon dots and their application as highly selective and sensitive nanoprobes for low level detection of uranyl ion in hair and water samples and application to cellular imaging. Sensor. Actuator. B Chem. 2018;257:772–782. [Google Scholar]
- 179.Bajpai V.K., et al. Multifunctional N-P-doped carbon dots for regulation of apoptosis and autophagy in B16F10 melanoma cancer cells and in vitro imaging applications. Theranostics. 2020;10(17):7841–7856. doi: 10.7150/thno.42291. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 180.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 doi: 10.1016/j.foodchem.2020.127677. [DOI] [PubMed] [Google Scholar]
- 181.Li X., et al. Dual-excitation and dual-emission carbon dots for Fe3+ detection, temperature sensing, and lysosome targeting. Anal. Methods. 2021;13 doi: 10.1039/d1ay01165k. [DOI] [PubMed] [Google Scholar]
- 182.Yang M., et al. Carbon dots co-doped with nitrogen and chlorine for "off-on" fluorometric determination of the activity of acetylcholinesterase and for quantification of organophosphate pesticides. Mikrochim. Acta. 2019;186(8):585. doi: 10.1007/s00604-019-3715-z. [DOI] [PubMed] [Google Scholar]
- 183.Huang J., et al. Nitrogen and chlorine co-doped carbon dots with synchronous excitation of multiple luminescence centers for blue-white emission. New J. Chem. 2021;45(16):7056–7059. [Google Scholar]
- 184.Wang N., et al. Deep eutectic solvent-assisted preparation of nitrogen/chloride-doped carbon dots for intracellular biological sensing and live cell imaging. ACS Appl. Mater. Interfaces. 2018;10 doi: 10.1021/acsami.8b00947. [DOI] [PubMed] [Google Scholar]
- 185.Chang S., et al. Label-free chlorine and nitrogen-doped fluorescent carbon dots for target imaging of lysosomes in living cells. Mikrochim. Acta. 2020;187(8):435. doi: 10.1007/s00604-020-04412-6. [DOI] [PubMed] [Google Scholar]
- 186.Hu Q., et al. Ultrafast and energy-saving synthesis of nitrogen and chlorine Co-doped carbon nanodots via neutralization heat for selective detection of Cr(VI) in aqueous phase. Sensors. 2018;18:3416. doi: 10.3390/s18103416. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 187.Peng B., et al. Smartphone colorimetric determination of hydrogen peroxide in real samples based on B, N, and S co-doped carbon dots probe. Anal. Bioanal. Chem. 2020;412(4):861–870. doi: 10.1007/s00216-019-02284-1. [DOI] [PubMed] [Google Scholar]
- 188.Hao Y., et al. Nitrogen, sulfur, phosphorus, and chlorine co-doped carbon nanodots as an "off-on" fluorescent probe for sequential detection of curcumin and europium ion and luxuriant applications. Mikrochim. Acta. 2021;188(1):16. doi: 10.1007/s00604-020-04618-8. [DOI] [PubMed] [Google Scholar]
- 189.Dong Y., et al. Chemiluminescence of carbon dots induced by diperiodato-nicklate (IV) in alkaline solution and its application to a quenchometric flow-injection assays of paracetamol, L-cysteine and glutathione. Microchim. Acta. 2014;182:1071–1077. [Google Scholar]
- 190.Yuan Y., et al. A fluorescence switch sensor used for D-Penicillamine sensing and logic gate based on the fluorescence recovery of carbon dots. Sensor. Actuator. B Chem. 2016;236:565–573. [Google Scholar]
- 191.Liu L., et al. Sensitive determination of kaempferol using carbon dots as a fluorescence probe. Talanta. 2015;144:390–397. doi: 10.1016/j.talanta.2015.07.004. [DOI] [PubMed] [Google Scholar]
- 192.Dong B., et al. Homogeneous fluorescent immunoassay for the simultaneous detection of chloramphenicol and amantadine via the duplex FRET between carbon dots and WS2 nanosheets. Food Chem. 2020;327 doi: 10.1016/j.foodchem.2020.127107. [DOI] [PubMed] [Google Scholar]
- 193.Liu L., et al. One-step synthesis of fluorescent carbon dots for sensitive and selective detection of hyperin. Talanta. 2018;186:315–321. doi: 10.1016/j.talanta.2018.04.065. [DOI] [PubMed] [Google Scholar]
