9
Recent Advances in Carbon Dots for Bioanalysis and the Future Perspectives

Jessica Fung Yee Fong1 Yann Huey Ng1 and Sing Muk Ng1,2

1 Swinburne Sarawak Research Centre for Sustainable Technologies, Swinburne University of Technology, Malaysia

2 Faculty of Engineering, Computing and Science, Swinburne University of Technology, Malaysia

9.1 Introduction

Nanomaterials are especially interesting in comparison to their bulk form due to their tunable physical, chemical, electronic, thermal, mechanical, and optical properties. Carbon‐based nanomaterials fullerene, carbon nanotubes (CNTs), graphene, carbon dots (CDs) and nanodiamond, have been studied extensively due to their intrinsic properties. Among these, CNTs, graphene, and CDs have attracted tremendous amount of attention for further bioanalytical applications due to their ease of functionalization for detection of analytes of interest. CNTs and graphene are similar in structure where CNTs can be viewed as seamless cylinders rolled by graphene sheets. Despite their electrical and electrochemical properties, these carbon nanomaterials typically requires complicated and costly instrumental setup, specific experimental conditions and synthesis methods such as arc discharge, laser ablation, and chemical vapor deposition. These are not favorable for laboratory settings with limited resources. In view of that, CDs that are easily attainable from renewable sources or waste materials via facile and simple synthesis routes are more ideal and cost‐effective option to be molded and extended for bioanalysis.

CDs are basically nanoparticles that are made up of majority carbon element and in some cases, blended with other impurities such as oxygen and nitrogen. It has caught the attention of the research community as it portrays very similar properties to the well‐known quantum dots (QDs), but with rather easier synthesis route and using less toxic starting precursors. It has been well demonstrated that synthesis of CDs can be from simple organic molecules [1], polymers [2, 3], or even waste biomass [4, 5]. Normally, it is not required to use a specific pure chemical as starting precursor, while the carbonization can be performed in dry form via thermal treatment [6] or just in pure water as solvent in the case of hydrothermal treatment [7]. There is also not much on the generation of toxic side products from the synthesis process, eventually making the whole synthesis process cleaner and greener. All these means that CDs are good alternative to replace the QDs in various applications. CDs have all the unique functionalities of the QDs for an intended application, but less harmful to the environments. It is even more promising when the mentioned environment refers to a bioenvironment in an organism that is often susceptible to toxic compounds. In other words, CDs can be applied directly for in vivo applications. Based on this conditional advantage, researchers have been exploring the practical usage of CDs for application in the biological system. One of the focuses is on bioanalysis, an analytical method that deals with the quantitative measurement of a compound or metabolites in biological fluids, primarily blood, plasma, serum, urine, or tissue extracts [8].

While the general acceptance on the potential application of CDs for bioanalysis is converging, the practical development involves various stages and also technology advancement before the final developed CDs can be adopted as a detection probe. In this case, crude CDs are only acting as signal tags to generate a measurable signal, such as in the form of optical or potential signals using a detector. In order to introduce the analytical characteristics, the CDs need to be first designed with functionality that enables its interaction with the analyte of interest. It can be just of the weaker Van der Waals forces, stronger ionic interactions or the formation of complex via covalent bonds. Such interactions eventually should trigger a change in the initial signal that is concentration dependent. Once this has been achieved, the signal change can be correlated with the concentration of the analyte into a calibration curve and used as model to predict unknown concentration of the analyte based on the change of signal recorded. Specifically, the interaction shall be targeting on small molecules found in the biological system for the bioanalysis application. It will be of further advantage for having the probe to be biocompatible and nontoxic, which will enable its direct in vivo application for real‐time monitoring.

This chapter will particularly discuss the advancements and approaches taken to design CDs into bioanalytical probes. The fundamentals of CDs will be first revealed and served as the foundation for the further discussion on the technology that can be applied to engineer the CDs into a sensing probe. The engineering approaches will be limited to only those that can lead to potential development into a biocompatible probe. In this case, it can be generally divided into two; the advancement on the unique features of the CDs into better suiting the bioanalysis applications and the surface modification with biomolecules to introduce specific interaction of the CDs toward a targeted bioanalyte in the body. The chapter will then discuss the sensing mechanisms that can be adopted with developed CDs, revealing the working principle that causes the change in the initial signals. Most of the transduction systems that are reported to date have adopted optical method, since the CDs show unique photoluminescence (PL) property. In addition, optical method seems to be very practical and involves less sophisticated instrumentation. Some successful examples reported in the literature on the applications of CDs as bioanalysis probe targeting on specific bioanalyte are also included in this chapter. The future prospect of CDs is very promising for their practicality, as real bioanalysis probes and continuous improvements of the system are required to address some of the existing limitations of CDs. These include the identification on the origin of the optical features, consistent protocol for homogenous production of CDs, enhancement of quantum yield (QY) and also improvement of analytical sensitivity.

9.2 Fundamentals of CDs

9.2.1 Synthesis Approaches

Various synthetic strategies to access CDs have been explored progressively ever since the first report of CDs in 2004 by Xu et al. [9]. The different methods developed are typically classified into two broad categories: top‐down and bottom‐up approaches. Top‐down synthesis usually involves cleaving or breaking down of large carbonaceous precursors into nano‐sized particles that eventually form the CDs. Conversely, bottom‐up synthesis primarily involves seeding of atomic or molecular organic precursors to grow into the form of CDs [10]. Both approaches can involve either single or in combination of chemical, physical, or electrochemical processes.

In the top‐down approach, laser ablation is one of the earliest methods pioneered by Sun et al. in the preparation of CDs from a carbon target made of graphite powder and cement [11]. The CDs obtained were varying in sizes and the PL could only be detected upon performing the surface passivation. CDs can also be isolated from the arc discharge technique and one simple example will be the isolation of CDs from carbon soot collected of a burning candle [12, 13]. Most of the CDs isolated from this method tend to show relatively lower QY. The enhancement of the luminescence property, however, can be made via surface passivation, such as by refluxing in strong acid. Another top‐down method is the electrochemical exfoliation that involves electrochemical etching of carbon precursor through applied potential to obtain CDs, which was first reported by Zhou et al. [14]. The images of the CDs obtained from high‐resolution transmission electron microscopy (HRTEM) have revealed that the sample is of high density, in homogenous spherical shape, and narrow size distribution. Despite the different routes employed during the synthesis, the formation mechanism of the CDs generally relies on breaking down bulk carbon sources into physically small carbon targets [15].

Generally, top‐down approaches involve sophisticated instrumentation and considerably pure carbon targets, while the bottom‐up routes are comparatively more straightforward and can be achieved using basic laboratory apparatuses and facilities. One of the common examples is the carbonization of organic precursors using strong mineral acid to form the CDs. The work by Peng and Travas‐Sejdic utilized strong sulfuric acid to dehydrate carbohydrates into carbon nanoparticles, and the sample was further treated with nitric acid to obtain fluorescent CDs [16]. An apparent limitation of such methods is the difficulty to thoroughly remove the excess acid while retaining the intrinsic PL properties of the CDs. Similar to acid carbonization, direct thermal decomposition can also be adopted to produce CDs. For instance, Bourlinos et al. [17, 18] produced surface functionalized CDs with average size of 7 nm from single‐step thermal decomposition of different ammonium citrate salts. These CDs were reported to be readily dispersible in various aqueous and organic solvents. The bottom‐up approach can be carried out economically and simply using conventional furnace and oven, while it also serving the motivation for “greener” synthesis since the process involved no toxic chemicals or organic solvents [1922]. Furthermore, the formation mechanism of CDs is related to the building up of CDs due to nucleation, condensation, and polymerization of the selected starting precursors [2326]. Under harsh conditions such as in the presence of strong acids or at elevated temperature, organic precursors can undergo bond breakage to form reactive monomers. These reactive monomers can then proceed with extensive cross‐linking reactions and formation of inter‐chain imide bonds to promote CDs build‐up [27]. Table 9.1 summarizes a list of recent examples of fluorescent CDs synthesized from different strategies and some of their respective physical properties.

Table 9.1 Some recent examples of CDs prepared from different synthetic strategies and their properties.

Synthesis method Precursor Dopant PL Color Maximum emission wavelength (nm) Size (nm) QY (%) Reference
Laser ablation Graphite powder N‐doped Green 543 2.87 ± 0.02 6.5 [28]
Toluene Blue to green 370–500 1.3–4.0 8.7 [29]
Arc discharge Flaxseed oil Green 523 2–10 ~5
N,N‐dimethylformamide Blue 420 2–6 3.0 [30]
Electrochemical exfoliation Amino acids N‐doped Blue 430 2.95 ± 0.12 46.2 [31]
1‐propanol Blue 450 3–4 [32]
Chemical oxidation Citric acid, thiourea S, N‐doped Red 594 4.42 ± 0.67 22 [33]
L‐glutamic acid Blue 437 5.42 ± 0.98 6.3 [34]
Micro‐plasma treatment Citric acid, ethylenediamine Blue 450 2.51 ± 0.49 [35]
Citric acid, ethylenediamine N‐doped Blue 450 1–3 5.1 [36]
Acid dehydration Activated carbon Green 513 8–10 [37]
Sucrose Yellow 524 ~10 [1]
Thermal decomposition Chicken egg white Blue 395 3.3 ± 0.4 43 [38]
Mangosteen pulp Blue 440 ~5 [39]
Aminophenylboronic acid B,N‐doped Green 520 1.5–4.0 1.6 [40]
Hydrothermal treatment Microalgae Blue 435 1–8 4.3 [41]
Chocolate Blue 354 ~6.41 [42]
Citric acid, polyethylene‐polyamines Blue 436 1–3 38 [43]
Microwave irradiation Trisodium citrate solution, sodium thiosulfate S‐doped Blue 425 ~5 58 [44]
Citric acid, ethylenediamine N‐doped Blue 445 2–7.6 80 [45]
Lysine Blue 432 5–10 23.3 [46]
Ultrasonication Glucose Yellow 516 3–5 [47]
Glucose Green 500 ~2.7 [48]
N‐methylethanolammonium thioglycolate N,S‐doped Blue 450 3–8 12.5 [49]
Templated route Ethylenediaminetetraacetic acid (EDTA) N‐doped Blue 420 2.3 ± 0.3 1–5 [50]
Glucose Blue 408, 419, 430 1.5–4.5 3.0 [51]

9.2.2 Optical Properties

9.2.2.1 Absorbance and Photoluminescence (PL)

Ultraviolet and visible (UV–Vis) spectroscopy is an essential tool that has been widely applied to study the electronic energy transition and structural properties of CDs. It is typical to observe CDs displaying strong absorption band in the UV region of 260–320 nm, with a tail extending into the visible range [52]. The absorption in the UV range is often suggested to be ascribed to the π π * transition of C=C bond and n π * transition of C=O bond [5355]. UV–Vis absorption peaks at longer wavelength were also reported at 340 and 451 nm [56]. The former spectrum is attributed to the n π * transition whereas the absence of a well‐defined band edge in the UV–Vis range was suggested for the latter broad absorption spectrum that extended to the visible range.

The unique PL characteristics of CDs include the effect of excitation wavelength (λex) on the emission wavelength and intensity of CDs. The λex‐dependent emission property of CDs is commonly reported [31, 34, 57]. A tentative explanation of such optical property has been proposed as to be due to the surface states created on the CDs [58]. The complex surface defects could be generated by surface oxidation to trap more excitons during the excited stage. This would subsequently result in radiative recombination of the trapped excitons to generate fluorescence emissions corresponding to the excitation energy, hence red‐shifting the emission wavelength. Different emissive states and heterogeneous types of π conjugations in nanodomains are also suggested to be responsible for the λex‐dependent PL [59, 60]. In contrast, λex‐independent fluorescence emission of CDs has also been observed [61], which is often claimed to be attributed to a lower degree of surface oxidation and surface uniformity [58, 62]. Such λex‐independent emission generally agrees with Kasha's rule, where the emission is originated from the ground vibrational level of the excited singlet state [44, 56].

In addition to λex, pH of the aqueous solution is reported to alter the fluorescence emission wavelength and intensity of CDs. One of the earliest reports has shown that maximum emission of CDs produced from ascorbic acid had red‐shifted from 441 to 550 nm when the pH was increased accordingly [63]. Zhang et al. also reported pH‐induced shift in emission wavelength of CDs from the blue (490 nm) to green (523 nm) region [64]. Moreover, pH change in aqueous solutions of CDs would affect the emission intensity due to the presence of surface functional groups such as carboxyl and amino groups [65, 66]. Wang et al. have observed that CDs exhibited smaller diameter at low pH and were readily dissolved in aqueous solution. However, when at high pH, CDs possess comparatively larger diameter and pH‐induced aggregations could occur and eventually lead to quenching in fluorescence [67].

Although quite a few studies have reported bright blue PL of CDs, their application in biological systems is usually restricted due to blue auto‐fluorescence of biological matrix. This reduces the contrast between the area tagged with CDs and the biological matrix. Besides, there is a potential of photodamage to biological tissues upon excitation of the CDs under UV light [68]. Hence, production of CDs with emission at the longer wavelength and of multicolor is greatly desired for applications dealing with in‐vivo biological environment. Precise control of the synthesis conditions such as reaction temperature and time could effectively produce CDs with two distinct fluorescence properties from the same precursor [2]. Multicolored CDs with relatively high QY of 41% (blue), 43% (yellow), and 39% (red) emission have been fabricated via solvothermal synthesis in dimethylformamide as solvent using citric acid and urea as carbon (C) and nitrogen (N) sources [69]. Besides, different precursors can also tune the color emission of the CDs. CDs with blue to near‐infrared (NIR) emissions, as shown in Figure 9.1 within the range of 480–680 nm have been prepared from various functional precursors based on polythiophene derivatives [70]. Besides starting precursors, the emission of CDs can also be adjusted by controlling the charge transfer process, as demonstrated by Hu et al. [71]. The emissive states of the CDs were tuned by grafting the CDs with molecules consisting of a benzene ring and different heteroatom substituents through peptide bonds. The resulting CDs exhibited bright blue, green, and yellow emission under UV excitation. Interestingly, a recent study has reported the solvatochromism effect where CDs emitted different PL colors in the visible region from green to red when dispersed in various solvents [72]. These CDs were claimed to display λex‐independent emission.

Synthetic route of CDs with 3 skeletal formulas for HOOC, PT1, and PT2 (left) having a right arrow labeled Hydrothermal (top side) and Carbonization (bottom side) pointing to circles of various shades labeled C-dots.; Image displaying the emission colors of various CDs under UV light exposure with 7 tube-like shapes labeled C-dots1, C-dots2, C-dots3, C-dots4, C-dots5, and C-dots 6 (from left to right).

Figure 9.1 (a) Synthetic route of CDs. (b) Emission colors of various CDs under UV light exposure. (See color plate section for the color representation of this figure.)

Source: Reprinted with permission from Ref. [70].

9.2.2.2 Quantum Yield (QY)

QY is one of the vital parameters measured to characterize the PL property of CDs. The CDs produced during the early stage of development have typically shown relatively low QY that is less than 10.0% [7375]. It was later found that the QY of CDs can be enhanced by doping of heteroatoms such as oxygen (O), sulfur (S), and nitrogen (N) into the CDs [7678]. When compared to the bare CDs obtained from hydrothermal treatment of citric acid (QY = 5.3%), the QY of N‐doped CDs and N,S‐co‐doped CDs were significantly enhanced to 16.9% and 73.0%, both using quinine sulfate as standard [79]. The work by Liu et al. that prepared N‐doped CDs in a confined layered‐double hydroxide (LDH) host claimed that the QY of CDs could be tuned by adjusting the charge density of the host layer and achieved maximum QY of 61.6% [80]. Recent advances in CDs synthesis have successfully prepared CDs with ultra‐high QY of up to 99% via simple one‐step microwave‐assisted synthesis without the need of additional step of surface passivation [81]. In addition to the aforementioned PL properties of CDs, nonblinking nature and excellent photostability are also the merits of CDs that have inspired tremendous research in this field of study compared to QDs.

9.2.2.3 Photoluminescence Origins

Intensive research has been focusing on unveiling the origin of the PL from CDs; as such knowledge is imperative for fine‐tuning the PL properties to suit a specific application. Although the exact origin of PL is still debatable, several general mechanisms explaining the root of the PL event have been suggested. These include quantum confinement effect, surface states, and molecular origins. Quantum confinement effect of π‐conjugated domains of carbon core can be referred to nano‐sized particles with diameter smaller than the exciton radius, causing the charge carriers to become spatially confined, leading to the rise in energy levels [82]. Through a series of spectroscopic analyses and theoretical calculations (Figure 9.2), Li et al. convincingly showed that the size‐dependent‐PL of CDs obtained from electrochemical approach was solely attributed to the quantum‐sized graphitic structure of CDs [83]. In addition to the particles of different sizes, quantum confinement of a distribution of different emissive energy traps on CDs upon surface passivation was also suggested [11, 14].

Left: eV vs. pasticle size/mm displaying box markers fitted on the descending curve. Right: Gap/C6H6 vs. diameter/A displaying a descending curve with 5 star markers labeled C6H6, C24H12, C54H18, C96H24, and C150H30.

Figure 9.2 (a) Optical images of typical CDs illuminated under white (left). and UV (right). light; (b) PL spectra of CDs where red, black, green, and blue lines are the PL spectra for CDs emitting blue, green, yellow, and red fluorescence, respectively; (c) relationship between the size of CDs and their electrochemical properties; (d) HOMO‐LUMO gap dependence on the size of graphitene fragments. (See color plate section for the color representation of this figure.)

Source: Reprinted with permission from Ref. [83].

There are also some suggestions that PL of CDs is due to the synergistic effect of the graphitic structures in the carbon core and the well‐distributed surface states [84, 85]. Regarding the surface states theory, the PL of CDs is claimed to be originated from the functional groups on the surface of the CDs. The presence of functionalities with O and N is responsible for the λex‐independent emission of CDs due to the different induced vibration relaxation modes [86]. Besides, doping with heteroatoms, especially the N atoms, can improve the emission efficiency of the PL with higher QY. A study by Yuan et al. has claimed that amino groups on the surface of CDs was governing the PL property and could enhance the conjugation degree of the CDs system [87]. As a result, it increases the electron transition probability from the ground state to the lowest excited singlet state that leads to higher QY.

Unlike the surface state where the PL origin was formed by hybridization structure of functional groups and carbon core, molecular‐state‐induced PL is solely due to the organic fluorophore attached on the surface or interior of the carbon backbone [88]. For instance, the dehydration of carbohydrate molecules can produce 5‐hydroxymethylfurfural (HMF) derivatives that serve as aggregated fluorescing units within the CDs, causing it to emit orange‐red fluorescence [89]. Others have suggested the generation of H‐aggregate type excitonic states that has caused the PL of CDs [90, 91]. In a recent study, PL mechanism of multi‐states has been reported for N‐doped CDs obtained via microwave‐assisted synthesis [92]. Treatment at 160 °C produced organic molecule clusters that were rich in amide bonds with molecular states being the PL center. Increasing treatment temperature to 200 °C has facilitated the formation of carbon core, and the PL was attributed to the synergy between the carbon core and surface states; whereas prolonged treatment resulted in PL arise from carbon core.

9.2.2.4 Up‐Conversion Photoluminescence (UCPL)

In addition to the usual PL (sometimes also known as down‐conversion fluorescence), up‐conversion photoluminescence (UCPL) of CDs has also been reported. UCPL refers to the phenomenon where the fluorophores could absorb two or more low‐energy photons and emit high‐energy photons at wavelengths shorter than the excitation wavelength [93]. The UCPL of CDs was suggested by Cao et al., who first observed that CDs emitted visible light upon excitation at NIR (800 nm) using a femtosecond pulsed laser [94]. Following that, UCPL is also reported in CDs obtained via alkali‐assisted electrochemical method [83], thermal decomposition [63], microwave‐assisted pyrolysis [95], etc.

The UCPL is utilizing NIR light as the excitation light source and this offers intriguing advantages especially in the field of biological applications. The benefits of UCPL include noninvasiveness, deep‐tissue penetration, reduced damage caused by photon, low photobleaching, and weak auto‐fluorescence background [96]. In a recent study, UCPL of a nontoxic CDs‐embedded hybrid microgel with good‐biocompatibility was demonstrated for glucose sensing at physiological pH [97]. This has shed light on applying CDs with UCPL property for theranostics purposes. UCPL property is commonly believed to be accounted by multiphoton active process [83, 94]. However, Shen et al. later claimed that multiphoton active process is insufficient to fully explain the mechanism of the UCPL of CDs and therefore speculated the theory of anti‐Stroke transition based on their spectral observation [98]. They have proposed that the UCPL of their CDs was due to anti‐Stokes PL, which involved the process of electrons relaxing back to σ orbital from a high‐energy state when the electrons in π orbital were excited by low‐energy photons.

It is worth mentioning that several reports have claimed that the UCPL observed in CDs systems is actually originated from the equipment artifact. The UCPL is the normal fluorescence excited by light radiation from the second‐order diffraction (λ/2) passing from the excitation monochromators of fluorescence spectrophotometer [99]. This type of misinterpreted PL due to the leaking light radiation can be eliminated by inserting a long‐pass filter in the excitation pathway of the spectrophotometer [100]. Additionally, the interference of second‐order diffraction light can also be eliminated by using a pulsed laser as an excitation source instead of a xenon lamp [101]. Therefore, UCPL of CDs observed by employing laser as the excitation source coupled with quadratic dependence relationship between the two‐photon luminescence intensity and the excitation laser power indicated the genuine UCPL [101]. Nevertheless, the UCPL property of CDs must be carefully evaluated, taking into consideration that second‐order diffraction light (λ/2) might lead to misinterpretation of normal fluorescence as UCPL.

9.2.2.5 Phosphorescence

Phosphorescent materials have been widely used in optoelectronic devices, sensors, and security systems over the past few decades. However, research on room temperature phosphorescence (RTP) of CDs is still at its infancy due to the relatively scarce reports in this area. In a pioneering work by Deng et al., the RTP has been observed for the first time from the CDs obtained via pyrolysis of ethylenediaminetetraacetic acid disodium salt (EDTA‐2Na) that were dispersed into a polyvinyl alcohol (PVA) matrix under UV light excitation [102]. The phosphorescence centered at approximately 500 nm and had a lifetime of 380 ms. Subsequently, RTP has also been observed when CDs were dispersed in different matrices such as potash alum [103], polyurethane [104], LDHs [105], and silica gel [106]. The CDs embedded in silica gel matrix displayed bright green after‐glow had a phosphorescence lifetime of about 1.8 seconds, which was claimed to be the highest magnitude reported for CDs in solid‐state matrices thus far [106].

Apart from single‐component matrices, composite matrices have also shown to activate the RTP of CDs. In one example, N‐doped CDs that were integrated into composite matrices consisting of the recrystallized urea and biuret from the heating of urea has exhibited the RTP [107]. This is due to the more rigid microenvironment of the CDs and the enhanced hydrogen bonding between biuret and CDs that promoted the RTP. The composite matrices have protected the triplet state from a long lifetime quenched by suppressing the nonradiative deactivation processes. The two‐component matrices were claimed to be more superior when compared to single‐component matrices due to the combination of the individual ingredient role [107]. RTP is generally observed when CDs are incorporated onto solid support. However, a ground‐breaking study by Yan et al. reported RTP of water‐soluble CDs in pure aqueous solution without any supporting matrices [108]. The RTP of the functionalized‐CDs could be quenched by ferric ions (Fe(III)) and effectively recovered in the presence of phosphate‐containing molecules such as adenosine‐5ʹ‐triphosphate (ATP).

In terms of potential applications, a novel RTP logic gate based on nucleic acid functionalized CDs and graphene oxide (GO) has been demonstrated [109]. Complementary DNA (cDNA) was first conjugated with CDs (cDNA‐CDs) through carbodiimide chemistry prior to adsorption onto a GO surface by π−π stacking interactions, yielding a cDNA‐CDs/GO complex. RTP of the cDNA‐CDs conjugates was “turned off” upon adsorption onto the GO surface due to photo‐induced electron transfer (PET) from CDs to GO. Subsequent addition of mercuric ions (Hg(II)) or transfer DNA (tDNA) that is complementary to the cDNA could turn the RTP of CDs back “on.” Another study by Chen et al., who demonstrated aggregation‐induced RTP in N‐doped CDs powder hydrothermally synthesized from PVA and ethylenediamine, found potential application in temperature sensing [110]. Interestingly, RTP of the CDs could only be observed in aggregation state but not in aqueous solution due to the postulate that the CDs themselves could serve as both the solidified host and the luminescent guest. The RTP intensity decreased in double exponential decay profile as the temperature increased from 30 to 90 °C; hence, phosphorescence intensity variation could be predicted within this temperature range by fitting relevant equations.

Suppression of radiationless relaxation processes is vital for the phosphorescence mechanism [105]. RTP of CDs is generally claimed to be originated from the triplet excited states of aromatic carbonyl (C=O) groups formed at the surface of CDs. Hydrogen bonding between the solid matrices and CDs could rigidify the C=O bonds on the surface of CDs, and this, in turn, minimizes the nonradiative transitions of triplet excitons and hence promotes phosphorescence [102, 104]. On the other hand, Li et al. argued that the C=N bonds on CDs surface were responsible for the RTP of CDs [107]. The C=N bonds were believed to create a new energy level structure and promote the generation of triplet excitons through n−π* transition. In a recent report, covalent bonds fixation of CDs onto colloidal nanosilica solution was evaluated to be a better alternative than hydrogen bonds due to the extension of long afterglow to solution forms. The presence of unsaturated bonds such as C=C bonds on CDs surface could make them self‐protecting agents from quenching effects as a result of energy transfer from excited triplet states of the chromophore to the dissolved oxygen [111].

9.2.3 Physical and Chemical Properties

CDs are zero‐dimensional, quasi‐spherical nanoparticles with a typical diameter of less than 10 nm, and the carbogenic cores can be amorphous or nanocrystalline [52]. In terms of morphology, scanning electron microscopy (SEM), transmission electron microscopy (TEM), and atomic force microscopy (AFM) can be employed to provide such information. CDs are commonly described as homogeneous, monodisperse nanoparticles with narrow size distribution [32, 37]. Besides, HRTEM, X‐ray diffraction (XRD), and Raman spectroscopy are powerful tools to characterize the structural properties of CDs. Several structural models such as diamond‐like, graphite/graphite oxide, and amorphous structures are often adopted to describe the CDs carbogenic cores. For instance, the selected area electron diffraction (SAED) pattern in TEM image of CDs obtained through laser irradiation indicated a diamond‐like structure [112]; the broad XRD diffraction peak at 2 θ = 22.6° was attributed to amorphous carbon structure [53]; signals of Raman spectrum at 1578 and 1346 cm−1 that corresponded to G and D band, respectively, indicated high degree of graphitization [33].

The carbogenic cores of CDs are mainly composed of C, H, O, and N elements. Microanalysis of the elemental composition and speciation of CDs can be achieved by Fourier transform infrared spectroscopy (FTIR), nuclear magnetic resonance (NMR), and X‐ray photoelectron spectroscopy (XPS). FTIR spectroscopy is widely employed to characterize the surface functional groups that are present on CDs. Carboxylic moieties (─COOH) and hydroxyl groups (─OH) are commonly found on the surface of CDs [112, 113]. CDs usually have high oxygen content due to the oxidized carbogenic cores as well as the presence of surface functional groups. For example, the elemental analysis of the oxidized CDs purified from candle soot have been revealed to have the composition of 36.8% C, 5.9% H, 9.6% N, and 44.7% O, which compares to raw candle soot that consists of 91.7% C, 1.8% H, 0.1% N, and only 4.4% O [12].

The chemical properties of CDs can also be identified by NMR measurements. The 1 H NMR spectrum can indicate the presence of sp3 C─H protons, aromatic or sp2 protons as well as protons attached with hydroxyl, ether, carbonyl, and aldehyde groups [23]. Meanwhile, their 13 C NMR spectrum indicated the presence of sp3 carbons and carbons attached with hydroxyl groups and ether linkages, aromatic C=C and C=O carbons.

Furthermore, a typical XPS as shown in Figure 9.3 by Miao et al. indicated the four peaks at approximately 284.3, 285.1, 286.1, and 288.5 eV were corresponding to sp2 hybridized carbon atoms in C─C/C=C, sp3 hybridized carbon atoms in C─S/C─N/C─O, ─CN, and ─COOH groups [33].

Image described by caption and surrounding text.

Figure 9.3 (a) Full survey X‐ray photoelectron spectroscopy (XPS): (b) C 1s, (c) N 1s, and (d) S 2p XPS spectra of S,N‐co‐doped CDs.

Source: Reprinted with permission from Ref. [33].

9.2.4 Biosafety Assessments

CDs are well known for their good biocompatibility and relatively nontoxic nature, which is also a prerequisite for bioanalysis application. The biotoxicity of CDs has been comprehensively studied both in vitro and in vivo. In a pioneering work by Sun et al., the application of CDs for optical imaging of biological cells in vitro was first demonstrated using human epithelial colorectal adenocarcinoma (Caco‐2) cells [11]. The result showed that CDs could be readily taken up by the cells, indicating excellent biocompatibility. This report later inspired considerable amount of studies to further explore the cytotoxicity of CDs for bioapplications. Cytotoxicity of CDs has been evaluated in various cell lines such as human cervical cancer (HeLa) cells [31], murine colorectal carcinoma (CT26.WT) cells [34], liver hepatocellular carcinoma (HepG2) cells [114], human umbilical vein endothelial (HUVEC) cells [115], human breast adenocarcinoma (MCF‐7) cells [116], and macrophages (RAW264.7) [117]. All in vitro studies have indicated that CDs exhibited none or low cytotoxicity.

Other than in vitro cytotoxicity, in vivo studies have also been widely investigated. Optical imaging in vivo utilizing CDs as fluorescent probes was first demonstrated by Yang et al. using mice as animal models [118]. CDs were introduced into female DBA/1 mice through subcutaneous and intravenous injection. The result showed that none of the animals exhibited any obvious sign of acute toxicological responses and the CDs injected intravenously were eventually excreted via urine [118]. In addition to mice models, in vivo studies have also been reported in guppy fish [119], genetic worms (Caenorhabditis elegans) [120], and fruit flies (Drosophila melanogaster) [121]. Toxicity in both plants and animals has also been conducted using bean sprouts and mice models [122]. All results suggested that CDs possess negligible toxicity.

In vivo biodistribution and tracking of CDs are particularly important in terms of clinical translation. Tao et al. employed a radiolabeling method by tagging the CDs with Iodine‐125 (125 I) to track the in vivo behavior of CDs [74]. It was realized that CDs mainly accumulated in the reticuloendothelial system (RES) such as liver and spleen as well as in kidney. Moreover, CDs would eventually clear out of the body through both renal and fecal excretions. In another study, fluorescence dye ZW800‐tagged CDs were tracked ex vivo and in vivo using mice models [123]. Ex vivo imaging revealed that CDs mainly accumulated in kidney and only small amount accumulated in liver an hour after CDs injection. CDs were found rapidly excreted from the body after injection through different routes, in which the clearance rate was ranked as intravenous > intramuscular > subcutaneous.

9.3 Bioengineering of CDs for Bioanalysis

9.3.1 Functionalization Mechanism and Strategies

9.3.1.1 Chemical Functionalization

Surface modification via chemical route is practical for tailoring the properties of CDs with desired characteristics to suit a specific application. Sometimes, it is also known as surface passivation or functionalization. The approach can be performed via various types of chemical interactions such as covalent bonding, π−π conjugation, and electrostatic interaction. Among these, the most popular one is the covalent linkage formed using the carbodiimide crosslinking chemistry. This technique is targeting on the activation of carboxylic group that is commonly found on the surface of CDs, in which the activated carboxylic group later can react with its conjugate such as amine group to form a covalent bond. The carbodiimide chemistry involves two‐step reactions where the first is activation of carboxyl groups (─COOH) and the second is peptide bond formation as a result of nucleophilic attack by primary amines (─NH2) [124, 125]. The new bond formation is often proven by the observation of new absorbance shown in the infrared spectrum. For instance, FTIR spectrum by Fu et al. showing the stretching vibration of N─C=O at 1651 cm−1 has confirmed the formation of amide bonds between ─COOH groups in bare CDs and ─NH2 groups in arginine (Arg) via carbodiimide chemistry [126]. It is noted that no part of the chemical structure of the carbodiimides is being added to the final product, making them an efficient zero‐length carboxyl‐to‐amine cross‐linker. In some cases, N‐hydroxysuccinimide (NHS) will be added to act as a stabilizer of the intermediate formed by 1‐ethyl‐3‐(3‐dimethylaminopropyl)carbodiimide hydrochloride (EDC). Upon quenching the reaction, the conjugated product is often isolated via various purification options such as column chromatography or dialysis to remove the urea byproducts and excess reagents. The carbodiimide crosslinking is typically favored to conjugate antibody or other protein compounds due to the presence of amine and carboxyl groups.

In addition to the aforementioned chemical interactions, weaker interaction forces can also be employed for the surface conjugation. This can be due to the surface property that is likely to induce attractive forces to some ligands or polymeric chains. For instance, some research reports the electrostatic interactions between negatively charged CDs with positively charged polyethylenimine (PEI) and folic acid (FA) in one of the surface modification efforts [127]. Furthermore, the presence of C=C as shown by the sp2 hybridization of CDs has given the nanoparticles leverage for π−π conjugation with desired compounds that are also present with conjugated system. For instance, doxorubicin (DOX), an anticancer drug that contains conjugated system within the chemical structure has been successfully conjugated to CDs through both π−π conjugation and electrostatic interactions for controlled drug release [128]. This characteristic is also useful to functionalize CDs with small molecule dyes that are typically rich with π‐conjugated domains such as rhodamine B [129].

9.3.1.2 Doping

Heteroatoms doping is generally considered as an effective approach to tune the band gap of semiconductors to achieve a particular electrical property of interest. This is by introduction of these heteroatom impurities to the semiconductors intentionally to alter the initial band gaps [130]. Heteroatoms are defined as elements other than C and H. In the case of CDs, doping has been also demonstrated to be efficacious to fine‐tune the properties of CDs. For examples, N‐doping could notably enhance the QY of CDs [77, 80]. Doping with boron (B) has significantly improved the PL intensity and nonlinear optical response of CDs [131, 132]. The presence of S atoms in the carbon precursor has played a vital role in promoting the formation of doped CDs from low‐molecular‐weight precursors [133]. Phosphorus (P)‐doped CDs exhibited aggregation‐induced red shift emission (AIRSE) from blue to orange yellow as schematically illustrated in Figure 9.4 [134].

Image described by caption and surrounding text.

Figure 9.4 Schematic illustration on the synthesis of P‐CDs and the induced effect due to AIRSE.

Source: Reprinted with permission from Ref. [134].

In addition to nonmetal elements, metal dopants such as zinc (Zn) [135], copper (Cu) [136], cobalt (Co) [137], manganese (Mn) [138], germanium (Ge) [139], and terbium (Tb) [140] have also been reported for CDs. Doping with gadolinium (Gd) endowed CDs to be a potential magnetic resonance imaging (MRI) contrast agent for imaging and radiotherapy of tumors [141, 142]. Moreover, co‐doping provides an effective means to further improve the properties of CDs. Some examples include CDs co‐doped with B, N, and S that later can be fabricated for colorimetric and fluorescent dual mode detection of Fe(III) ions [143] and selenium (Se) ion. Besides, N‐co‐doped CDs exhibited bright green emission with high QY (~52%) evidenced high performance as a dye for fluorescein fundus angiography (FFA) [130].

9.3.1.3 Coupling with Gold Nanoparticles

Gold nanoparticles (AuNPs) have been used for centuries in arts due to their vibrant colors produced as a result of their interaction with visible light. These unique optoelectronics properties have then been studied and exploited for various biomedical applications such as sensory probes, therapeutic agents, drug delivery, etc. The optical and electronic properties of AuNPs can be tuned by altering the size, shape, surface chemistry, or aggregation state. More recently, research on these nanoparticles is extended by the integration with other nanoparticles such as with the CDs for biosensing applications [144, 145]. The CDs have showed an enhanced PL in the presence of AuNPs and this is explained as due to the two synergistic mechanisms of electric field and the induced surface plasmon effects [146]. The same study has also reported that the consistent distance in the proximity of less than 10 nm between CDs and AuNPs as due to the linked polymer, N‐(β‐aminoethyl)‐γ‐aminopropyl methyl‐dimethoxy silane (AEAPMS) has exposed the nanoparticles with increased electric fields. Such effect has improved the local electromagnetic field and, in turn, strengthened the PL of CDs. Simultaneously, the partial transfer of energies from excited state of CDs to AuNPs surface plasmons has induced surface plasmon effects and subsequently increased the PL of CDs. This study has shown that the two nanoparticles can support and enhance each other's properties as a result of coupling reaction. The simplest coupling or modification with AuNPs was by mixing the precursors for both CDs and AuNPs in one‐pot synthesis. A study by Wu's group has utilized citric acid and cysteine as a C source for CDs, which these precursors have concurrently served as reducing agent and could easily form the graphene framework through intermolecular dehydration [147]. Another method of coupling with AuNPs would be by assembly via electrostatic interactions between the AuNPs and CDs. In such cases, both nanoparticles were synthesized individually and assembled in subsequent step. For instance, Wang et al. synthesized CDs via hydrothermal treatment of pancreatin, while AuNPs were obtained from its precursor HAuCl4 followed by modification with aptamer [148]. The electrostatic interactions were formed upon mixing the two nanoparticles in 10 mM phosphate buffer saline (PBS) buffer.

9.3.1.4 Fabrication onto Solid Polymeric Matrices

The blending of CDs into various solid matrices through sol–gel technology has also attracted immense attention due to the materialization of CDs for real bioanalytical device applications. Molecularly imprinted polymers (MIPs) can be prepared by simulating receptor active sites using target molecules as templates, from which the template molecules will later be removed after the formation of the synthetic polymer [149]. CDs were successfully integrated into a silica matrix through sol–gel based molecular imprinting technique for fluorometric determination of nicotinic acid [150]. Besides, CDs have also been successfully incorporated into other types of polymeric matrices such as agarose hydrogel [151], chitosan hydrogel [152, 153], poly(methyl methacrylate) (PMMA) [154, 155], PVA [156], and gel glass [157]. In a recent study, a polymer/silica hybrid film has been successfully fabricated to entrap CDs [158]. The transparent free‐standing hybrid film that possessed great flexibility could protect the embedded CDs even when treated at a high temperature (550 °C), and its potential in photonic application has also been demonstrated.

9.3.2 Biomolecules Grafted on CDs as Sensing Receptors

9.3.2.1 Deoxyribonucleic Acid (DNA)

Nucleic acids have been innovatively employed as capping agents and engineered onto the surface of CDs. DNA is a linear biopolymer made up of nucleotides consisting three molecular fragments – namely, a sugar, a phosphate group, and a nitrogenous base [159]. Nucleic acids conjugated CDs possess great potential for biomedical applications. For instance, genomic DNA isolated from Escherichia coli was successfully conjugated to CDs and employed as a fluorescent vehicle for cell imaging and drug delivery [160]. In another study, a hybrid hydrogel composed of phosphoramidate‐linked amine functionalized CDs and single‐stranded DNA (ssDNA) were loaded with DOX for sustained release of the drug [161].