- 194.Azizi B., Farhadi K., Samadi N. Functionalized carbon dots from zein biopolymer as a sensitive and selective fluorescent probe for determination of sumatriptan. Microchem. J. 2019;146:965–973. [Google Scholar]
- 195.Zhang T., et al. vol. 134. 2022. (Ratiometric Fluorescent Probe Based on Carbon Dots and Zn-Doped CdTe QDs for Detection of 6-Mercaptopurine). [Google Scholar]
- 196.Wang S., et al. vol. 371. 2022. (On-site, Rapid, and Facile Determination of Gentamicin Using a Fluorescent Resonance Energy Transfer Sensor Constructed from Nitrogen-Carbon Quantum Dots Functionalized by 4, 5-imidazole Dicarboxylic Acid). [DOI] [PubMed] [Google Scholar]
- 197.Sajwan R.K., Solanki P.R.J.F.C. vol. 415. 2023. (Gold@ Carbon Quantum Dots Nanocomposites Based Two-In-One Sensor: a Novel Approach for Sensitive Detection of Aminoglycosides Antibiotics in Food Samples). [DOI] [PubMed] [Google Scholar]
- 198.Kong W., et al. vol. 973. 2017. pp. 91–99. (Carbon Dots for Fluorescent Detection of α-glucosidase Activity Using Enzyme Activated Inner Filter Effect and its Application to Anti-diabetic Drug Discovery). [DOI] [PubMed] [Google Scholar]
- 199.Zhang X., et al. Rapid microwave synthesis of N-doped carbon nanodots with high fluorescence brightness for cell imaging and sensitive detection of iron (III) Opt. Mater. 2017;64:1–8. [Google Scholar]
- 200.Yu Z.-y., Duo S.J.O.M. Facile, ultrafast synthesis of up- and down-conversion fluorescent carbon dots for highly sensitive detection of hematin. 2020;107 [Google Scholar]
- 201.Akhgari F., Samadi N., Farhadi K.J.J.o.F. vol. 27. 2017. pp. 921–927. (Fluorescent Carbon Dot as Nanosensor for Sensitive and Selective Detection of Cefixime Based on Inner Filter Effect). [DOI] [PubMed] [Google Scholar]
- 202.Uriarte D., Domini C., Garrido M. New carbon dots based on glycerol and urea and its application in the determination of tetracycline in urine samples. Talanta. 2019;201 doi: 10.1016/j.talanta.2019.04.001. [DOI] [PubMed] [Google Scholar]
- 203.Zhang H., et al. vol. 255. 2021. (Nitrogen-doped Carbon Dots Derived from Hawthorn for the Rapid Determination of Chlortetracycline in Pork Samples). [DOI] [PubMed] [Google Scholar]
- 204.Sang Y., et al. vol. 362. 2022. (Color-multiplexing Europium Doped Carbon Dots for Highly Selective and Dosage-Sensitive Cascade Visualization of Tetracycline and Al3+). [Google Scholar]
- 205.Li Y., et al. vol. 358. 2022. (Fabrication of Carbon Dots@ Hierarchical Mesoporous ZIF-8 for Simultaneous Ratiometric Fluorescence Detection and Removal of Tetracycline Antibiotics). [Google Scholar]
- 206.Zhang J., et al. A molecularly imprinted fluorescence sensor for sensitive detection of tetracycline using nitrogen-doped carbon dots-embedded zinc-based metal-organic frameworks as signal-amplifying tags. 2023;1251 doi: 10.1016/j.aca.2023.341032. [DOI] [PubMed] [Google Scholar]
- 207.Jia Y., et al. vol. 186. 2019. pp. 1–10. (Boron Doped Carbon Dots as a Multifunctional Fluorescent Probe for Sorbate and Vitamin B12). [DOI] [PubMed] [Google Scholar]
- 208.Song J., et al. Fluorescent boron and nitrogen co-doped carbon dots with high quantum yield for the detection of nimesulide and fluorescence staining. Spectrochim. Acta Mol. Biomol. Spectrosc. 2019;216:296–302. doi: 10.1016/j.saa.2019.03.074. [DOI] [PubMed] [Google Scholar]
- 209.Hu Q., et al. Nitrogen and chlorine dual-doped carbon nanodots for determination of curcumin in food matrix via inner filter effect. Food Chem. 2019;280:195–202. doi: 10.1016/j.foodchem.2018.12.050. [DOI] [PubMed] [Google Scholar]
- 210.Zhu J., et al. 2020. Nitrogen and Fluorine Co-doped Green Fluorescence Carbon Dots as a Label-free Probe for Determination of Cytochrome C in Serum and Temperature Sensing. [DOI] [PubMed] [Google Scholar]