A novel paper‐based electrochemiluminescence (ECL) origami device for detection of IgG antigen based on rolling circle amplification (RCA) using oligonucleotide functionalized CDs as nanotags has been reported [162]. Sensitive detection of 6‐mercaptopurine and Hg(II) ions utilizing N‐doped CDs conjugated with carboxyfluorescein (FAM)‐labeled DNA [163] as well as “off−on” detection of Hg(II) ions based on oligodeoxyribonucleotide (ODN) functionalized CDs and grapheme oxide have also been demonstrated [164]. Furthermore, several recent studies show that DNA conjugated CDs could be utilized as robust fluorescent probes for detection of microRNAs (miRNAs) based on different detection modes such as fluorescence resonance energy transfer (FRET) [165], and ECL [166].

9.3.2.2 Aptamers

Aptamers are oligonucleotide sequences that bind specifically to target molecules with high affinity. The specific sequences of aptamers can be generated by systematic evolution of ligands by exponential enrichment (SELEX) process [167]. These synthetic single‐stranded nucleic acids could form a well‐ordered tertiary structure that is capable of recognizing various targets such as inorganic ions, small molecules, proteins, and cells with high affinity and specificity. Although the selection of aptamers in the market is still not comprehensive, the demand is increasing these days due to their great potential in the biological applications. Thrombin aptamer has been commonly grafted on CDs for selected applications [168, 169]. For instance, Xu et al. constructed aptamer sandwich using two different thrombin aptamers; namely the TBA15 and TBA29, which could form an intramolecular G‐quadruplex that recognizes the different sites of thrombin [170]. TBA15 has been conjugated with silica nanoparticles while TBA29 has been covalently linked to CDs. These two surface engineered nanoparticles were assembled into a fluorescent sandwich structure bridged by thrombin. This enables the selective and sensitive detection of thrombin by referring to the correlation of the PL with the quantity of thrombin. In another separate study, a specific DNA aptamer (HB5) targeting human epithelial growth factor receptor 2 (HER2) was assembled on mesoporous silica‐CDs (MSCN)‐based system loaded with DOX for chemo‐photothermal combined therapy of HER2‐positive breast cancer cells [171]. Besides, carboxyl‐modified CDs conjugated with amino‐modified aptamers for qualitative detection of pathogenic bacteria, Salmonella typhimurium in egg and water samples have also been reported [172]. In addition, aptamer‐based CDs sensor have also been demonstrated for detection of Mucin 1 (MUC1) protein [173, 174], Aflatoxin B1 (AFB1 ) [148], carcinoembryonic antigen (CEA) [175], dopamine (DA) [176] as well as cancer cells such as MCF‐7 [177, 178], HeLa, and C6 glioma cells [179].

9.3.2.3 Proteins/Peptides

CDs tagged with proteins and peptides have also been studied more intensively as catalysts for metabolic reactions, DNA replication, responses to stimuli, biomolecule transport, and biological defense system. Proteins such as antibodies and enzymes are extensively studied due to their specificity and strong affinity toward respective antigens and substrates. CDs have been utilized as protein carrier to deliver enhanced green fluorescent protein (EGFP) into HeLa cells [180]. Circular dichroism spectroscopy could be employed to determine the composition and concentration of various protein‐carrier conjugates [181, 182]. There is a report on the success of linking bovine serum albumin (BSA) to CDs functionalized by NHS for actualizing fluorescence labeling of the protein [183]. Peptides such as the nuclear localization signal peptide (NLS) representing a short sequence of amino acid that can transport cargo proteins into cell nucleus through nuclear pore complex have also been conjugated with CDs. For example, the simian virus 40 (SV40) large T‐antigen NLS (PKKKRKVG) was employed as a targeting ligand and has been covalently conjugated to CDs through EDC/NHS coupling reaction for nucleus targeting cell staining [184]. Besides, Sharma group has successfully developed nanoliposomes containing cell penetrating peptides such as polyarginine (R9) and trans‐activator of transcription (TAT) conjugated with CDs for multiple applications including multicolor cell imaging, nucleus targeting, NIR‐triggered drug delivery, and NIR photothermal therapy (PTT) [185, 186]. Amino acids serving as primary building blocks for proteins are also favored in targeting studies due to their interactions with biological structures. For instance, arginine (Arg)‐conjugated CDs was reported with higher cellular uptake efficiency than bare CDs due to electrostatic binding between guanidyl group in Arg and the plasma membrane components in cells [126].

In addition to biomolecules sensing, CDs have also been assembled with enzyme to modulate specific enzymatic activity. For instance, porcine pancreatic lipase (PPL)/CDs hybrids were fabricated to modulate the catalytic activity of PPL [187]. When compared to free PPL, the activity of PPL/CDs was enhanced by 10% upon visible light irradiation; whereas the PPL/CDs activity was decreased by 30% in the absence of the light source. Nano‐bioconjugates of CDs and several different enzymes such as soybean peroxidase (SBP), Chromobacterium viscosum (CV) lipase, trypsin, and cytochrome c (cyt c) were also investigated for their feasibility to probe the enzyme location in water‐in‐oil microemulsion [188].

9.3.2.4 Biopolymers

Biopolymers are another option available to passivate the surface of CDs to further introduce the intended functionality and property to suit an application. Polyethylene glycol (PEG) with different molecular weights has been a popular functional polymer for surface engineering of CDs. PEG functionalized CDs are often claimed to exhibit enhanced fluorescence with excellent biocompatibility, which are suitable to serve as fluorescence probes for targeted cell imaging [189191]. For example, PEG‐passivated CDs were further functionalized with NLS as cancer cell nucleus imaging probes [184]. PEI is a well‐known high‐efficient transfection vector to mediate gene delivery [192]. Hence, it is not surprising that CDs surface engineered with PEI could effectively transfect genetic materials such as DNA and RNA into cells for cancer therapy [193195]. PEI‐anchored CDs has also been reported as sensitive probes for detection of sulfide anion (S2− ) [196], copper (Cu(II)), and Hg(II) ions [197] in biological fluids.

Engineering of different organic amino molecules onto the surface of CDs could alter their fluorescence properties as well as hydrophilicity and hydrophobicity [198]. Ethylenediamine (EDA) is one of the most commonly employed amino compounds to functionalize the surface of CDs with the purpose for bio‐labeling and fluorescence imaging of cells such as mesenchymal stem cells (MSCs) [191] and human gastric carcinoma cells (MGC‐803) [199]. CDs have also been successfully passivated with 1,6‐hexamethylenediamine (HMDA) that were conjugated with mouse anti‐α‐fetoprotein (AFP) antibody and goat anti‐mouse IgG to label HepG2 cells [200]. Bis(3‐pyridylmethyl)amine (BPMA) modified CDs could serve as chemosensor for glutathione (GSH) detection [201]. Besides, poly(amido amine) (PAMAM) dendrimers were assembled with CDs via noncovalent interactions and loaded with chemo‐drug epirubicin (EPI) for simultaneous intracellular imaging and drug delivery in cancer cells [202]. CDs encapsulated in biodegradable poly(lactic‐co‐glycolic acid) (PLGA) have also been fabricated as imaging agent [203].

Other than polyamines, organosilane such as AEAPMS has also been engineered with CDs for detection of quercetin [204] and Hg(II) ions [205]. The advantages of organosilane over typical polyamines as surface passivating agents are that the resulting organosilane‐passivated CDs could be directly fabricated into hybrid films or nanoliths without requiring any additional polymers. In addition, they possess high stability in various nonaqueous solutions and have the capability to be further fabricated into silica‐encapsulated nanoparticles [206]. Polyhedral oligomeric silsesquioxane (POSS), a hybrid molecule of silicon and oxygen, has also been employed as surface passivating agent for CDs. The nanocomposites of CDs and POSS was claimed to exhibit enhanced PL and thermal stability [207]. Moreover, the CDs functionalized with POSS have also been successfully utilized as cell imaging probes with high resistance to photobleaching and excellent PL stability in the presence of biological sample matrix [208].

9.4 Bioanalysis Applications of CDs

CDs that have been carefully designed and engineered with bioanalytical functionality can be then utilized for specific bioanalytical applications. Bioanalysis has become an important part of biomedical diagnosis as it involves quantification measurement of a drug compound or a metabolite in the biological fluids such as in blood, plasma, serum, urine, or tissue extracts. The excellent properties of CDs, such as good biocompatibility, low cytotoxicity, and rapid clearance from body have prompted many to switch focus from organic dyes or QDs to CDs for bioanalytical applications. These days, the scope of bioanalysis has even extended to the cellular studies via bioimaging, monitoring of the drug delivery process via intracellular tracking, and also the therapeutics including the photothermal or photodynamic therapy (PDT), and theranostics that integrated both diagnostics and therapy. All these will be discussed in greater detail in this section.

9.4.1 Biosensing Mechanism/Transduction Schemes

With the advancement and uprising field of biotechnology and nanotechnology, research efforts have been focusing on the development of biosensors as an alternative to bioassays that require observation on the effect of a substance toward a living cell or system. According to Turner, a biosensor is analytical device that uses biological components such as tissue, microorganisms, organelles, cell receptors, enzymes, antibodies, nucleic acids, natural products, etc., biologically derives material (e.g. recombinant antibodies, engineered proteins, aptamers, and etc.) or a biomimic – such as synthetic receptors, biomimetic catalysts, combinatorial ligands, and imprinted polymers [209]. The biological components will be conjugated or integrated within a physicochemical transducer or transducing microsystem. The transduction system maybe dealing with optical, electrochemical, thermometric, piezoelectric, magnetic, or micromechanical signal. The readings from transducing system will then be analyzed by correlation with the concentration of analyte being tested with the change in the signal caused by the presence of the analyte.

9.4.1.1 Fluorescence

Fluorescence is the luminescence process that takes place when a molecule or substance gets excited by a given energy and subsequently releases photon in the form of light when it relaxes back to the ground state. Fluorescence usually has a relatively short lifetime, meaning the relaxation process will happen very rapidly after the excitation has occurred. Typically, the emitted light is of lower energy compared to the absorbed light, where there will be an energy loss in the form of heat during the excitation and relaxation processes. Uniquely, CDs are showing fluorescence property when excited with sufficient amount of energy. This fluorescence property has enabled the CDs to be employed as an optical transducer for sensing applications.

Direct monitoring on the change in fluorescence property is the most common technique being applied for optical based detection. The fluorescence property is often affected by its surrounding parameters such as solvent types, existence of other impurities, temperature, etc. Therefore, any chemical or physical interactions of an external species with the fluorophore would eventually lead to changes in the property of fluorescence emission [10]. This can be in the form of intensity change or the shift in the emission wavelength and the change should be correlating with the amount of the effective external species that is present in the system. Intensity change is more commonly observed where it can be either quenching or enhancement of the emission signal in the presence of a targeted species. In the case of CDs as fluorescence species, most of the reported sensing system are based on fluorescence quenching, whether it is static quenching (ground‐state complex formation) or dynamic quenching (collisional or nonradiative energy transfer). Static quenching can be verified by the change in the PL profile of the fluorophore, typically the absorbance. For instance, the interaction between the analyte such as metal ions and CDs could result in formation of nonfluorescent ground‐state complex. The complex may absorb UV light and immediately return to the ground state without emission of photon, therefore causing a reduction in absorbance value [210212]. By understanding the quenching mechanism, the fluorophore can be exploited as sensor more efficiently and effectively. The effective quenching can be caused by metal ion such as Hg(II) due to the soft‐soft acid–base interaction between the S terminal on N and S co‐doped CDs. This has allowed more specific and selective detection of Hg(II) with minimal interferences from other metal ions [213]. The limit of detection (LOD) was found to be 0.05 nM within a linear dynamic range of 0–0.1 μM.

Dynamic quenching is due to the effective collision between the fluorophore and a quencher and resulted in a nonradiative energy transfer such as the Foster Resonance Energy Transfer (FRET). The process occurs from an excited donor fluorophore to a proximal ground‐state acceptor fluorophore within close proximity, typically 1–10 nm [214]. The FRET process highly depends on the degree of overlapping in the spectral between the fluorescence emission of donor and absorption of acceptor, meaning the acceptor must be able to absorb the energy from the donor. Recently, there has been increasing number of sensing systems that have adopted the FRET utilizing CDs with other nanoparticles forming nanocomposites. The close proximity within the nanocomposites construct can favor the FRET process, causing the turning off of the fluorescence that subsequently can be adopted for analytical applications. For example in a FRET‐pair system reported, two emissions were observed for CDs coated with vitamin B12: blue region (417 nm) attributed to CDs itself and red region (550 nm) due to vitamin B12 [215]. In such system, it was observed that the red emission was quenched while the blue emission of CDs increased accordingly in the presence of the sensing analyte, carbofuran phenol 3‐keto, which is a phenolic metabolite of carbofuran. The ratio of two intensities denoted as I417/I550 decreased linearly with the increasing amount of carbofuran phenol 3‐keto in the range of 9.8 μM to 14.0 mM with a limit of detection (LOD) of 12.2 μM. Besides ratiometric analysis, detection can also be made based on the change in intensity on a single wavelength due to FRET. This has been applied in a number of sensing systems namely CDs and manganese dioxide (MnO2) nanosheets for the detection of GSH [216]; CDs and AuNPs for the sensing of melamine; [217] and CDs with silver nanoparticles (AgNPs) for the detection of cysteine [218].

9.4.1.2 Chemiluminescence (CL)

CL is the emission of light as a result of a radiant relaxation from a fluorophore that has been excited from a chemical reaction [219]. Nanoparticles can typically take part in CL reactions as reductant, energy acceptor, luminophore, or catalyst. In many instances, the direct CL of CDs can be induced or enhanced by classical oxidants such as potassium permanganate and reactive oxygen species (ROS) [220, 221]. Specifically, there is a study that demonstrated that the direct CL of CDs can be enhanced by potassium ferricyanide, K3Fe(CN)6 under basic conditions, in greater manner as compared to other oxidants such as hydrogen peroxide (H2O2), KMnO4, and cerium (Ce(IV)) [222]. Ferricyanide can serve as the hole injector to convert CDs into positively charged radical. Meanwhile, superoxide anion radical formed by dissolving oxygen in a strong alkaline medium with CDs as reductant can act as an electron injector to produce negatively charged CDs radical. The CL emission of CDs can be interrupted by the presence of inorganic cations that will compete for the oxygenated radical intermediates. Among some of the metal cations and organic compounds studied, chromium, Cr(VI) was found to be one of the most effective quenchers with the efficiency of 44% while noradrenaline managed to quench up to 16%. This phenomena was exploited for the sensing of Cr(VI) and noradrenaline, with evaluated LODs of 0.004 and at 0.003 mg l−1 , respectively [222].

The CDs can also play the role as an acceptor in CL system for detection of cobalt, Co(II) when paired with electronic spinning resonance [223]. In this work, CDs were first passivated with PEG, followed by the modulation with cationic cetyltrimethylammonium bromide (CTAB) surfactant. This is to form polymer‐surfactant microenvironment that can amplify the CL of the Fenton‐like system. CDs have served as an energy acceptor from the radicals, causing strong CL at 535 nm upon its returning to the ground state. Co(II) was found to generate the strongest CL intensity on the system, and the enhancement has allowed the quantification of Co(II) even in liver hepatocellular carcinoma cells (HepG2). The strong enhancement of CL by CDs has also been applied for detection of DA, where DA was able to reduce effect on the formation of CL. [224]. The quenching of CL intensities that were initially enhanced by about 150 times in the presence of CDs was then linearly correlated with the concentrations of DA within the range of 2.5 nM to 20 μM with a LOD of 1.0 nM.

9.4.1.3 Electrochemiluminescence (ECL)

ECL is a redox‐induced optical phenomenon that involves high‐energy electron transfer between two luminous agents (emitters) or between a luminous agent and a coreactant generated at the electrodes that subsequently leads to a light emission [225]. The energy transfer between two emitters takes place based on an annihilation mechanism, whereby the application of oxidative conditions to a luminophore followed by reductive conditions (or vice versa) will generate high‐energy species that react with one another to produce ECL. In cases where the energy transfer occurs between emitter and a coreactant, both are first oxidized or reduced at the electrode to form radicals and intermediate states. The coreactant radical then oxidizes or reduces the emitter to produce its excited state that later emits ECL. The main advantage of having co‐reactant is that the formation of radicals is in aqueous phase and the subsequent ECL being generated is attainable without potential cycling and at less extreme potentials as compared to the common organic solvents [226]. This allows the ECL to be operated or applied for bioanalytical applications. In several studies, CDs acted as the coreactant that enhanced the ECL of the system [227230]. Particularly, Carrara et al. have reportedly adopted CDs as both the coreactant and nano‐carrier for the luminophor, ruthenium(II) tris(2,2ʹ‐bipyridyl) (Ru(bpy)3 2+ ) that have activated the “oxidation‐reduction” ECL mechanism [226]. The covalently linked CDs and Ru(bpy)3 2+ system was shown with self‐enhanced ECL due to intramolecular electron transfer reaction. This enhancement effect has been further applied for biosensing of natural products such as sophoridine on a solid‐state sensor platform [156]. The electrode has been modified with a coating of CDs embedded in PVA layer trapped with the Ru(bpy)3 2+ . While CDs have enhanced the ECL signal, the entrapment of Ru(bpy)3 2+ within PVA has yielded stable ECL signals that allowed the detection of sophoridine down to a LOD of 5.0 × 10−11  M. This probe can potentially be applied for the detection of clinical samples such as human serum. In a separate study, a dual‐peak ECL system from CDs in the presence of organic solution such as 0.1 M tetrabutyl ammonium bromide (TBAB) ethanol solution was found useful for metallic ions detections [231].

9.4.1.4 Electrochemical

CDs also portray some electrochemical properties, although this has not been widely explored for sensing as compared to optical method. Electrochemical sensors include three main types namely potentiometric, amperometric, and conductometric. Potentiometric sensors are based on the potential change on the working electrode in the presence of the targeted analyte compared to the reference electrode [232]. As for amperometric sensors, a potential will be applied across a reference and a working electrode to cause a redox reaction on the analyte that can generate current signal that is dependent on the analyte concentration. Lastly, conductometric sensors measure the conductivity of sample across the two electrodes. In recent years, CDs have been explored for their roles primarily in amperometric sensors. Canervari et al. optimized an electrochemical oxidation process for 1‐propanol to form nanocrystalline CDs that were later fabricated onto glassy carbon electrode (GCE) for the sensitive detection of DA and epinephrine [32].

In addition, CDs are often paired with another nanoparticle to form nanocomposite that modifies the electrode to increase total active surface area. For instance, cuprous oxide (Cu2O) is a p‐type semiconductor with low‐index planes of (1,1,1)‐crystal face that possess excellent electrochemical properties [233, 234]. The morphology of Cu2O affects its electrocatalytic performance, and an octahedral Cu2O has been coupled with CDs in a nonenzymatic electrochemical biosensing system for detection of glucose and H2O2. The nanocomposite was mixed with Nafion before being fabricated onto GCE. Due to the synergistic effect between CDs and Cu2O, it has further lowered the LOD for glucose sensing from the uncoupled system of 128–8.4 μM and H2O2 from 6.4 μM to 2.8 μM [233]. Huang et al. constructed a similar Cu2O/CDs nanocomposite sensor by directly adding CDs to reduce copper hydroxide to form Cu2O/CDs without using other reducers or stabilizers [234]. The sensor formed was successfully demonstrated for DA sensing with low LOD of 1.1 nM and high recovery rates in the range of 99.4% and 103.8% when applied for human serum samples.

CDs have also been coupled with other nanomaterials such as Pd─Au nanoalloys [235], Au nanocrystals [236], and reduced GO [237] for several applications. CDs were discovered to significantly increase the reduction and oxidation peak current when being coupled to Pd─Au nanoalloys, suggesting that CDs film was capable of accelerating the electron transfer kinetic and therefore improving the sensitivity of a biosensor [235].

Besides acting as stabilizing and reducing agents, CDs are commonly found to be rich with carboxyl groups that can serve as good anchor for biological moieties such as enzymes and antibodies without changing its initial biological properties. Furthermore, the addition of CDs provided desired microenvironment for the enzyme to transfer electrons back to the electrode for electrochemical signal detection [238].

9.4.2 Uses of CDs in Bioanalysis

Owing to the unique optoelectronic characteristics of CDs, there are a lot of biosensors being studied and reported in the literature for bioanalysis applications. Among those reported, the most common CDs‐based sensing systems are adopted for detection of heavy metals associated with diseases, DNA, small molecules, pharmaceutical drugs, proteins, antibodies, pH, and other biomarkers as part of disease screening. This section will explore and reveal some examples reported on the use of CDs for bioanalysis.

9.4.2.1 Heavy Metals/Elements

While some metals such as iron and zinc are essential elements for human physiological health, other heavy metals might portray a high toxicity effect to the environment and human health. Most metals when ingested in large amount result in immediate direct poisoning or damaging diseases in a longer run [239]. For instance, heavy metals such as Hg(II), cadmium (Cd(II)), lead (Pb(II)), and arsenic (Ar(II)) pose a high toxicity risk to human and other organisms. CDs can be developed to detect these heavy metals, mostly in their ionic state. In fact, some of the earliest demonstrations on the applications of CDs as sensing probes are on the detection of Hg(II) [240244]. Most of these CDs‐based sensors adopt cysteine as sensing receptor for the detection of Hg(II) since the thiol group can bind well to Hg(II) [245]. It is worth noting that most of the Hg(II) detection systems reported to date are performed in liquid phase directly using the CDs, while only Gonçalves et al. have reported the development of CDs into a solid platform for the sensing application [242, 246]. The CDs were immobilized on an optical fiber via layer‐by‐layer approach, and the resulting optical fiber sensor had achieved an LOD of 0.01 μM for Hg(II).

Since their practicality for Hg(II) sensing, CDs have also been explored for the detection of other metal ions based on similar sensing mechanism. For example, CDs have been developed for the detection of Pb(II) ions with LOD of 5.05 μM [247] and ferric ions (Fe(III)) with LOD of 1.06 μM [3]. Usually, the surface of the CDs is reported to be oxygen rich due to the presence of hydroxyl (OH ) and carboxyl (COO ) groups that are introduced during the carbonization process. This interface is of high affinity to attract the positively charged metal cations. Such interactions could result in the disturbances of the initial origin of the fluorescence, leading to quenching of the fluorescence. The degree of quenching can be correlated to the amount of metal ions present to establish a calibration curve, which later can be analyzed to predict the concentration of metal ions in unknown samples. Most of the CDs synthesized via aerobic carbonization or hydrothermal method have shown fluorescence quenching by Fe (III) ions [24, 248].

More recently, CDs sensing systems have also been developed for selective detection of other toxic metals such as beryllium ions [249], copper ions [95, 250255], zinc ions [256, 257], arsenic [258], palladium [259], and aluminum ions [260]. Interestingly, the known interaction between CDs and metal cations can also be exploited for selective detection of anions such as sulfides, as demonstrated by Barati et al. [261]. While sulfide ions themselves did not pose any changes on fluorescence of CDs, metal‐mediated sensing has been developed via two pathways. First, Hg(II) ions worked as quencher of the fluorescence of CDs and secondly, the presence of sulfide ions could restore the fluorescence. This approach has allowed the detection of sulfides in the range of 2–10 μM. On the other hand, Ag(I) ions that were found to have no effect on the fluorescence of CDs have facilitated the quenching by sulfides due to the formation of Ag2S particles, enabling detection within linear dynamic range of 1–100 μM.

9.4.2.2 Reactive Oxygen/Nitrogen Species (ROS/RNS)

ROS is an oxygen‐containing chemical species that is highly reactive, which typically includes peroxides, superoxide, hydroxyl radicals, and singlet oxygen. Similarly, reactive nitrogen species (RNS) is a nitrogen‐containing species that is chemically reactive – namely, peroxynitrite and nitric oxide (NO). ROS and RNS could jointly or individually cause damage to cells, resulting in cellular oxidative or nitrosative stress, and are often associated with blood homeostasis, inflammatory responses during cardiovascular events, or cell apoptosis related to cancer killing [262]. Due to their important roles in biological system, it is essential to monitor the amount of ROS/RNS. Under this motivation, CDs have also been explored for detection of these reactive species. For instance, S and N co‐doped CDs has been developed to detect NO, a class of radical that works as physiological messenger and effector in mammalian cells [263]. The sensing system has achieved a linear response range from 1 to 25 μM and an LOD of 0.3 μM. Another group has also reported sensors for NO using CDs, employing a bimodal sensing system based the colorimetric and fluorimetric observation as a result of reaction between NO and aminoguanidine residues in CDs [264]. The reaction between NO and CDs produced azo residues that give rise to the observable yellow‐red color changes. In the aqueous phase, absorbance increment at 490 nm and fluorescence quenching at 430 nm were observed with the increasing concentration of NO. To further explore the applications in cellular detection, NO generation was tested on treated and untreated mouse leukemic monocyte macrophage cell line RAW 264.7. NO was produced by the addition of lipopolysaccharide (LPS). The CDs have shown fluorescence quenching as an indication of NO production, while not observed by the control cells that were not treated with LPS. In another study, fluorescence detection of both peroxynitrite (ONOO ) and NO was demonstrated using ethylenediamine doped CDs [265]. The sensing of ONOO was optimized at pH 10 while the sensitivity of NO was highest under acidic condition at pH 4. The same research team has synthesized tryptophan‐doped CDs for ONOO sensing at pH 7.4 and reported linear dynamic range within 5–25 μM and LOD of 1.5 μM [266].

CDs have been reported to detect superoxide anion (O2 ˙− ) with very minimal interferences (<5%) from other coexisting ROS and small molecules such as amino acids and metal cations [267]. This system has combined CDs with hydroethidine (HE) and it was observed to give an emission at 525 nm upon excitation at 488 nm. HE has served as recognition receptor due to the high binding affinity for O2 ˙− . Upon the introduction of O2 ˙− to the system, a new fluorescent peak at 610 nm was observed while peak at 525 nm has remained constant. This has allowed a ratiometric detection to be adopted for the detection of O2 ˙− . The LOD was evaluated to be 0.1 μM. Besides O2 ˙− , CDs have also been explored for the detection of hydroxide radicals (OH˙ ) by CDs‐ascorbic acid (AA) hydrogel [268]. The hydrogel was constructed from AA derivative as the building block and then further used to encapsulate the amphiphilic CDs. When the OH˙ was added to the hydrogels, the ROS oxidized the AA constituents and disintegrated the hydrogel scaffold, which, in turn, quenched the strong fluorescence of CDs. This hydrogel sensor has successfully demonstrated for screening the apoptotic activity of 5‐fluorouracil, which is a chemotherapeutic drug compound.

9.4.2.3 Oligonucleotides

Oligonucleotides are short nucleic acid chains containing DNA or RNA molecules. DNA and RNA carry genetic information in biological systems and are often applied in genetic testing and gene‐related bioapplications. Mutagenic changes of DNA are often associated with cancers or other genetically affected diseases. For instance, long‐term exposure to UV has been known to cause DNA damage and can lead to skin cancer. In view of this, there is a need for early detection of DNA damages. There are studies reported that design CDs to monitor the DNA damage. This is by monitoring the FRET interaction between the fluorophore and ethidium bromide (EtBr) that has been preloaded to stain the electrophoretic gel [269]. Although EtBr has been in practice for a long time as a fluorescent tag for DNA detection, there have been some pressing concerns of the health risks caused by its carcinogenic nature and its reputation as mutagen that has high binding affinity toward DNA. The strong fluorescence emission by CDs has allowed the replacement for EtBr in gel electrophoresis [270].

Besides detecting DNA damages, CDs have been demonstrated for the identification of the segment of DNA or genes responsible for particular diseases. For instance, ssDNA coded for human immunodeficiency virus (HIV) has been studied for fluorescence detection based on CDs. Most of these biosensors makes use of FRET between CDs and other fluorescent nanoparticles such as GO/AuNPs [271] and Ag nanoclusters [272]. While most of the HIV genes detections are studied using ssDNA, Liang et al. reported a CDs‐based biosensor for sensing of HIV‐1 dsDNA [273]. In the presence of the dsDNA, the fluorescence of cadmium tellurium (CdTe) QDs pretagged with mitoxantrone (MTX), which could intercalate with DNA was quenched while the fluorescence of CDs remained constant. This has served as a good reference point for ratiometric analysis of fluorescence intensities between the two nanoparticles. Based on this system, the signal‐to‐noise ratio recorded was showing a correlation with the targeted dsDNA and achieved excellent recoveries ranged between 98% and 110% when tested on spiked serum samples of healthy volunteers.

miRNAs are a group of small, noncoding, endogenous, regulatory single‐stranded RNAs that are involved in gene expression regulation. As such, they are often selected as biomarkers for cardiovascular diseases, cancers, and viral infections. Detection of a miRNA, miR9‐1 associated with breast cancer has been reported by observation of fluorescence quenching of CDs [165]. The sensing was designed by adding FAM‐labeled ssDNA to CDs, which no or low fluorescence was observed due to FRET mechanism. However, the emission can be recovered when miR9‐1 is bound to the FAM‐labeled ssDNA forming a duplex helix and being released from the surface of CDs. Furthermore, a DNA‐labeled CDs sensing system based on ECL transduction has been demonstrated for miRNA‐21 with linear dynamic range between 10 aM and 104  fM [166].

9.4.2.4 Small Molecules/Pharmaceutical Drugs/Natural Compounds

Small molecules are organic compounds with typical molecular weight below 900 Da. Some examples include primary and secondary metabolites that assist in regulating the normal biological processes such as amino acids, vitamins, nucleotides, etc. Monitoring of these small molecules in the biological system could be useful in understanding more of a particular disease or their metabolic pathways in finding a cure for the disease. On the other hand, monitoring of pharmaceutical drugs is becoming more important especially in cases where the misuse or overuse of drugs. For instance, the misuse of antibiotic has led to more serious problems such as the birth of superbugs, which may threaten the survival of humankind.

Recently, there have been more reports on utilizing CDs for the detection small molecules. One of the latest is by Fong et al., which demonstrated a simple CDs “turn‐on” fluorescence system for the sensing of vitamin C, also known as AA [274]. In designing the system, the fluorescence signal of the isolated CDs was first effectively quenched by using Fe(III) ions. Once this is achieved, the addition of AA into the system will restore back the fluorescence due to the high affinity of AA toward the Fe(III) ions, where the degree of enhancement in terms of the intensity can be correlated to the amount of AA added. Based on the similar concept, Lin's group has demonstrated a “turn‐on” sensing of AA in rat brain following brain ischemia [275]. The CDs synthesized have been premodified with hexagonal cobalt oxyhydroxide (CoOOH) nanoflakes. During the FRET process, the fluorescence of CDs (λem – 450 nm) was turned off by CoOOH nanoflakes when being excited at 380 nm and eventually turned back on when the recognizing unit of CoOOH selectively binds to the enediol group of AA. Cerebral AA in brain microdialysate was determined by implanting the microdialysis probe in the brain cortex region through the guide cannula and was perfused with artificial cerebrospinal fluid. Upon equilibration via microinjection pump, every 30 μL of brain dialysate was collected for AA sensing. This method has enabled in vivo AA sensing from the cortex of rat brain under different physiological conditions including calm and ischemia. Another group has also reported AA sensing in with LOD as low as 5 nM and further quantifying AA in human serum samples and rat brain microdialysate samples [276].

Cholesterol checking is critical, as a high level could result in health complications such as cardiovascular diseases, obesity, and hypertension. Cholesterol can be detected using CDs without involving of enzyme cholesterol oxidase that is commonly adopted to identify cholesterol in existing bioassay. The mechanism is based on the functionalization of β‐cyclodextrin on CDs, in which the core of the β‐cyclodextrin can entrap various hydrophobic molecules with different affinities [37]. In this system, p‐nitrophenol was first bound to the CDs sensing probe, causing quenching of the fluorescence. The fluorescence was then recovered by cholesterol, which has higher affinity toward the β‐cyclodextrin. Another enzyme‐free biosensor has been reported on using Au/CDs nanoconjugates for a dual function (colorimetric and fluorimetric) detection scheme [277]. The Au particles in this system took part in the colorimetric responses and fluorescence quenching in the presence of cholesterol within concentration of 0.208–2.08 mM. When cholesterol is present, a visible color change can be observed together with fluorescence quenching of CDs. The system has very minimal interference from other coexisting entities and recorded a LOD of 2.5 μM. Similarly, a hemoglobin (Hb) conjugated‐CDs (CDs/Hb) complex was constructed for cholesterol sensing [278]. The detection of cholesterol was performed as the CDs break free from the CDs/Hb complex when Hb favors the hydrophobic interaction with cholesterol. A LOD of 56 μM with response time of less than five minutes in human plasma samples has been reported.

Nervous system−related diseases such as Parkinson's disease has been a major concern especially the average age of the world population has grown older. In view of this, DA serves as an important biomarker for such diseases. For the past decades, many works have been invested in search of an accurate sensing system for DA and CDs have also been studied for such purpose. Most of the reported systems for DA determination have shown good correlation between the changes of fluorescence of CDs with the concentration of DA. More recently, Fang et al. have reported a carbon‐fiber microelectrode that has been co‐deposited with CDs and GO nanosheets for selective sensing of DA [279]. In this system, CDs serve as the physical barrier to avoid the complete restacking of GO nanosheets while in the meantime increasing the surface area of accessible sheet surface for DA interaction. This approach has been reported to improve the performance of the sensor. Besides, the group claimed that CDs themselves could interact with DA molecules and increase charge storage capability of the electrodes for signal amplification. Besides using CDs as part of sensing transducer, CDs can be used as matrix for matrix‐assisted laser desorption‐ionization mass spectroscopy (MALDI MS) as demonstrated for screening of three compounds: serotonin (Sr), glutamic acid (Glu), and DA that serve as biomarkers for neurological disorder [280]. The strong absorbance in the UV range (220–350 nm) of CDs was exploited to facilitate the energy transfer from Nc laser (337 nm) of MALDI MS to the three targeted analytes. Due to the significantly reduced background noises, the LOD of Sr, Glu, and DA could be achieved at values as low as 3, 5, and 8 nM, respectively.

Most pharmaceutical drugs or natural compounds are categorized under small molecules. Quantification of such organic compounds known with medicinal benefits is essential for clinical analysis. As such, a substantial amount of effort has been invested to develop biosensors for monitoring these compounds, especially in vivo condition. CDs‐based sensing systems have been explored for this and have successfully demonstrated detection of different drugs such as plant‐derived sinapine [281], melamine [217, 282], sophoridine of traditional medicine [156], isotretinoin or vitamin A [283], allopurinol as uric acid reducing drug [284], patulin as an antibiotic (which was then found to be toxic [285]), amoxicillin as a common antibiotic [286], cisplatin as the anticancer drug [287], and neurotoxins such as tetrodotoxin [288] and hydrazine [289].

9.4.2.5 Proteins

Immunosorbent assay is a widely adopted technique in biochemistry to detect biomolecules based on antibody–antigen interactions. Some efforts have been taken to incorporate CDs into the working principle of this assay to obtain a biosensor that can give direct quantification signal. This is such as the development of a CDs‐linked immunosorbent assay for the detection of human AFP [290]. In this sensing probe, the captured anti‐AFP (Ab1) was first immobilized onto polystyrene well plates in the presence of BSA as the inhibitor for the unsaturated binding sites. Following that, a CDs‐labeled anti‐AFP pair (Ab2) was added to the well to form sandwich immunocomplexes with the bound AFP. The concentration of AFP antigen could then be determined based on the direct correlation with the fluorescence intensity measured. Other examples include CD‐labeled antibodies (IgG‐CDs) for in situ visualization of glyphosate distribution in plant tissues as well as detection of glyphosate in river water, tea, and soil samples [291].

There have been innovative explorations with CDs as biocatalysts in immunosorbent assay for the replacement of enzymes in the existing approach. Studies have shown that CDs possess superior peroxidase activity as a result of Fe doping, which mimics the nature of ferriporphyrin [292]. As such, the Fe‐doped CDs (above 70%) were further proven to exhibit excellent catalytic activity of above 70% when bound to antibody, in comparison with the natural horseradish peroxidase (HRP) at below 10%. The catalytic activity of these modified CDs was observed by the color change upon oxidation of 3,3,5,5‐tetramethylbenzidine (TMB) in the presence of H2O2. Besides the high catalytic performance, these CDs displayed superiority against the HRP as they are able to catalyze under wider range of pH and temperature conditions. In this classical enzyme‐linked immunosorbent assay (ELISA) kit, CDs labeled with antibody have been employed as biocatalysts to replace HRP for efficient and specific quantification of CEA. In another report, a sandwiched ECL immunosensor for CEA has also been reported with LOD of 1.67 pg ml−1 with dynamic linear range of 5– pgml−1 to 500 ng ml−1 [293]. For this biosensor, primary antibody, Ab1 was conjugated to the polydopamine and AuNPs nanocomposites while the secondary antibody, Ab2 was linked to CDs that has been premodified with poly(ethylenimine) functionalized GO. This immunosensor has been applied for real samples analysis in human serum with relative errors below 10% when validated with ELISA kit.

Thrombin is a trypsin‐like serine protease that is involved in converting soluble fibrinogen into insoluble strands of fibrin and catalyzing a variety of other coagulation‐related activities [294]. Due to their pro‐inflammatory character, they are believed to be involved in the onset and progression of atherosclerosis and therefore always being chosen as a diagnosis tool for cardiovascular diseases. Most detections make use of thrombin aptamer to improve the targeting efficiency of the biosensor. A number of thrombin sensors have been demonstrated based on PL phenomena such as color change [295], fluorescence [170], phosphorescence energy transfer [296], and FRET [168]. During a coagulation reaction, the concentration of thrombin in the biological system can range from below 1 to 500 nM [297]. These CDs‐based biosensors for thrombin have mostly reported LOD in the nanomolar concentrations and matched with the practical range for the coagulation process in blood. Hb in red blood cells is involved in transporting oxygen to maintain aerobic physiological activities in human body. Low concentration of Hb will lead to anemia and can be fatal. Therefore, it is important to monitor the concentrations of Hb for clinical diagnosis, especially detecting hemolysis. Huang et al. developed a simple mix and detect fluorescence sensing of Hb by observing the fluorescence changes in the blue region [298]. When added with Hb sample, the initial CDs emission at 430 nm was quenched upon excitation at 340 nm. The quenching was based on static quenching mechanism where ground‐state complex between CDs and Hb was formed. This was further proven by the structural changes in Hb upon interaction with CDs as the α‐helix content of Hb decreased and the secondary structure of Hb rearranged. With low LOD of 0.12 nM, this sensing probe was tested for human urine and blood samples and showed high recovery rates of 95–100%. Besides the direct interaction between Hb and CDs, the presence of a strong oxidizing agent such as H2O2 can facilitate the detection of Hb [299]. ROS such as OH˙ is generated during the reaction between Hb and H2O2, which then will quench the fluorescence of CDs. The quenching by OH˙ has a direct correlation with the quantity of Hb in the linear range of 1–100 nM and LOD of 0.4 nM.