- 211.Tan Q., et al. One-pot hydrothermal method synthesis carbon dots as a fluorescent sensor for the sensitive and selective detection of clioquinol. 2022;123 [Google Scholar]
- 212.Liang J.-M., et al. vol. 275. 2022. (One-pot Hydrothermal Synthesis of Si-Doped Carbon Quantum Dots with Up-Conversion Fluorescence as Fluorescent Probes for Dual-Readout Detection of Berberine Hydrochloride). [DOI] [PubMed] [Google Scholar]
- 213.Rizk M., et al. vol. 278. 2022. (Fluorescent Carbon Dots as Selective Nano Probe for Determination of Diacerein in Presence of Co-formulated Drugs). [DOI] [PubMed] [Google Scholar]
- 214.Wu M., et al. vol. 282. 2022. (Dual-mode Colorimetric and Fluorescence Sensing System for the Detection of Captopril Based on Fe/NC Nanozymes and Carbon Dots). [DOI] [PubMed] [Google Scholar]
- 215.Mohammadinejad A., et al. Application of green-synthesized carbon dots for imaging of cancerous cell lines and detection of anthraquinone drugs using silica-coated CdTe quantum dots-based ratiometric fluorescence sensor. 2023;288 doi: 10.1016/j.saa.2022.122200. [DOI] [PubMed] [Google Scholar]
- 216.Zhong Y., et al. vol. 382. 2023. (Red Emission Carbon Dots for Mitoxantrone Detection). [Google Scholar]
- 217.Hu Z., Jiao X.-Y., Xu L., The N. S co-doped carbon dots with excellent luminescent properties from green tea leaf residue and its sensing of gefitinib. Microchem. J. 2020;154 [Google Scholar]
- 218.Zhu J., et al. vol. 114. 2021. (Green Preparation of Carbon Dots from Plum as a Ratiometric Fluorescent Probe for Detection of Doxorubicin). [Google Scholar]
- 219.Qandeel N.A., et al. Fast one-pot microwave-assisted green synthesis of highly fluorescent plant-inspired S, N-self-doped carbon quantum dots as a sensitive probe for the antiviral drug nitazoxanide and hemoglobin. 2023;1237 doi: 10.1016/j.aca.2022.340592. [DOI] [PubMed] [Google Scholar]
- 220.Chen J., et al. vol. 417. 2023. (Highly Photoluminescent Nitrogen-Doped Carbon Quantum Dots as a Green Fluorescence Probe for Determination of Myricetin). [DOI] [PubMed] [Google Scholar]
- 221.Atulbhai S.V., et al. vol. 287. 2023. (Microwave Synthesis of Blue Emissive Carbon Dots from 5-sulpho Anthranilic Acid and 1, 5-diphenyl Carbazide for Sensing of Levocetirizine and Niflumic Acid). [DOI] [PubMed] [Google Scholar]
- 222.Meng Y.-t., et al. 2020. Facile Synthesis of Orange Fluorescence Multifunctional Carbon Dots for Label-free Detection of Vitamin B12 and Endogenous/exogenous Peroxynitrite. [DOI] [PubMed] [Google Scholar]
- 223.Liu H., et al. vol. 209. 2020. (Fabrication of Carbon Dots@restricted Access Molecularly Imprinted Polymers for Selective Detection of Metronidazole in Serum). [DOI] [PubMed] [Google Scholar]
- 224.Purbia R., Paria S.J.B. vol. 79. 2016. pp. 467–475. (Bioelectronics, A Simple Turn on Fluorescent Sensor for the Selective Detection of Thiamine Using Coconut Water Derived Luminescent Carbon Dots). [DOI] [PubMed] [Google Scholar]
- 225.Amin N., et al. vol. 1030. 2018. pp. 183–193. (Green and Cost-Effective Synthesis of Carbon Dots from Date Kernel and Their Application as a Novel Switchable Fluorescence Probe for Sensitive Assay of Zoledronic Acid Drug in Human Serum and Cellular Imaging). [DOI] [PubMed] [Google Scholar]
- 226.Ensafi A.A., et al. A novel one-step and green synthesis of highly fluorescent carbon dots from saffron for cell imaging and sensing of prilocaine. Sensor. Actuator. B Chem. 2017;253:451–460. [Google Scholar]
- 227.Guo X.-j., et al. vol. 219. 2019. pp. 15–22. (Fluorescent Carbon Dots Based Sensing System for Detection of Enrofloxacin in Water Solutions). [DOI] [PubMed] [Google Scholar]
- 228.Chen X., et al. vol. 270. 2022. (Dual-mode Turn-On Ratiometric Fluorescence Sensor Based on Carbon Dots and CuInS2/ZnS Quantum Dots for Detection of Chlorotetracycline). [DOI] [PubMed] [Google Scholar]