9.4.2.6 Enzyme Activities and Inhibitor Screening

Enzyme activities or kinetics and enzyme inhibition have been important processes employed in clinical diagnosis to derive the cause of a disease. The information harvested will provide insight on the role of the enzyme in metabolism. A drug can later be designed to control or inhibit this enzyme activity to drive to whole metabolism process toward an intended outcome. Many enzymes have a close relationship with certain diseases, and eventually most of them can serve as biomarkers for diseases such as breast and prostate cancer, diabetes, and bone diseases [300]. CDs have been exploited for their applications in enzyme assays. One of the most commonly tested enzymes in the clinical practices is the alkaline phosphatase (ALP), which has also been detected using CDs as the bioanalysis probe. The main aim is to be able to detect ALP in the concentration range of 40–190 U l−1 in human serum [301]. Feng et al. have reported a recyclable biosensor for ALP based on aggregation and disaggregation of CDs via competitive approach [302]. The sensing mechanism follows two steps outlined as followed: (i) CDs tend to aggregate as triggered by cerium ions that can strongly coordinate with the carboxyl groups on the surface of CD and lead to fluorescence quenching; (ii) ALP catalyzes the hydrolysis of ATP to phosphate ions, which has higher affinity to cerium ions than the carboxyl groups on CDs. This leads to the disaggregation of the initial complex and thus will recover the fluorescence of the CDs. The study showed that there is a linear correlation between the degrees of recovery of fluorescence with the ALP concentrations within the range of 4.6–383 U l−1 . Similarly, Cu(II) ions could turn off the fluorescence of CDs, which was then restored by phosphate ions when ALP catalyzed the transformation reaction from pyrophosphate ions to phosphate ions [303].

The fluorescence “off−on” sensing mechanism making use of the catalytic properties of the enzyme has also been demonstrated for detection of other enzymatic activities such as ALP [304306], acid phosphatase [307], and α‐glucosidase [308]. The enzyme inhibitor screening can be carried as a continuation of the aforementioned two‐step system. This has been applied for acetylcholinesterase (AChE) and its inhibitor screening [309]. Cu(II) ions were first added to aggregate and quench fluorescence of CDs, followed by AChE catalyzing the hydrolysis of acetylthiocholine into thiocholine, inducing the fluorescence recovery. When the inhibitor tacrine is present, AChE loses its catalytic ability to hydrolyze acetylthiocholine and therefore fluorescence of CDs remained quenched instead of being restored due to the absence of thiocholine. This system has demonstrated detection of AChE activity as low as 4.25 U l−1 within linear range of 14.2–121.8 U l−1 . The FRET between CDs and hexagonal cobalt oxyhydroxide nanoflakes has successfully established ratiometric fluorescence analysis for α‐glucosidase inhibitor screening [310]. Within close proximity, the CDs transfer energy to the nanoflakes, which results in fluorescence turn “off.” Subsequently, L‐ascorbic acid‐2‐O‐α‐D‐glucopyranosyl (AAG), which has no reducing power, would be hydrolyzed by the enzyme to release AA that could rapidly reduce the nanoflakes to Co(II) ions. As a result, the FRET was halted and fluorescence of CDs was restored. Acarbose, which was one of the inhibitors for the enzyme, was studied as model and a series of natural compounds have been screened for their inhibitory effects on the enzyme activity as part of anti‐diabetic drug discovery. The tumor‐invasive biomarker, β‐glucuronidase (GLU), was also detected in a similar manner but via the inner‐filter effect of GLU's catalytic product, p‐nitrophenol, which absorbed the fluorescence of CDs due to the complementary spectral overlap between p‐nitrophenol and CDs [311]. This CDs‐based enzyme assay has been applied for tumor screening and the inhibitors screening from natural products that could be potential anti‐tumor drug. Tyrosinase (TYR) activity was monitored using DA functionalized CDs (DA‐CDs), and arbutin was selected as an inhibitor of TYR for screening [312]. It can sensitively monitor the intracellular TYR level in melanoma cells and intracellular pH changes in living cells. Due to the spectral overlap between CDs and neutral red (NR), these two fluorophores were incorporated in a FRET‐sensing system for hyaluronidase (HAase) activity [313]. By a ratiometric analysis of fluorescence emissions of CDs at 525 nm and NR at 630 nm, this system has exhibited LOD as low as 0.05 U ml−1 in aqueous form. Another FRET system of CDs and naphthalimide has been developed for detection of thioredoxin reductase (TrxR), which is commonly observed in most cancers [314]. From the linear correlation between the ratiometric intensities (I450/I525) and TrxR concentrations, LOD of 0.72 nM was obtained.

9.4.2.7 pH

pH plays a huge role in biological systems especially when the protonation or deprotonation of a biomolecule would lead to a trigger or halt a biological process. pH changes can be due to the presence of foreign species in the body such as drugs during the course of treatment or infection of diseases. Therefore, monitoring of intracellular pH plays an important role besides detecting biomarkers as part of disease screening. This fact holds true especially for the cases of cancerous diseases where a change in the pH is an indication of the presence of tumors [315]. Intracellular pH can be monitored using CDs as a strong emitting fluorophore. The change in pH can lead to variations in the fluorescence emission of CDs, whether by direct monitoring of fluorescence intensity changes or by performing ratiometric measurement of the fluorescence emissions. The surface chemistry of CDs has made them useful for pH sensing due to the presence of reactive oxygenated species such as hydroxyl and carboxyl groups that are sensitive to the uptake and release of hydrogen ions. For instance, the reversible fluorescence changes in response to pH changes exhibited by CDs has been demonstrated and suggested to be due to the presence of acidic/basic sites on the surface [316, 317]. Ratiometric measurement of fluorescence for pH sensing has been a popular practice since the fluorescence changes can be maximally correlated to the pH variations. This sensing technique is applicable to the fluorophores displaying dual fluorescence emissions peaks, where the emissions can each respond specifically to the change of pH. For instance, a two‐photon fluorescent nanoprobe based on CDs has been reported for physiological pH sensing in the range of 6.0–8.5 in living cells and tissues [318]. As shown in Figure 9.5, the nanoprobe consisted of CDs that have been surface bound by a 4ʹ‐(aminomethylphenyl)‐2,2ʹ:6ʹ,2ʺ‐terpyridine (AE‐TPY) molecule, which acted as a selective receptor to detect the change in the concentration of H+ . In a surge of H+ , a strong luminescence was observed upon two‐photon NIR excitation at wavelength of 800 nm. This observation could be reversed when there was a rise of pH, resulting in weak emission at 498 nm. A similar ratiometric sensor for pH has also been reported based on dual fluorescence emissions between CDs and other fluorophores such as fluorescein isothiocyanate (FITC) [319, 320].

Image described by caption and surrounding text.

Figure 9.5 Schematic illustration of pH sensing and bio‐imaging based on two‐photon fluorescence interaction of AE‐TPY bound CDs.

Source: Reprinted with permission from Ref. [318].

9.4.2.8 Temperature

Temperature is one of the most important parameters for a living organism, as it governs a wide variety of biological reactions from gene expression to cellular metabolism. From a clinical viewpoint, pathological cells are present at slightly higher temperatures than that of normal healthy cells due to their enhanced metabolic activity [321]. Therefore, monitoring cellular temperatures could be useful for better understanding diseases and for developing novel diagnosis and treatment options. In recent years, the concept of a nano‐thermometer has been introduced to measure cellular temperature. Fluorescent nanoparticles can be adopted to map the temperature in living cells since their optical properties such as emission band shape, peak position, intensity, or lifetimes are strongly affected by temperature [322, 323]. The fluorescence of CDs is responsive to the changes of temperature and can be employed to sense change of temperature in liquid from 15 to 60 °C [67]. More recently, CDs have been studied for application as nanoscale thermometry in living cells. Zhang's group has looked into dual‐emission fluorescent responses of a nanohybrid consisting of CDs and gold nanoclusters with single excitation, which was well facilitated within the physiological temperatures of 25–45 °C [324]. The gold nanoclusters were first functionalized with a targeting moiety, GSH to improve the targeting efficiency. The system was then conjugated with amino‐functionalized CDs to form the nanohybrids. The ratiometric analysis of fluorescence dual‐emission at 430 and 605 nm upon single excitation at 367 nm was shown to be more efficient and accurate compared to that of single fluorescence emission. When tested in human embryonic kidney 293T cells, the blue emission from CDs was observed to remain unchanged and served as a reference point, while the red emission corresponding with the gold nanoclusters decreased when the incubating temperature was elevated to 45 °C. Similarly, ratiometric temperature sensing has been demonstrated in vitro when CDs have been coupled with a temperature‐sensitive dye, rhodamine B [325]. On the other hand, Zhang's team has also developed a single‐emission CDs‐based thermometer by observation of the fluorescence “off‐on” using microscopy imaging [326]. When the MC3T3‐E1 cells were incubated with GSH at room temperature, the red fluorescence (λem – 615 nm) of CDs was found to decrease but was slowly restored when the temperature was increased to 30 °C and further intensified when the temperature was raised to 40 °C as a result of dissociation of GSH from CDs. This observation has proven that the CDs‐based sensing system could readily enter the cytoplasm of the cell and detect temperature changes in the subcellular level. Besides, since PL lifetimes of fluorophores are dependent on temperature, it could also be applied as nanothermometers in cells, which was demonstrated by Kalytchuk et al. between 15 and 45 °C in human cervical cancer HeLa cells [327]. The PL lifetimes of the synthesized CDs decreased monotonically with the increasing temperatures, without being affected by the pH, concentration of CDs, and the environmental ionic strengths.

9.4.3 Solid‐State Sensing for Point‐of‐Care Diagnostic Kits

Most medical diagnoses are carried out in pathology lab setting, which means that special laboratory knowledge, facilities, and skills are required for the screening process. Besides, longer processing and waiting times are needed before the medical team and patients can find out the cause of a sickness. Point‐of‐care diagnosis has been proposed as a feasible solution to improve the healthcare system. In addition, microfluidics have been developed as miniaturized portable analytical devices for instant test outcome with high accuracy. Biosensors are being studied and improved as part of the process to translate sensory systems into solid platforms for the miniature devices.

There are reported efforts on improving different problems faced while developing solid‐state sensors. The main issue in transferring the sensing probe from liquid phase onto solid platform is the leaching of the fluorophore that can cause tremendous loss in sensing signals. Fabrication techniques become an important factor with the intention to introduce more fluorophores onto the sensor matrix without hindering the interaction with the sensing analytes. Besides layer‐by‐layer immobilization methods as demonstrated by Gonçalves et al. for Hg(II) optical fiber sensor [246], another group has established a fabrication method using colloidal photonic crystals (CPhCs) as a solid matrix without compromising the sensitivity toward the targeted analytes [328]. CPhCs films can act as a Bragg mirror and enhance the fluorescence or incident light with the wavelength located into their photonic band gap, thus improving the sensitivity of the sensor [329]. They can also concurrently modulate electromagnetic waves and exhibit a photonic band gap that is similar to the effect of a semiconductor band gap. Li et al. adopted CPhC comprising three layers of monolithic CPhCs for their solid‐state sensor for Hg(II) [328]. The double hetero‐structure of CPhC was formed by successive deposition of CPhCs in a specific sequence from bottom to top. The Hg(II) targeting CDs were immobilized on the surface of polystyrene spheres as the middle layer in between two monolithic CPhC layers with a periodicity overlapping the excitation wavelength. The excitation efficiency was enhanced as the layers were able to reflect any excitation light that propagate in the CDs layer. Due to the fluorescence enhancement effect from the CPhC structure, the LOD of the solid‐state sensor was obtained as low as 91 pM for Hg(II).

Paper‐based microdevices have attracted a lot of attention as low‐cost sensing platforms since the discovery of microfluidic paper‐based analytical devices (μPADs) [330, 331]. The advancement in the screen‐printing technology has aided the research focus on μPADs in the area of fabrication of designed sensing systems. Wax screen printing is an inexpensive method that has been adopted for fabrication of μPADs, and it involves two fundamental steps: (i) printing patterns of wax on the surface of the paper and (ii) melting the wax onto the paper with heat to form the hydrophobic barrier [332]. A few studies have reported on screen‐printed recyclable biosensors based on ECL transduction for in situ screening for IgG antigen [162], MCF‐7 cancer cells [177], and K562 leukemic cells [333]. Wu et al. coupled RCA method with oligonucleotide functionalized CDs with sensitive ECl signal, within a μPAD for detection of larger target analytes such as proteins [162]. In this system, a three‐dimensional (3D) macroporous Au paper electrode has been employed as the working electrode to capture antibodies. Meanwhile, the RCA is an isothermal DNA amplification strategy that produces tandem‐repeat sequences to serve as template for periodic assembly of CDs, in which each protein recognition event is presented to a few CDs tags for ECL signal readout. The sensing system incorporating the CDs and antibody conjugated AuNPs electrodes were printed onto the paper origami device as depicted in Figure 9.6. The incorporation of RCA has significantly enhanced the sensitivity of the device for monitoring the concentration of IgG as the ECL intensity of CDs increases alongside the cascade DNA nanotags. This device has demonstrated linear dynamic response to IgG in the range of 1.0 fM to 25 pM with LOD of 0.15 fM. Another paper device based on ECL responses of CDs has been reported for observation of MCF‐7 cancer cells and the electrode was then further applied as in situ screening for anti‐cancer drugs [177]. This extension of application was achieved by monitoring the number of apoptotic cancer cells after in situ 3D culturing of the captured cancer cells on the electrode to a drug containing medium. Anticancer drugs such as cisplatin, paclitaxel, fluorouracil have been selected as study model. The relative respiratory activity of captured MCF‐7 cells was investigated to test the chemosensitivity of the μPAD. The result obtained was found in agreement with the validation of drugs induced apoptosis via fluorescence imaging technique. Besides ECL being the commonly applied transduction for paper‐based biosensors, CL‐based paper devices have also been reported for DNA sensing [334]. Similar to other CDs‐based CL sensing mechanisms applied in solution form, this μPAD employed CDs‐dotted nanoporous Au as a signal‐amplification label. N,N‐disuccinimidyl carbonate (DSC) was in charge of capturing DNA by covalently immobilizing on the paper device. In the presence of an oxidizing agent, the CL reaction was induced and produced CL signals corresponding to the targeted DNA and successfully performed at an LOD as low as 8.56 × 10−19  M.

Image described by caption.

Figure 9.6 A schematic illustration of the fabrication and assay procedure that are based on paper origami device. The first step involved patterning paper sheets in bulk using a wax printer as shown in Sheet A, (a) showing 3D origami device without the screen‐printed electrodes. The electrodes on sheet A are screen‐printed in bulk (b, c). The last step involves cutting and folding sheet A into individual 3D origami device (d) and being modified for IgG detection.

Source: Reprinted with permission from Ref. [162].

9.4.4 Bioimaging/Real‐Time Monitoring

Bioimaging is referred to as noninvasive method to visualize biological processes in real time. It serves as an important part of clinical diagnosis. Most CDs‐based sensing systems described earlier are based on ex vivo testing, meaning the detection is applied outside of a body using biological samples such as urine and blood serum. However, very few studies have applied in vivo sensing or tracking of biomolecules or other potential analytes directly in body. Therefore, other techniques such as bioimaging are required to work alongside with bioanalysis to enable real‐time monitoring of intracellular processes. Many studies have reported the use of CDs as fluorescent labels in bioimaging due to their excellent optical stability and biocompatibility. In recent years, a few groups have studied CDs sensing in living cells via cellular imaging methods. Most of these studies focus on qualitative detection of heavy metals in living cells, observing the quenching of CDs in the presence of metal ions such as Hg(II) [335] and Fe(III) [336]. A fluorescence turn “off−on” intracellular sensing of Hg(II) and GSH has been demonstrated in HeLa cells based on observation of CDs fluorescence [337]. Most CDs are blue emitting, which could limit their practical in vivo applications due to the universal blue auto‐fluorescence of biological matrices. In addition, UV utilized for excitation can cause photodamage to the biological tissues [338]. Alternatively, CDs emitting at longer wavelength would be preferable, as they can penetrate into deeper tissues. Motivated by these concerns, there has been research into the viability of red‐emitting CDs (λem – 600 nm) synthesized via hydrothermal using a mixture of two chemical precursors: 2,5‐diaminobenzenesulfonic acid and 4‐aminophenylboronic acid hydrochloride [143]. They have performed cellular sensing of Fe(III) ions by incubating cervical cancer HeLa cells with CDs and Fe(III) ions. As shown in Figure 9.7, the red fluorescence of CDs was quenched by Fe(III) ions, which served as a qualitative indication of the presence of Fe(III) ions. Fluorescence ratio imaging is an effective approach to quantitative detection of biological molecules in living cells. This is made possible for fluorophores that emit at two different wavelengths upon single excitation. Although excitation‐dependent CDs can provide a series of different fluorescence emissions upon varied excitation wavelengths, it is still a challenge to harness single excitation/dual peak emissions from CDs itself. One possible way is via nanocomposite formation where CDs are coupled with other fluorescing nanoparticles. Zhang et al. grew gold nanoclusters on CDs that served as the skeleton by mixing molecular precursors for both individual nanoparticles in one synthesis [339]. With single excitation at 390 nm, two emissions at 470 (blue) and 565 nm (red) were observed. The fluorescence ratiometric imaging was performed by adding 30 μM of Fe(III) to the normal rat osteoblast MC3T3‐E1 cells. The results showed that no detectable change of blue fluorescence was observed but there was a significant decrease in the red emission. The fluorescence ratio at the two wavelengths has depicted a significant increment after addition of Fe(III).

Image described by caption.

Figure 9.7 Fluorescence microscopy images (a, b) and their corresponding bright‐field transmission images (c, d) of HeLa cells: (a, c) were incubated with CDs for 24 hours at 37 °C and (b, d) were first incubated with CDs followed by 100 μM of Fe(III). solutions. (See color plate section for the color representation of this figure.)

Source: Reprinted with permission from Ref. [143].

Other than metal ions in cells, the monitoring on the activation of anticancer prodrug such as cisplatin in real time for living cells using CDs is also possible using ratiometric method. It is based on observing the changes in the multi‐fluorescence peaks of CDs during the activation. Zhao et al. established such a sensing system based on the FRET pairing between CDs and Pt(IV) [287]. In this system, CDs also acted as nanocarriers for the drug and FRET donor while for cisplatin (IV), Pt(IV) was selected as the model drug and the linker to load Dabsyl quencher unit on the surface of CDs. The CDs were capable of emitting at blue (460 nm), green (535 nm), and red (610 nm) fluorescence upon excitation at 360, 480, and 545 nm, respectively. The CDs‐based fluorescent biosensor was tested on the human ovarian carcinoma cell line, A2780 cells under reductive conditions. When the sensor was constructed by the sequence of CDs‐Pt(IV)‐Dabsyl, the blue fluorescence of CDs was effectively quenched by the Dabsyl unit as the FRET acceptor while red and green fluorescence remained unaffected. In the presence of the reducing agent AA, the prodrug Pt(IV) was successfully reduced to Pt(II) species, which resulted in the prodrug activation and at the same time breaking of the bond from the sensor. Consequently, the FRET process became ineffective as the Dabsyl unit was no longer in close proximity with CDs. Therefore, the blue fluorescence of CDs was restored over time, but the red and green fluorescence remained constant throughout and can be used as effective internal reference. The fluorescence ratios, I460/I610 and I460/I535 increased gradually and was indicative of the real‐time reduction of Pt(IV) to cytotoxic Pt(II) species due to the activation of the prodrug.

9.4.5 Theranostics

Theranostics is a branch of biomedical applications where both therapy and diagnosis are achieved simultaneously within the same system. This biotechnological advancement has a lot of potential benefits as it combines both important components of detecting and curing at the same time. The bioanalytical possibilities of CDs regarding theranostics have been more broadly explored. As described earlier, CDs can play the role as fluorescence labels for bioimaging given the unique fluorescence and biocompatibility properties. This could help to achieve the diagnostics part. In addition, CDs can also serve as nanocarriers due to their small sizes in the nanometer range and ease of functionalization. A number of studies have explored on cancer theranostics using CDs as the fluorescent imaging agent. The real‐time imaging of targeted drug delivery using CDs as fluorescent label has been demonstrated by applying CDs labeled hollow mesoporous silica nanoparticles (HMSNs) as the drug carrier. Kang et al. successfully loaded DOX, which was selected as the model anticancer drug within the hollow structure and demonstrated the pH‐controlled drug release in the tumor bearing mice [340]. The pH‐controlled drug release was optimized at slightly acidic condition at pH 5.0, which was suited for the cancer cells because tumor sites are generally known to be slightly acidic. For real‐time imaging of the CDs‐based drug delivery system (DDS) in the tumor model, fluorescence scan of the mice was performed at two wavelengths, 455 and 523 nm, which was shown to be localized at the tumor sites within the mice. Similar systems have been constructed by loading the drugs onto CDs‐tagged HMSN. Besides targeting drug release at tumor sites using pH control, redox and enzyme triggering was adopted by labeling the DDS with hyaluronic acid (HA) responsive to HAase that are abundant in cancer cells [341, 342]. Cell‐penetrating peptides such as TAT peptides are another excellent option as cancer‐targeting moieties because they can efficiently translocate nanoparticles into cell nucleus by binding import receptors importin α and β (karyopherin) and then target the nuclear pore complexes of cancer cells to enter their nuclei [343]. In a nucleus targeting DDS, TAT peptides have been conjugated to CDs in a thermosensitive liposome as nanocarrier for the improvement on cell permeating capability and targeting specificity [186]. When CDs were excited with NIR in the range of 980–660 nm, the emission was obtained in the range of 440–450 nm with excellent PL intensity, supporting UCPL properties of CDs. This characteristic of CDs enabled it to act as the NIR absorber as part of heat stimulus that triggered the drug release via bursting of bubble inside the liposomes, leading to apoptosis of cancer cells. Furthermore, inclusion of CDs within the DDS allowed multi‐colored cell imaging of the liposomes as part of optical monitoring of the DDS in vivo. Apart from the chemotherapy using anticancer drug, gene‐silencing therapy can be another ideal cancer treatment option. This is due to the adjustable sequences of small interference RNA (siRNA) that can specifically reduce different oncogene expression and lower the growth of heterogeneous tumor cells accordingly [344]. A theranostics nanoagent based on CDs being integrated with multiple siRNAs such as epidermal growth factor receptors (EGFRs) and cyclin B1 has been constructed for gene delivery and intracellular tracking in lung cancer cells [194]. The surface of CDs were prefunctionalized with PEI, which could absorb protons under acidic condition typically of advantage in cancer environment would lead to electrical repulsion of the CDs. This will, in turn, minimize the light‐shielding effect by particle aggregation and increase the fluorescence of CDs, enabling more efficient imaging of the CDs in acidic endosome after receptor mediated endocytosis by the lung cancer cells.

PDT is a noninvasive treatment that involves a photosensitizer that will generate cytotoxic singlet oxygen when being exposed to a specific wavelength of light [345]. Serving as an extension from PDT, PTT adopts electromagnetic radiation typically NIR region for treatment, whereby a specific NIR wavelength excites a photosensitizer and releases vibrational energy in the form of heat and eventually kill the targeted cells [346]. These treatments have been joined with bioimaging to concurrently achieve therapy and diagnosis. CDs are often adopted as both the nanocarrier and fluorescent imaging agent. A unique structural assembly of CDs with clathrates (denoted as CDsCL) has been demonstrated for a targeted delivery of methotrexate (MTX), an anti‐cancer drug under physiological conditions followed by photothermic and photodynamic treatment on cancer cells [347]. This assembled complex is found to be dependent on pH and thus poses an advantage to precisely control the release of an anticancer drug in vitro. Due to the high absorbance of NIR wavelengths, a linear correlation between the temperature and the concentration of CDs was observed upon laser irradiation. Breast cancer stem cells (BCSCs) were treated with CDs complex and irradiated with NIR laser. The fluorescence observed in the cells was indicative of the cytoplasmic intake of the CDs complex. Upon NIR laser irradiation, the cells were detected as slightly deformed due to the local thermal stress created by CDs in the cytoplasm, changing the internal pressure within the cells. Furthermore, the change in the cell morphology to round‐like cells was due to the anti‐proliferative effect of MTX drug and the high temperature rise due to laser irradiation to CDs complex. Besides, the fluorescence of CDs remained constant throughout the study, proving its photostability for such applications. A similar working concept has been adopted in construction of a magnetofluorescent CDs complex for the controlled drug release of DOX for chemotherapy combined with PTT induced by the NIR laser irradiation at cancer cells [348]. An integrated system involving CDs and transition‐metal dichalcogenides (TMDs) such as WS2 nanorods has exhibited great potential for PTT [349]. The pronounced cell death observed in the treated cells has proven that this system was capable of inducing heat that caused the cell death upon NIR laser irradiation. In the meantime, Beack et al. constructed CDs'chlorine e6‐hyaluronate conjugates as part of PDT for melanoma skin cancer treatment [350]. The transdermal delivery of the conjugate has been validated with fluorescence imaging, confocal microscopy, and two‐photon microscopy. Upon the laser irradiation, the singlet oxygen generation was enhanced in the presence of the CDs‐based conjugates and therefore resulted in significant suppression of tumor growth, indicating the effectiveness of the PDT system.

9.5 Future Perspectives

9.5.1 Better Understanding of PL Mechanisms

Despite the bright and promising future of CDs for bioanalysis, there are still limited understanding concerning the fundamental PL mechanism of CDs. The origin of fluorescence of CDs is still vague and questionable, unlike the well understood mechanism for QDs. This to some extent will limit the progress to further develop CDs for bioanalysis purpose. Some studies have been performed to reveal the mechanism and this has ended up with various suggestions without any clear consistency. Perhaps, these studies were using different conditions and starting materials, which make it difficult to make direct comparison or coming out with a general conclusion.

Recently, a study has compared the fluorescence properties of CDs with organic fluorophores synthesized from the same molecular precursors under similar conditions [78]. The findings showed that the CDs synthesized by the usual bottom‐up hydrothermal method exhibited similar optical properties as the organic fluorophores with known chemical structures and characterizations. It has raised the question of the fluorescence nature of other CDs produced from the same manner (i.e. via bottom‐up approaches). There is hypothesis that organic fluorophores may be the actual fluorescence origins of CDs obtained from similar bottom‐up routes. Therefore, a concrete understanding of the origin of CDs PL is essential in order to be able to efficiently exploit the PL properties of CDs in various applications.

9.5.2 Establishment of Systematic Synthesis Protocol

The literature contains many reports of synthesis methods, each using a wide variety of starting materials. Typically, it has been proven that CDs can be obtained from basically any organic matters containing carbon, whether it is food, drinks, or organic wastes such as agriculture biomass. Although they exhibit excellent PL properties, this group of fluorophore is often reported with broad spectrum, unlike the narrow ones typically reported for QDs. The nonstandard synthetic pathways and the application of noncontrolled starting precursors may be the underlying causes. Although the facile synthesis procedures to obtain fluorescent CDs may be simple and convenient, the inconsistency in the starting precursors and the carbonization methods has yielded CDs of various similar but not exactly the same physical, chemical, and optical properties. For example, although most studies have reported the excitation‐dependent fluorescence emissions of CDs some still accounted for excitation‐independent fluorescence. In another instance, the physical morphology findings of CDs have varied from study to study. CDs are generally defined as zero‐dimensional quasi‐spherical carbon materials having the appearance of dots with sizes below 10 nm. However, some studies have reported CDs with sizes range above 10 nm exhibiting similar fluorescence properties. Therefore, a systematic and comprehensive characterization of CDs is required. Meanwhile, majority of the synthesis procedures studied for CDs are only suitable for small scale production to cater for fundamental proof‐of‐concept study purposes. In view of the great potential of CDs in bioanalysis applications, large‐scale synthesis of CDs in pure and high‐quality form is required. Upon establishing how the different synthesis factors affect the outcome on the properties of CDs, a standard synthesis and purification protocol for CDs should be established. Ultimately, there must be a standard synthesis method to produce CDs with intended properties, rather than just following the trial‐and‐error approach that is currently used.

9.5.3 QY Improvement and Spectral Expansion to Longer Wavelength

Moving on from the fundamental PL origins and synthesis studies, it has been well established that CDs can be an excellent alternative to QDs in many areas of applications, owing to their interesting PL properties and biocompatibility issues. With little concern for the health risks or cytotoxicity problems, CDs have been explored for biomedical field typically in sensing and imaging modals. However, a majority of the CDs have shown relatively low QY as compared to QDs. Particularly in optical‐sensing applications, high QY is essential for improving the baseline signal for detection that can, in turn, affect the sensitivity performance of the sensor. With that in mind, studies should be carried out to investigate the factors affecting the QY of CDs and subsequently improving the QY values.

In the meantime, research should also focus on expanding the spectral coverage of CDs to longer wavelengths. It will be useful in providing strategic advantages to the usage of CDs in bioimaging or PDT applications. CDs are presented with excitation‐dependent fluorescence emissions when excited with a range of wavelengths. To date, most of the CDs exhibit blue fluorescence in the visible range of 450–480 nm upon excitation by UV light. In vivo bioimaging and PDT are still limited to superficial structures due to light‐penetration limitations up to 10 mm deep and absorbance or reflectance issues by tissues and other biomolecules. For instance, absorbance by water molecules, proteins, Hb, and melanin are relatively high, between 200 and 650 nm, covering the whole visible range. In order to solve this issue, nanocarrier or imaging agents must both be excited and emit in the longer wavelength window above 650 nm, which has the lowest absorption in tissue and thus enables the deepest tissue penetration. Up to now, most of the CDs have been emitting blue fluorescence with excitation in the UV range. More recently, some red‐emitting CDs have been adopted for in vivo cellular studies. Therefore, to cater to such biological applications, the spectral coverage of CDs should be further expanded to the longer wavelengths.

9.5.4 Sensitivity Improvement for Solid‐State Sensing

In spite of the many research reports on the CDs for bioanalytical sensing, most of them are performed in solution phase. There are still limited studies on fabrication of CDs sensing systems onto a solid platform with high sensitivity. In general, fluorophores such as CDs are able to achieve high sensitivity and high fluorescence due to solvation effect. In dry solid conditions, CDs often lose the optical properties and therefore, a significant loss in their sensing performance due to the drop‐in sensing signal. However, CDs can be entrapped in some polymeric matrix, which can retain the optical and sensing characteristics. Therefore, more investigations in this area should be performed to fabricate the sensing systems on a solid platform and improve sensitivity toward analyte in dry form for safe development using high‐throughput biosensors.

9.6 Conclusions

In summary, CDs have achieved a lot in terms of bioanalytical applications over the past decade since its serendipitous discovery in 2004. As outlined in this chapter, the contributions of knowledge on CDs have been tremendous and vast; however, there are still various areas to be improved and explored. The unique fluorescence properties, excellent biocompatibility, and stability and ease of production from a variety of carbon‐rich sources have been the main driving force for the advancement of CDs. This new type of carbon nanomaterial has a lot of potential as a great alternative to QDs in many areas, especially the biological analysis field. In view of that, more in‐depth research on development of large‐scale synthesis of good‐quality CDs with high QY suitable for bioanalytical applications should be explored for commercialization purposes. On top of that, research and development of CDs‐based biosensors and diagnostics should be further explored and implemented to benefit the medical field.