- 229.Wang Y., et al. Ionothermal synthesis of carbon dots from cellulose in deep eutectic solvent: A sensitive probe for detecting Cu2+ and glutathione with “off-on” pattern. 2022;599 [Google Scholar]
- 230.Chen J., et al. vol. 405. 2023. (A Facile “Off–on” Fluorescence Sensor for Pentachlorophenol Detection Based on Natural N and S Co-doped Carbon Dots from Crawfish Shells). [DOI] [PubMed] [Google Scholar]
- 231.Mishra R.K., et al. In situ formation of carbon dots aids ampicillin sensing. 2016;8:2441–2447. [Google Scholar]
- 232.Al-Hashimi B., et al. vol. 244. 2020. (Inner Filter Effect as a Sensitive Sensing Platform for Detection of Nitrofurantoin Using Luminescent Drug-Based Carbon Nanodots). [DOI] [PubMed] [Google Scholar]
- 233.Shariati R., et al. Application of coated green source carbon dots with silica molecularly imprinted polymers as a fluorescence probe for selective and sensitive determination of phenobarbital. Talanta. 2019;194:143–149. doi: 10.1016/j.talanta.2018.09.069. [DOI] [PubMed] [Google Scholar]
- 234.Ghafarloo A., et al. vol. 388. 2020. (Sensitive and Selective Spectrofluorimetric Determination of Clonazepam Using Nitrogen-Doped Carbon Dots). [Google Scholar]
- 235.Bu X., et al. 2020. Self-assembly of DNA-Templated Copper Nanoclusters and Carbon Dots for Ratiometric Fluorometric and Visual Determination of Arginine and Acetaminophen with a Logic-Gate Operation; p. 187. [DOI] [PubMed] [Google Scholar]
- 236.Nazari F., Tabaraki R.J. Molecular, and b. spectroscopy, Sensitive fluorescence detection of atorvastatin by doped carbon dots synthesized in deep eutectic media. 2020;236 doi: 10.1016/j.saa.2020.118341. [DOI] [PubMed] [Google Scholar]
- 237.Zhang W., et al. vol. 128. 2022. (Nitrogen and Phosphorus Co-doped Carbon Dots as an Effective Fluorescence Probe for the Detection of Doxorubicin and Cell Imaging). [Google Scholar]
- 238.Abd Elhaleem S.M., et al. vol. 272. 2022. Turn-off fluorescence of nitrogen and sulfur carbon quantum dots as effective fluorescent probes for determination of imatinib. (Application to Biological Fluids). [DOI] [PubMed] [Google Scholar]
- 239.Fang F., et al. vol. 139. 2023. (A Fluorescent Nanoprobe Based on N/S Co-doped Carbon Dots Coupled with Molecularly Imprinted Polymers for the Detection of Curcumin). [Google Scholar]
- 240.Venugopalan P., Vidya N., Spectroscopy B. Microwave-assisted green synthesis of carbon dots derived from wild lemon (Citrus pennivesiculata) leaves as a fluorescent probe for tetracycline sensing in water. 2023;286 doi: 10.1016/j.saa.2022.122024. [DOI] [PubMed] [Google Scholar]
- 241.Qian J., et al. 2018. Aconitic Acid Derived Carbon Dots: Conjugated Interaction for the Detection of Folic Acid and Fluorescence Targeted Imaging of Folate Receptor Overexpressed Cancer Cells. [Google Scholar]
- 242.Qin J., Zhang L., Yang R. Solid pyrolysis synthesis of excitation-independent emission carbon dots and its application to isoniazid detection. J. Nanoparticle Res. 2019;21 [Google Scholar]
- 243.Yu C., et al. vol. 192. 2020. (N-Doped Carbon Quantum Dots from Osmanthus Fragrans as a Novel Off-On Fluorescent Nanosensor for Highly Sensitive Detection of Quercetin and Aluminium Ion, and Cell Imaging). [DOI] [PubMed] [Google Scholar]
- 244.Cai R., et al. vol. 361. 2022. (Determination and the Pharmacokinetic Study of Tigecycline by Fluorescence Strategy with F, N Codoping Carbon Dots as Probe). [Google Scholar]
- 245.Miao J., et al. vol. 384. 2023. (A Triple-Emission Ratiometric Fluorescence Sensor Based on Carbon Dots-Au Nanoclusters Nanocomposite for Detection of Tetracycline). [Google Scholar]
- 246.Camilus N., et al. vol. 279. 2022. (Selective Detection of Nitrotyrosine Using Dual-Fluorescent Carbon Dots). [DOI] [PubMed] [Google Scholar]