References

  1. 1 Chang, M.M.F., Ginjom, I.R., and Ng, S.M. (2016). Single‐shot ‘turn‐off’ optical probe for rapid detection of paraoxon‐ethyl pesticide on vegetable utilising fluorescence carbon dots. Sensors and Actuators B: Chemical 242: 1050–1056.
  2. 2 Zou, S., Hou, C., Fa, H. et al. (2017). An efficient fluorescent probe for fluazinam using N, S co‐doped carbon dots from l‐cysteine. Sensors and Actuators B: Chemical 239: 1033–1041.
  3. 3 Fong, J.F.Y., Chin, S.F., and Ng, S.M. (2015). Facile synthesis of carbon nanoparticles from sodium alginate via ultrasonic‐assisted nano‐precipitation and thermal acid dehydration for ferric ion sensing. Sensors and Actuators B: Chemical 209: 997–1004.
  4. 4 Tan, X.W., Romainor, A.N.B., Chin, S.F. et al. (2014). Carbon dots production via pyrolysis of sago waste as potential probe for metal ions sensing. Journal of Analytical and Applied Pyrolysis 105: 157–165.
  5. 5 Tiong, A.N.L., Wong, N.K.H., Fong, J.F.Y. et al. (2015). A sustainable alternative to synthesis optical sensing receptor for the detection of metal ions. Optical Materials 40: 132–138.
  6. 6 Vaz, R., Bettini, J., Júnior, J.G.F. et al. (2017). High luminescent carbon dots as an eco‐friendly fluorescence sensor for Cr(VI). determination in water and soil samples. Journal of Photochemistry and Photobiology A: Chemistry 346 (Supplement C): 502–511.
  7. 7 Li, L., Wang, X., Fu, Z. et al. (2017). One‐step hydrothermal synthesis of nitrogen‐ and sulfur‐co‐doped carbon dots from ginkgo leaves and application in biology. Materials Letters 196 (Supplement C): 300–303.
  8. 8 Pandey, S., Pandey, P., Tiwari, G. et al. (2010). Bioanalysis in drug discovery and development. Pharmaceutical Methods 1 (1): 14–24.
  9. 9 Xu, X., Ray, R., Gu, Y. et al. (2004). Electrophoretic analysis and purification of fluorescent single‐walled carbon nanotube fragments. Journal of the American Chemical Society 126 (40): 12736–12737.
  10. 10 Zhu, X., Zhao, T., Nie, Z. et al. (2016). Nitrogen‐doped carbon nanoparticle modulated turn‐on fluorescent probes for histidine detection and its imaging in living cells. Nanoscale 8 (4): 2205–2211.
  11. 11 Sun, Y.‐P., Zhou, B., Lin, Y. et al. (2006). Quantum‐sized carbon dots for bright and colorful photoluminescence. Journal of the American Chemical Society 128 (24): 7756–7757.
  12. 12 Liu, H., Ye, T., and Mao, C. (2007). Fluorescent carbon nanoparticles derived from candle soot. Angewandte Chemie International Edition 46 (34): 6473–6475.
  13. 13 Ray, S.C., Saha, A., Jana, N.R. et al. (2009). Fluorescent carbon nanoparticles: synthesis, characterization, and bioimaging application. The Journal of Physical Chemistry C 113 (43): 18546–18551.
  14. 14 Zhou, J., Booker, C., Li, R. et al. (2007). An electrochemical avenue to blue luminescent nanocrystals from multiwalled carbon nanotubes (MWCNTs). Journal of the American Chemical Society 129 (4): 744–745.
  15. 15 Du, Y. and Guo, S. (2016). Chemically doped fluorescent carbon and graphene quantum dots for bioimaging, sensor, catalytic and photoelectronic applications. Nanoscale 8 (5): 2532–2543.
  16. 16 Peng, H. and Travas‐Sejdic, J. (2009). Simple aqueous solution route to luminescent carbogenic dots from carbohydrates. Chemistry of Materials 21 (23): 5563–5565.
  17. 17 Bourlinos, A.B., Stassinopoulos, A., Anglos, D. et al. (2008). Photoluminescent carbogenic dots. Chemistry of Materials 20 (14): 4539–4541.
  18. 18 Bourlinos, A.B., Stassinopoulos, A., Anglos, D. et al. (2008). Surface functionalized carbogenic quantum dots. Small 4 (4): 455–458.
  19. 19 Shi, Q.‐Q., Li, Y.‐H., Xu, Y. et al. (2014). High‐yield and high‐solubility nitrogen‐doped carbon dots: formation, fluorescence mechanism and imaging application. RSC Advances 4 (4): 1563–1566.
  20. 20 Fan, R.‐J., Sun, Q., Zhang, L. et al. (2014). Photoluminescent carbon dots directly derived from polyethylene glycol and their application for cellular imaging. Carbon 71 (Supplement C): 87–93.
  21. 21 Sachdev, A., Matai, I., Kumar, S.U. et al. (2013). A novel one‐step synthesis of PEG passivated multicolour fluorescent carbon dots for potential biolabeling application. RSC Advances 3 (38): 16958–16961.
  22. 22 Song, L., Cui, Y., Zhang, C. et al. (2016). Microwave‐assisted facile synthesis of yellow fluorescent carbon dots from o‐phenylenediamine for cell imaging and sensitive detection of Fe3+ and H2O2. RSC Advances 6 (21): 17704–17712.
  23. 23 De, B. and Karak, N. (2013). A green and facile approach for the synthesis of water soluble fluorescent carbon dots from banana juice. RSC Advances 3 (22): 8286–8290.
  24. 24 Qu, K., Wang, J., Ren, J. et al. (2013). Carbon dots prepared by hydrothermal treatment of dopamine as an effective fluorescent sensing platform for the label‐free detection of iron(III). ions and dopamine. Chemistry – A European Journal 19 (22): 7243–7299.
  25. 25 Sahu, S., Behera, B., Maiti, T.K. et al. (2012). Simple one‐step synthesis of highly luminescent carbon dots from orange juice: application as excellent bio‐imaging agents. Chemical Communications 48 (70): 8835–8837.
  26. 26 Yang, Z., Xu, M., Liu, Y. et al. (2014). Nitrogen‐doped, carbon‐rich, highly photoluminescent carbon dots from ammonium citrate. Nanoscale 6 (3): 1890–1895.
  27. 27 Krysmann, M.J., Kelarakis, A., Dallas, P. et al. (2011). Formation mechanism of carbogenic nanoparticles with dual photoluminescence emission. Journal of the American Chemical Society 134 (2): 747–750.
  28. 28 Xu, H., Yan, L., Nguyen, V. et al. (2017). One‐step synthesis of nitrogen‐doped carbon nanodots for ratiometric pH sensing by femtosecond laser ablation method. Applied Surface Science 414 (Supplement C): 238–243.
  29. 29 Zhu, Z., Wang, S., Chang, Y. et al. (2016). Direct photodissociation of toluene molecules to photoluminescent carbon dots under pulsed laser irradiation. Carbon 105: 416–423.
  30. 30 He, D., Zheng, C., Wang, Q. et al. (2015). Dielectric barrier discharge‐assisted one‐pot synthesis of carbon quantum dots as fluorescent probes for selective and sensitive detection of hydrogen peroxide and glucose. Talanta 142 (Supplement C): 51–56.
  31. 31 Niu, F., Xu, Y., Liu, J. et al. (2017). Controllable electrochemical/electroanalytical approach to generate nitrogen‐doped carbon quantum dots from varied amino acids: pinpointing the utmost quantum yield and the versatile photoluminescent and electrochemiluminescent applications. Electrochimica Acta 236 (Supplement C): 239–251.
  32. 32 Canevari, T.C., Nakamura, M., Cincotto, F.H. et al. (2016). High performance electrochemical sensors for dopamine and epinephrine using nanocrystalline carbon quantum dots obtained under controlled chronoamperometric conditions. Electrochimica Acta 209: 464–470.
  33. 33 Miao, X., Yan, X., Qu, D. et al. (2017). Red emissive sulfur, nitrogen codoped carbon dots and their application in ion detection and theraonostics. ACS Applied Materials & Interfaces 9 (22): 18549–18556.
  34. 34 Zhang, X., Lu, J., Li, X. et al. (2017). Facile heat reflux synthesis of blue luminescent carbon dots as optical nanoprobes for cellular imaging. New Journal of Chemistry 41 (2): 702–708.
  35. 35 Wang, Z., Xu, C., Lu, Y. et al. (2017). Visualization of adsorption: luminescent mesoporous silica‐carbon dots composite for rapid and selective removal of U(VI). and in situ monitoring the adsorption behavior. ACS Applied Materials & Interfaces 9 (7398): 7392–7398.
  36. 36 Wang, Z., Lu, Y., Yuan, H. et al. (2015). Microplasma‐assisted rapid synthesis of luminescent nitrogen‐doped carbon dots and their application in pH sensing and uranium detection. Nanoscale 7 (48): 20743–20748.
  37. 37 Sun, Q., Fang, S., Fang, Y. et al. (2017). Fluorometric detection of cholesterol based on β‐cyclodextrin functionalized carbon quantum dots via competitive host‐guest recognition. Talanta 167: 513–519.
  38. 38 Baig, M.M.F. and Chen, Y.‐C. (2017). Bright carbon dots as fluorescence sensing agents for bacteria and curcumin. Journal of Colloid and Interface Science 501: 341–349.
  39. 39 Yang, R., Guo, X., Jia, L. et al. (2017). Green preparation of carbon dots with mangosteen pulp for the selective detection of Fe3+ ions and cell imaging. Applied Surface Science 423 (Supplement C): 426–432.
  40. 40 Rong, M.‐C., Zhang, K.‐X., Wang, Y.‐R. et al. (2017). The synthesis of B, N‐carbon dots by a combustion method and the application of fluorescence detection for Cu2+ . Chinese Chemistry Letters 28 (5): 1119–1124.
  41. 41 Guo, L.‐P., Zhang, Y., and Li, W.‐C. (2017). Sustainable microalgae for the simultaneous synthesis of carbon quantum dots for cellular imaging and porous carbon for CO2 capture. Journal of Colloid and Interface Science 493 (Supplement C): 257–264.
  42. 42 Liu, Y., Zhou, Q., Li, J. et al. (2016). Selective and sensitive chemosensor for lead ions using fluorescent carbon dots prepared from chocolate by one‐step hydrothermal method. Sensors and Actuators B: Chemical 237 (Supplement C): 597–604.
  43. 43 Yao, W., Wu, N., Lin, Z. et al. (2017). Fluorescent turn‐off competitive immunoassay for biotin based on hydrothermally synthesized carbon dots. Microchim Acta 184 (3): 907–914.
  44. 44 Gupta, A. and Nandi, C.K. (2017). PC12 live cell ultrasensitive neurotransmitter signaling using high quantum yield Sulphur doped carbon dots and its extracellular Ca2+ ion dependence. Sensors and Actuators B: Chemical 245 (Supplement C): 137–145.
  45. 45 Xiao, Q., Liang, Y., Zhu, F. et al. (2017). Microwave‐assisted one‐pot synthesis of highly luminescent N‐doped carbon dots for cellular imaging and multi‐ion probing. Microchim Acta 184 (7): 2429–2438.
  46. 46 Choi, Y., Thongsai, N., Chae, A. et al. (2017). Microwave‐assisted synthesis of luminescent and biocompatible lysine‐based carbon quantum dots. Journal of Industrial and Engineering Chemistry 47 (Supplement C): 329–335.
  47. 47 Wang, J., Han, S., Fan, Z. et al. (2017). Carbon dots‐catalyzed chemiluminescence for the determination of trace isonaphthol. Journal of the Chinese Chemical Society 64 (5): 486–492.
  48. 48 Tian, T., Zhong, Y., Deng, C. et al. (2017). Brightly near‐infrared to blue emission tunable silver‐carbon dot nanohybrid for sensing of ascorbic acid and construction of logic gate. Talanta 162 (Supplement C): 135–142.
  49. 49 Li, H., Sun, C., Vijayaraghavan, R. et al. (2016). Long lifetime photoluminescence in N, S co‐doped carbon quantum dots from an ionic liquid and their applications in ultrasensitive detection of pesticides. Carbon 104 (Supplement C): 33–39.
  50. 50 Ortega‐Liebana, M.C., Chung, N.X., Limpens, R. et al. (2017). Uniform luminescent carbon nanodots prepared by rapid pyrolysis of organic precursors confined within nanoporous templating structures. Carbon 117 (Supplement C): 437–446.
  51. 51 Gu, Z.‐G., Li, D.‐J., Zheng, C. et al. (2017). MOF‐templated synthesis of ultrasmall photoluminescent carbon‐nanodot arrays for optical applications. Angewandte Chemie International Edition 56 (24): 6853–6858.
  52. 52 Baker, S.N. and Baker, G.A. (2010). Luminescent carbon nanodots: emergent nanolights. Angewandte Chemie International Edition 49 (38): 6726–6744.
  53. 53 Lv, P., Yao, Y., Li, D. et al. (2017). Self‐assembly of nitrogen‐doped carbon dots anchored on bacterial cellulose and their application in iron ion detection. Carbohydrate Polymers 172: 93–101.
  54. 54 Wang, T.‐Y., Chen, C.‐Y., Wang, C.‐M. et al. (2017). Multicolor functional carbon dots via one‐step refluxing synthesis. ACS Sensors 2 (3): 354–363.
  55. 55 Tang, Q., Zhu, W., He, B. et al. (2017). Rapid conversion from carbohydrates to large‐scale carbon quantum dots for all‐weather solar cells. ACS Nano 11 (2): 1540–1547.
  56. 56 Zhao, Y., Liu, X., Yang, Y. et al. (2015). Carbon dots: from intense absorption in visible range to excitation‐independent and excitation‐dependent photoluminescence. Fullerenes, Nanotubes and Carbon Nanostructures 23 (11): 922–929.
  57. 57 Wang, Y., Li, Y., Yan, Y. et al. (2013). Luminescent carbon dots in a new magnesium aluminophosphate zeolite. Chemical Communications 49 (79): 9006–9008.
  58. 58 Liu, M., Xu, Y., Niu, F. et al. (2016). Carbon quantum dots directly generated from electrochemical oxidation of graphite electrodes in alkaline alcohols and the applications for specific ferric ion detection and cell imaging. Analyst 141 (9): 2657–2664.
  59. 59 Bera, K., Sau, A., Mondal, P. et al. (2016). Metamorphosis of ruthenium‐doped carbon dots: in search of the origin of photoluminescence and beyond. Chemistry of Materials 28 (20): 7404–7413.
  60. 60 Sharma, A., Gadly, T., Gupta, A. et al. (2016). Origin of excitation dependent fluorescence in carbon nanodots. The Journal of Physical Chemistry Letters 7 (18): 3695–3702.
  61. 61 Huang, J.J., Zhong, Z.F., Rong, M.Z. et al. (2014). An easy approach of preparing strongly luminescent carbon dots and their polymer based composites for enhancing solar cell efficiency. Carbon 70 (Supplement C): 190–198.
  62. 62 Liu, Y., Xiao, N., Gong, N. et al. (2014). One‐step microwave‐assisted polyol synthesis of green luminescent carbon dots as optical nanoprobes. Carbon 68 (Supplement C): 258–264.
  63. 63 Jia, X., Li, J., and Wang, E. (2012). One‐pot green synthesis of optically pH‐sensitive carbon dots with upconversion luminescence. Nanoscale 4 (18): 5572–5575.
  64. 64 Zhang, Y.‐Y., Wu, M., Wang, Y.‐Q. et al. (2013). A new hydrothermal refluxing route to strong fluorescent carbon dots and its application as fluorescent imaging agent. Talanta 117 (Supplement C): 196–202.
  65. 65 Rong, M., Feng, Y., Wang, Y. et al. (2017). One‐pot solid phase pyrolysis synthesis of nitrogen‐doped carbon dots for Fe3+ sensing and bioimaging. Sensors and Actuators B: Chemical 245: 868–874.
  66. 66 Hou, J., Yan, J., Zhao, Q. et al. (2013). A novel one‐pot route for large‐scale preparation of highly photoluminescent carbon quantum dots powders. Nanoscale 5 (20): 9558–9561.
  67. 67 Wang, C., Xu, Z., Cheng, H. et al. (2015). A hydrothermal route to water‐stable luminescent carbon dots as nanosensors for pH and temperature. Carbon 82 (Supplement C): 87–95.
  68. 68 Jiang, K., Sun, S., Zhang, L. et al. (2015). Bright‐yellow‐emissive N‐doped carbon dots: preparation, cellular imaging, and bifunctional sensing. ACS Applied Materials & Interfaces 7 (41): 23231–23238.
  69. 69 Chen, D., Xu, M., Wu, W. et al. (2017). Multi‐color fluorescent carbon dots for wavelength‐selective and ultrasensitive Cu2+ sensing. Journal of Alloy and Compounds 701 (Supplement C): 75–81.
  70. 70 Guo, L., Ge, J., Liu, W. et al. (2016). Tunable multicolor carbon dots prepared from well‐defined polythiophene derivatives and their emission mechanism. Nanoscale 8 (2): 729–734.
  71. 71 Hu, S., Wang, Y., Zhang, W. et al. (2017). Multicolour emission states from charge transfer between carbon dots and surface molecules. Materials 10 (2): 165.
  72. 72 Wang, H., Sun, C., Chen, X. et al. (2017). Excitation wavelength independent visible color emission of carbon dots. Nanoscale 9 (5): 1909–1915.
  73. 73 Zhu, H., Wang, X., Li, Y. et al. (2009). Microwave synthesis of fluorescent carbon nanoparticles with electrochemiluminescence properties. Chemisrty of Communications (34): 5118–5120.
  74. 74 Tao, H.Q., Yang, K., Ma, Z. et al. (2012). In vivo NIR fluorescence imaging, biodistribution, and toxicology of photoluminescent carbon dots produced from carbon nanotubes and graphite. Small 8 (2): 281–290.
  75. 75 Wang, J., Wang, C.‐F., and Chen, S. (2012). Amphiphilic egg‐derived carbon dots: rapid plasma fabrication, pyrolysis process, and multicolor printing patterns. Angewandte Chemie International Edition 51 (37): 9297–9301.
  76. 76 Ding, H., Wei, J.‐S., and Xiong, H.‐M. (2014). Nitrogen and sulfur co‐doped carbon dots with strong blue luminescence. Nanoscale 6 (22): 13817–13823.
  77. 77 Hou, J., Wang, W., Zhou, T. et al. (2016). Synthesis and formation mechanistic investigation of nitrogen‐doped carbon dots with high quantum yields and yellowish‐green fluorescence. Nanoscale 8 (21): 11185–11193.
  78. 78 Shi, L., Yang, J.H., Zeng, H.B. et al. (2016). Carbon dots with high fluorescence quantum yield: the fluorescence originates from organic fluorophores. Nanoscale 8 (30): 14374–14378.
  79. 79 Dong, Y., Pang, H., Yang, H.B. et al. (2013). Carbon‐based dots co‐doped with nitrogen and sulfur for high quantum yield and excitation‐independent emission. Angewandte Chemie International Edition 52 (30): 7800–7804.
  80. 80 Liu, W., Xu, S., Liang, R. et al. (2017). In situ synthesis of nitrogen‐doped carbon dots in the interlayer region of a layered double hydroxide with tunable quantum yield. Journal of Materials Chemistry C 5 (14): 3536–3541.
  81. 81 Zhang, Y., Liu, X., Fan, Y. et al. (2016). One‐step microwave synthesis of N‐doped hydroxyl‐functionalized carbon dots with ultra‐high fluorescence quantum yields. Nanoscale 8 (33): 15281–15287.
  82. 82 Smith, A.M. and Nie, S. (2010). Semiconductor nanocrystals: structure, properties, and band gap engineering. Accounts of Chemical Research 43 (2): 190–200.
  83. 83 Li, H., He, X., Kang, Z. et al. (2010). Water‐soluble fluorescent carbon quantum dots and photocatalyst design. Angewandte Chemie International Edition 49 (26): 4430–4434.
  84. 84 Choi, Y., Kang, B., Lee, J. et al. (2016). Integrative approach toward uncovering the origin of photoluminescence in dual heteroatom‐doped carbon nanodots. Chemistry of Materials 28 (19): 6840–6847.
  85. 85 Dutta Choudhury, S., Chethodil, J.M., Gharat, P.M. et al. (2017). pH‐elicited luminescence functionalities of carbon dots: mechanistic insights. The Journal of Physical Chemistry Letters 8 (7): 1389–1395.
  86. 86 Wen, Z.‐H. and Yin, X.‐B. (2016). Excitation‐independent carbon dots, from photoluminescence mechanism to single‐color application. RSC Advances 6 (33): 27829–27835.
  87. 87 Yuan, Y.H., Liu, Z.X., Li, R.S. et al. (2016). Synthesis of nitrogen‐doping carbon dots with different photoluminescence properties by controlling the surface states. Nanoscale 8 (12): 6770–6776.
  88. 88 Zhu, S., Song, Y., Zhao, X. et al. (2015). The photoluminescence mechanism in carbon dots (graphene quantum dots, carbon nanodots, and polymer dots): current state and future perspective. Nano Research 8 (2): 355–381.
  89. 89 Gude, V., Das, A., Chatterjee, T. et al. (2016). Molecular origin of photoluminescence of carbon dots: aggregation‐induced orange‐red emission. Physical Chemistry Chemical Physics 18 (40): 28274–28280.
  90. 90 Demchenko, A.P. and Dekaliuk, M.O. (2016). The origin of emissive states of carbon nanoparticles derived from ensemble‐averaged and single‐molecular studies. Nanoscale 8 (29): 14057–14069.
  91. 91 Sharma, A., Gadly, T., Neogy, S. et al. (2017). Molecular origin and self‐assembly of fluorescent carbon nanodots in polar solvents. The Journal of Physical Chemistry Letters 8 (5): 1044–1052.
  92. 92 He, G., Shu, M., Yang, Z. et al. (2017). Microwave formation and photoluminescence mechanisms of multi‐states nitrogen doped carbon dots. Applied Surface Science 422 (Supplement C): 257–265.
  93. 93 Barati, A., Shamsipur, M., and Abdollahi, H. (2015). A misunderstanding about upconversion luminescence of carbon quantum dots. Journal of the Iranian Chemical Society 12 (3): 441–446.
  94. 94 Cao, L., Wang, X., Meziani, M.J. et al. (2007). Carbon dots for multiphoton bioimaging. Journal of the American Chemical Society 129 (37): 11318–11319.
  95. 95 Salinas‐Castillo, A., Ariza‐Avidad, M., Pritz, C. et al. (2013). Carbon dots for copper detection with down and upconversion fluorescent properties as excitation sources. Chemical Communications 49 (11): 1103–1105.
  96. 96 Denk, W., Strickler, J., and Webb, W. (1990). Two‐photon laser scanning fluorescence microscopy. Science 248 (4951): 73–76.
  97. 97 Wang, H., Yi, J., Yu, Y. et al. (2017). NIR upconversion fluorescence glucose sensing and glucose‐responsive insulin release of carbon dot‐immobilized hybrid microgels at physiological pH. Nanoscale 9 (2): 509–516.
  98. 98 Shen, J., Zhu, Y., Chen, C. et al. (2011). Facile preparation and upconversion luminescence of graphene quantum dots. Chemical Communications 47 (9): 2580–2582.
  99. 99 Barati, A., Shamsipur, M., and Abdollahi, H. (2015). A misunderstanding about upconversion luminescence of carbon quantum dots. Journal of the Iranian Chemical Society 12 (3): 441–446.