- 247.Ren X., et al. Sensitive detection of dopamine and quinone drugs based on the quenching of the fluorescence of carbon dots. Sci. Bull. 2016;61(20):1615–1623. [Google Scholar]
- 248.Xiao J., et al. vol. 192. 2020. (Determination of Chondroitin Sulfate in Synovial Fluid and Drug by Ratiometric Fluorescence Strategy Based on Carbon Dots Quenched FAM-Labeled ssDNA). [DOI] [PubMed] [Google Scholar]
- 249.Muthusankar G., Devi R.K., Gopu G. Nitrogen-doped carbon quantum dots embedded Co3O4 with multiwall carbon nanotubes: an efficient probe for the simultaneous determination of anticancer and antibiotic drugs. Biosens. Bioelectron. 2020;150 doi: 10.1016/j.bios.2019.111947. [DOI] [PubMed] [Google Scholar]
- 250.Amin N., et al. Green and cost-effective synthesis of carbon dots from date kernel and their application as a novel switchable fluorescence probe for sensitive assay of Zoledronic acid drug in human serum and cellular imaging. Anal. Chim. Acta. 2018;1030:183–193. doi: 10.1016/j.aca.2018.05.014. [DOI] [PubMed] [Google Scholar]
- 251.Mutuyimana F.P., et al. Synthesis of orange-red emissive carbon dots for fluorometric enzymatic determination of glucose. Microchim. Acta. 2018;185(11):518. doi: 10.1007/s00604-018-3041-x. [DOI] [PubMed] [Google Scholar]
- 252.Khasawneh O.F.S., Palaniandy P. Occurrence and removal of pharmaceuticals in wastewater treatment plants. Process Saf. Environ. Protect. 2021;150:532–556. [Google Scholar]
- 253.OECD . 2019. Pharmaceutical Residues in Freshwater. [Google Scholar]
- 254.Drugs in the water. 2011. https://www.health.harvard.edu/newsletter_article/drugs-in-the-water Available from:
- 255.Dutta J., Ahmad A. Removal of antibiotic from the water environment by the adsorption technologies: a review. Water Sci. Technol. 2020;82 doi: 10.2166/wst.2020.335. [DOI] [PubMed] [Google Scholar]
- 256.Karungamye P. 2020. Methods Used for Removal of Pharmaceuticals from Wastewater: A Review; pp. 412–428. [Google Scholar]
- 257.Sadiq S.T., Yaşa İ. New techniques used for removing antibiotic residues and antibiotic resistance genes from water. Rec. Advan. Biol. Med. 2019;5 [Google Scholar]
- 258.Chen J., et al. Removal of steroid hormones and biocides from rural wastewater by an integrated constructed wetland. Sci. Total Environ. 2019;660:358–365. doi: 10.1016/j.scitotenv.2019.01.049. [DOI] [PubMed] [Google Scholar]
- 259.Gómez-Espinosa R., Arizmendi-Cotero D. 2020. Use of Membrane for Removal of Nonsteroidal Anti-inflammatory Drugs. [Google Scholar]
- 260.Khan A.N., et al. The role of graphene oxide and graphene oxide-based nanomaterials in the removal of pharmaceuticals from aqueous media: a review. 2017;24:7938–7958. doi: 10.1007/s11356-017-8388-8. [DOI] [PubMed] [Google Scholar]
- 261.Mi X., et al. Electro-Fenton degradation of ciprofloxacin with highly ordered mesoporous MnCo2O4-CF cathode: enhanced redox capacity and accelerated electron transfer. Chem. Eng. J. 2019;358:299–309. [Google Scholar]
- 262.Chen Q., Xin Y., Zhu X. Au-Pd nanoparticles-decorated TiO2 nanobelts for photocatalytic degradation of antibiotic levofloxacin in aqueous solution. Electrochim. Acta. 2015;186:34–42. [Google Scholar]
- 263.Boakye-Yiadom K.O., et al. Carbon dots: applications in bioimaging and theranostics. Int. J. Pharm. 2019;564:308–317. doi: 10.1016/j.ijpharm.2019.04.055. [DOI] [PubMed] [Google Scholar]
- 264.Wan J., et al. Advanced nanomaterials for energy-related applications. J. Nanomater. 2015;2015:1–2. [Google Scholar]
- 265.Kang C., et al. A review of carbon dots produced from biomass wastes. Nanomaterials. 2020;10:2316. doi: 10.3390/nano10112316. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 266.Kang Z., Lee S.-T., dots Carbon. Advances in nanocarbon applications. Nanoscale. 2019;11 doi: 10.1039/c9nr05647e. [DOI] [PubMed] [Google Scholar]