  100. 100 Wen, X., Yu, P., Toh, Y.‐R. et al. (2014). On the upconversion fluorescence in carbon nanodots and graphene quantum dots. Chemical Communications 50 (36): 4703–4706.
  101. 101 Gan, Z., Wu, X., Zhou, G. et al. (2013). Is there real upconversion photoluminescence from graphene quantum dots? Advanced Optical Materials 1 (8): 554–558.
  102. 102 Deng, Y., Zhao, D., Chen, X. et al. (2013). Long lifetime pure organic phosphorescence based on water soluble carbon dots. Chemical Communications 49 (51): 5751–5753.
  103. 103 Dong, X., Wei, L., Su, Y. et al. (2015). Efficient long lifetime room temperature phosphorescence of carbon dots in a potash alum matrix. Journal of Materials Chemistry C 3 (12): 2798–2801.
  104. 104 Tan, J., Zou, R., Zhang, J. et al. (2016). Large‐scale synthesis of N‐doped carbon quantum dots and their phosphorescence properties in a polyurethane matrix. Nanoscale 8 (8): 4742–4747.
  105. 105 Bai, L.Q., Xue, N., Wang, X.R. et al. (2017). Activating efficient room temperature phosphorescence of carbon dots by synergism of orderly non‐noble metals and dual structural confinements. Nanoscale 9 (20): 6658–6664.
  106. 106 Joseph, J. and Anappara, A.A. (2017). Cool white, persistent room‐temperature phosphorescence in carbon dots embedded in a silica gel matrix. Physical Chemistry Chemical Physics 19 (23): 15137–15144.
  107. 107 Li, Q., Zhou, M., Yang, Q. et al. (2016). Efficient room‐temperature phosphorescence from nitrogen‐doped carbon dots in composite matrices. Chemistry of Materials 28 (22): 8221–8227.
  108. 108 Yan, X., Chen, J.‐L., Su, M.‐X. et al. (2014). Phosphate‐containing metabolites switch on phosphorescence of ferric ion engineered carbon dots in aqueous solution. RSC Advances 4 (43): 22318–22323.
  109. 109 Gui, R., Jin, H., Wang, Z. et al. (2015). Room‐temperature phosphorescence logic gates developed from nucleic acid functionalized carbon dots and graphene oxide. Nanoscale 7 (18): 8289–8293.
  110. 110 Chen, Y., He, J., Hu, C. et al. (2017). Room temperature phosphorescence from moisture‐resistant and oxygen‐barred carbon dot aggregates. Journal of Materials Chemistry C 5 (25): 6243–6250.
  111. 111 Jiang, K., Wang, Y., Cai, C. et al. (2017). Activating room temperature long afterglow of carbon dots via covalent fixation. Chemistry of Materials 29 (11): 4866–4873.
  112. 112 Hu, S.‐L., Niu, K.‐Y., Sun, J. et al. (2009). One‐step synthesis of fluorescent carbon nanoparticles by laser irradiation. Journal of Materials Chemistry 19 (4): 484–488.
  113. 113 Xue, M., Zhan, Z., Zou, M. et al. (2016). Green synthesis of stable and biocompatible fluorescent carbon dots from peanut shells for multicolor living cell imaging. New Journal of Chemistry 40 (2): 1698–1703.
  114. 114 Yan, F., Kong, D., Luo, Y. et al. (2016). Carbon dots serve as an effective probe for the quantitative determination and for intracellular imaging of mercury(II). Microchim Acta 183 (5): 1611–1618.
  115. 115 Ramanan, V., Thiyagarajan, S.K., Raji, K. et al. (2016). Outright green synthesis of fluorescent carbon dots from eutrophic algal blooms for in vitro imaging. ACS Sustainable Chemistry & Engineering 4 (9): 4724–4761.
  116. 116 Gong, Y., Yu, B., Yang, W. et al. (2016). Phosphorus, and nitrogen co‐doped carbon dots as a fluorescent probe for real‐time measurement of reactive oxygen and nitrogen species inside macrophages. Biosensors and Bioelectronics 79: 822–828.
  117. 117 Yang, S.‐T., Cao, L., Luo, P.G. et al. (2009). Carbon dots for optical imaging in vivo. Journal of the American Chemical Society 131 (32): 11308–11309.
  118. 118 Yang, S.‐T., Wang, X., Wang, H. et al. (2009). Carbon dots as nontoxic and high‐performance fluorescence imaging agents. The Journal of Physical Chemistry C 113 (42): 18110–18114.
  119. 119 Jiang, C., Wu, H., Song, X. et al. (2014). Presence of photoluminescent carbon dots in Nescafe® original instant coffee: applications to bioimaging. Talanta 127: 68–74.
  120. 120 Yuan, M., Zhong, R., Gao, H. et al. (2015). One‐step, green, and economic synthesis of water‐soluble photoluminescent carbon dots by hydrothermal treatment of wheat straw, and their bio‐applications in labeling, imaging, and sensing. Applied Surface Science 355 (Supplement C): 1136–1144.
  121. 121 Ghosh, M., Sonkar, S.K., Saxena, M. et al. (2011). Carbon nano‐onions for imaging the life cycle of drosophila melanogaster. Small 7 (22): 3170–3177.
  122. 122 Qu, S., Wang, X., Lu, Q. et al. (2012). A biocompatible fluorescent ink based on water‐soluble luminescent carbon nanodots. Angewandte Chemie International Edition 51 (49): 12215–12218.
  123. 123 Huang, X., Zhang, F., Zhu, L. et al. (2013). Effect of injection routes on the biodistribution, clearance, and tumor uptake of carbon dots. ACS Nano 7 (7): 5684–5693.
  124. 124 Johnsson, B., Löfås, S., and Lindquist, G. (1991). Immobilization of proteins to a carboxymethyldextran‐modified gold surface for biospecific interaction analysis in surface plasmon resonance sensors. Analytical Biochemistry 198 (2): 268–277.
  125. 125 Montalbetti, C.A.G.N. and Falque, V. (2005). Amide bond formation and peptide coupling. Tetrahedron 61 (46): 10827–10852.
  126. 126 Fu, C., Qian, K., and Fu, A. (2017). Arginine‐modified carbon dots probe for live cell imaging and sensing by increasing cellular uptake efficiency. Materials Science and Engineering: C 76 (Supplement C): 350–355.
  127. 127 Lei, D., Yang, W., Gong, Y. et al. (2016). Non‐covalent decoration of carbon dots with folic acid via a polymer‐assisted strategy for fast and targeted cancer cell fluorescence imaging. Sensors and Actuators B: Chemical 230: 714–720.
  128. 128 Sun, T., Zheng, M., Xie, Z. et al. (2017). Supramolecular hybrids of carbon dots with doxorubicin: synthesis, stability and cellular trafficking. Materials Chemistry Frontiers 1 (2): 354–360.
  129. 129 Diac, A., Focsan, M., Socaci, C. et al. (2015). Covalent conjugation of carbon dots with Rhodamine B and assessment of their photophysical properties. RSC Advances 5 (95): 77662–77669.
  130. 130 Qu, D., Miao, X., Wang, X. et al. (2017). Se & N co‐doped carbon dots for high‐performance fluorescence imaging agent of angiography. Journal of Materials Chemistry B 5 (25): 4988–4992.
  131. 131 Wang, F., Hao, Q., Zhang, Y. et al. (2016). Fluorescence quenchometric method for determination of ferric ion using boron‐doped carbon dots. Microchim Acta 183 (1): 273–279.
  132. 132 Bourlinos, A.B., Trivizas, G., Karakassides, M.A. et al. (2015). Green and simple route toward boron doped carbon dots with significantly enhanced non‐linear optical properties. Carbon 83 (Supplement C): 173–179.
  133. 133 Dai, B., Wu, C., Lu, Y. et al. (2017). Synthesis and formation mechanism of s‐doped carbon dots from low‐molecule‐weight organics. Journal of Luminescence 190 (Supplement C): 108–114.
  134. 134 Li, Y., Lin, H., Luo, C. et al. (2017). Aggregation induced red shift emission of phosphorus doped carbon dots. RSC Advances 7 (51): 32225–32228.
  135. 135 Cheng, J., Wang, C.‐F., Zhang, Y. et al. (2016). Zinc ion‐doped carbon dots with strong yellow photoluminescence. RSC Advances 6 (43): 37189–37194.
  136. 136 Xu, Q., Wei, J., Wang, J. et al. (2016). Facile synthesis of copper doped carbon dots and their application as a “turn‐off” fluorescent probe in the detection of Fe3+ ions. RSC Advances 6 (34): 28745–28750.
  137. 137 Zhang, H.‐Y., Wang, Y., Xiao, S. et al. (2017). Rapid detection of Cr(VI). ions based on cobalt(II)‐doped carbon dots. Biosensors and Bioelectronics 87 (Supplement C): 46–52.
  138. 138 Wang, Y., Meng, H., Jia, M. et al. (2016). Intraparticle FRET of Mn(II)‐doped carbon dots and its application in discrimination of volatile organic compounds. Nanoscale 8 (39): 17190–17195.
  139. 139 Yuan, Y.H., Li, R.S., Wang, Q. et al. (2015). Germanium‐doped carbon dots as a new type of fluorescent probe for visualizing the dynamic invasions of mercury(II). ions into cancer cells. Nanoscale 7 (40): 16841–16847.
  140. 140 Chen, B.B., Liu, Z.X., Zou, H.Y. et al. (2016). Highly selective detection of 2,4,6‐trinitrophenol by using newly developed terbium‐doped blue carbon dots. Analyst 141 (9): 2676–2681.
  141. 141 Du, F., Zhang, L., Zhang, L. et al. (2017). Engineered gadolinium‐doped carbon dots for magnetic resonance imaging‐guided radiotherapy of tumors. Biomaterials 121: 109–120.
  142. 142 Yuan, Y., Zhao, X., Qiao, M. et al. (2016). Determination of sunset yellow in soft drinks based on fluorescence quenching of carbon dots. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 167: 106–110.
  143. 143 Liu, Y., Duan, W., Song, W. et al. (2017). 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 Applied Materials & Interfaces 9 (14): 12663–12672.
  144. 144 Zhong, D., Yang, K., Wang, Y. et al. (2017). Dual‐channel sensing strategy based on gold nanoparticles cooperating with carbon dots and hairpin structure for assaying RNA and DNA. Talanta 175: 217–223.
  145. 145 Yang, K., Wang, S., Wang, Y. et al. (2017). Dual‐channel probe of carbon dots cooperating with gold nanoclusters employed for assaying multiple targets. Biosensors and Bioelectronics 91: 566–573.
  146. 146 Zhang, Y., Zhang, J., Zhang, J. et al. (2017). Intense enhancement of yellow luminescent carbon dots coupled with gold nanoparticles toward white LED. Dyes and Pigments 140: 122–130.
  147. 147 Abdelhamid, H.N., Talib, A., and Wu, H.‐F. (2017). One pot synthesis of gold – carbon dots nanocomposite and its application for cytosensing of metals for cancer cells. Talanta 166: 357–363.
  148. 148 Wang, B., Chen, Y., Wu, Y. et al. (2016). Aptamer induced assembly of fluorescent nitrogen‐doped carbon dots on gold nanoparticles for sensitive detection of AFB1. Biosensors and Bioelectronics 78: 23–30.
  149. 149 Sellergren, B. and Hall, A.J. (2012). Molecularly imprinted polymers. In: Supramolecular Chemistry . John Wiley & Sons, Ltd.
  150. 150 Zuo, P., Gao, J., Peng, J. et al. (2016). A sol‐gel based molecular imprint incorporating carbon dots for fluorometric determination of nicotinic acid. Microchim Acta 183 (1): 329–336.
  151. 151 Gogoi, N., Barooah, M., Majumdar, G. et al. (2015). Carbon dots rooted agarose hydrogel hybrid platform for optical detection and separation of heavy metal ions. ACS Applied Materials & Interfaces 7 (5): 3058–3067.
  152. 152 Konwar, A., Gogoi, N., Majumdar, G. et al. (2015). Green chitosan–carbon dots nanocomposite hydrogel film with superior properties. Carbohydrate Polymers 115 (Supplement C): 238–245.
  153. 153 Lu, H., Ren, S., Zhang, P. et al. (2017). Laser‐textured surface storing a carbon dots/poly(ethylene glycol)/chitosan gel with slow‐release lubrication effect. RSC Advances 7 (35): 21600–21606.
  154. 154 Chen, B. and Feng, J. (2015). White‐light‐emitting polymer composite film based on carbon dots and lanthanide complexes. The Journal of Physical Chemistry C 119 (14): 7865–7872.
  155. 155 Sun, C., Zhang, Y., Kalytchuk, S. et al. (2015). Down‐conversion monochromatic light‐emitting diodes with the color determined by the active layer thickness and concentration of carbon dots. Journal of Materials Chemistry C 3 (26): 6613–6615.
  156. 156 Liu, Z., Zhang, X., Cui, L. et al. (2017). Development of a highly sensitive electrochemiluminescence sophoridine sensor using Ru(bpy)3 2+ integrated carbon quantum dots – polyvinyl alcohol composite film. Sensors and Actuators B: Chemical 248 (Supplement C): 402–410.
  157. 157 Ma, L., Xiang, W., Gao, H. et al. (2015). Carbon dot‐doped sodium borosilicate gel glasses with emission tunability and their application in white light emitting diodes. Journal of Materials Chemistry C 3 (26): 6764–6770.
  158. 158 Vassilakopoulou, A., Georgakilas, V., Vainos, N. et al. (2017). Successful entrapment of carbon dots within flexible free‐standing transparent mesoporous organic‐inorganic silica hybrid films for photonic applications. Journal of Physics and Chemistry Solids 103 (Supplement C): 190–196.
  159. 159 Saenger, W. (1984). DNA structure. In: Principles of Nucleic Acid Structure , 253–282. New York, NY: Springer New York.
  160. 160 Ding, H., Du, F., Liu, P. et al. (2015). DNA–carbon dots function as fluorescent vehicles for drug delivery. ACS Applied Materials & Interfaces 7 (12): 6889–6897.
  161. 161 Singh, S., Mishra, A., Kumari, R. et al. (2017). Carbon dots assisted formation of DNA hydrogel for sustained release of drug. Carbon 114 (Supplement C): 169–176.
  162. 162 Wu, L., Ma, C., Zheng, X. et al. (2015). Paper‐based electrochemiluminescence origami device for protein detection using assembled cascade DNA–carbon dots nanotags based on rolling circle amplification. Biosensors and Bioelectronics 68: 413–420.
  163. 163 Li, Z., Ni, Y., and Kokot, S. (2015). A new fluorescent nitrogen‐doped carbon dot system modified by the fluorophore‐labeled ssDNA for the analysis of 6‐mercaptopurine and Hg(II). Biosensors and Bioelectronics 74: 91–97.
  164. 164 Cui, X., Zhu, L., Wu, J. et al. (2015). A fluorescent biosensor based on carbon dots‐labeled oligodeoxyribonucleotide and graphene oxide for mercury (II). detection. Biosensors and Bioelectronics 63: 506–512.
  165. 165 Khakbaz, F. and Mahani, M. (2017). Micro‐RNA detection based on fluorescence resonance energy transfer of DNA‐carbon quantum dots probes. Analytical Biochemistry 523: 32–38.
  166. 166 Liu, Q., Ma, C., Liu, X.‐P. et al. (2017). A novel electrochemiluminescence biosensor for the detection of microRNAs based on a DNA functionalized nitrogen doped carbon quantum dots as signal enhancers. Biosensors and Bioelectronics 92: 273–279.
  167. 167 Jayasena, S.D. (1999). Aptamers: an emerging class of molecules that rival antibodies in diagnostics. Clinical Chemistry 45 (9): 1628–1650.
  168. 168 Wang, Y., Bao, L., Liu, Z. et al. (2011). Aptamer biosensor based on fluorescence resonance energy transfer from upconverting phosphors to carbon nanoparticles for thrombin detection in human plasma. Analytical Chemistry 83 (21): 8130–8137.
  169. 169 Zhang, L., Cui, P., Zhang, B. et al. (2013). Aptamer‐based turn‐on detection of thrombin in biological fluids based on efficient phosphorescence energy transfer from Mn‐doped ZnS quantum dots to carbon nanodots. Chemistry – A European Journal 19 (28): 9242–9250.
  170. 170 Xu, B., Zhao, C., Wei, W. et al. (2012). Aptamer carbon nanodot sandwich used for fluorescent detection of protein. Analyst 137 (23): 5483–5466.
  171. 171 Wang, K., Yao, H., Meng, Y. et al. (2015). Specific aptamer‐conjugated mesoporous silica–carbon nanoparticles for HER2‐targeted chemo‐photothermal combined therapy. Acta Biomaterialia 16 (Supplement C): 196–205.
  172. 172 Wang, R., Xu, Y., Zhang, T. et al. (2015). Rapid and sensitive detection of Salmonella typhimurium using aptamer‐conjugated carbon dots as fluorescence probe. Analytical Methods 7 (5): 1701–1706.
  173. 173 Ma, N., Jiang, W., Li, T. et al. (2015). Fluorescence aggregation assay for the protein biomarker Mucin 1 using carbon dot‐labeled antibodies and aptamers. Microchim Acta 182 (1): 443–447.
  174. 174 Ding, Y., Ling, J., Wang, H. et al. (2015). Fluorescent detection of Mucin 1 protein based on aptamer functionalized biocompatible carbon dots and graphene oxide. Analytical Methods 7 (18): 7792–7798.
  175. 175 Miao, H., Wang, L., Zhuo, Y. et al. (2016). Label‐free fluorimetric detection of CEA using carbon dots derived from tomato juice. Biosensors and Bioelectronics 86: 83–89.
  176. 176 Zhu, L., Xu, G., Song, Q. et al. (2016). Highly sensitive determination of dopamine by a turn‐on fluorescent biosensor based on aptamer labeled carbon dots and nano‐graphite. Sensors and Actuators B: Chemical 231: 506–512.
  177. 177 Wu, L., Zhang, Y., Wang, Y. et al. (2016). A paper‐based electrochemiluminescence electrode as an aptamer‐based cytosensor using PtNi@carbon dots as nanolabels for detection of cancer cells and for in‐situ screening of anticancer drugs. Microchim Acta 183 (6): 1873–1880.
  178. 178 Su, M., Liu, H., Ge, L. et al. (2014). Aptamer‐based electrochemiluminescent detection of MCF‐7 cancer cells based on carbon quantum dots coated mesoporous silica nanoparticles. Electrochimica Acta 146: 262–269.
  179. 179 Lee, C.H., Rajendran, R., Jeong, M.‐S. et al. (2013). Bioimaging of targeting cancers using aptamer‐conjugated carbon nanodots. Chemstry of Communications 49 (58): 6543–6555.
  180. 180 Zhang, J., Zheng, M., and Xie, Z. (2016). Co‐assembled hybrids of proteins and carbon dots for intracellular protein delivery. Journal of Materials Chemistry B 4 (34): 5659–5663.
  181. 181 Li, S., Peng, Z., and Leblanc, R.M. (2015). Method to determine protein concentration in the protein–nanoparticle conjugates aqueous solution using circular dichroism spectroscopy. Analytical Chemistry 87 (13): 6455–6459.
  182. 182 Peng, Z., Li, S., Han, X. et al. (2016). Determination of the composition, encapsulation efficiency and loading capacity in protein drug delivery systems using circular dichroism spectroscopy. Analytica Chimica Acta 937 (Supplement C): 113–118.
  183. 183 Liu, P., Zhang, C., Liu, X. et al. (2016). Preparation of carbon quantum dots with a high quantum yield and the application in labeling bovine serum albumin. Applied Surface Science 368 (Supplement C): 122–128.
  184. 184 Yang, L., Jiang, W., Qiu, L. et al. (2015). One pot synthesis of highly luminescent polyethylene glycol anchored carbon dots functionalized with a nuclear localization signal peptide for cell nucleus imaging. Nanoscale 7 (14): 6104–6113.
  185. 185 Patra, S., Roy, E., Madhuri, R. et al. (2016). The next generation cell‐penetrating peptide and carbon dot conjugated nano‐liposome for transdermal delivery of curcumin. Biomaterials Science 4 (3): 418–429.
  186. 186 Roy, E., Patra, S., Madhuri, R. et al. (2017). Carbon dot/TAT peptide co‐conjugated bubble nanoliposome for multicolor cell imaging, nuclear‐targeted delivery, and chemo/photothermal synergistic therapy. Chemical Engineering Journal 312: 144–157.
  187. 187 Li, H., Kong, W., Liu, J. et al. (2014). Carbon dots for photoswitching enzyme catalytic activity. Journal of Materials Chemistry B 2 (34): 5652–5658.
  188. 188 Das, K., Maiti, S., and Das, P.K. (2014). Probing enzyme location in water‐in‐oil microemulsion using enzyme–carbon dot conjugates. Langmuir 30 (9): 2448–2459.
  189. 189 Anilkumar, P., Cao, L., Yu, J.‐J. et al. (2013). Crosslinked carbon dots as ultra‐bright fluorescence probes. Small 9 (4): 545–551.
  190. 190 Li, Q., Ohulchanskyy, T.Y., Liu, R. et al. (2010). Photoluminescent carbon dots as biocompatible nanoprobes for targeting cancer cells in vitro. The Journal of Physical Chemistry C 114 (28): 12062–12068.
  191. 191 Liu, J.‐H., Cao, L., LeCroy, G.E. et al. (2015). Carbon “quantum” dots for fluorescence labeling of cells. ACS Applied Materials & Interfaces 7 (34): 19439–19445.
  192. 192 Kichler, A., Leborgne, C., Coeytaux, E. et al. (2001). Polyethylenimine‐mediated gene delivery: a mechanistic study. Journal of Gene Medicine 3 (2): 135–144.
  193. 193 Wang, H.‐J., He, X., Luo, T.‐Y. et al. (2017). Amphiphilic carbon dots as versatile vectors for nucleic acid and drug delivery. Nanoscale 9 (18): 5935–5947.
  194. 194 Wu, Y.‐F., Wu, H.‐C., Kuan, C.‐H. et al. (2016). Multi‐functionalized carbon dots as theranostic nanoagent for gene delivery in lung cancer therapy. Scientific Reports 6: 21170.
  195. 195 Wang, L., Wang, X., Bhirde, A. et al. (2014). Carbon‐dot‐based two‐photon visible nanocarriers for safe and highly efficient delivery of siRNA and DNA. Advanced Healthcare Materials 3 (8): 1203–1209.
  196. 196 Jin, H., Gui, R., Wang, Y. et al. (2017). Carrot‐derived carbon dots modified with polyethyleneimine and nile blue for ratiometric two‐photon fluorescence turn‐on sensing of sulfide anion in biological fluids. Talanta 169: 141–148.
  197. 197 Gui, R., Jin, H., Wang, Y. et al. (2017). Ions‐induced two‐photon fluorescence dual‐switching for reversible and simultaneous sensing of Cu2+ and Hg2+ based on dual‐emitting carbon dot/carbon dot conjugates. Sensors and Actuators B: Chemical 245 (Supplement C): 386–394.
  198. 198 Li, M., Hu, C., Yu, C. et al. (2015). Organic amine‐grafted carbon quantum dots with tailored surface and enhanced photoluminescence properties. Carbon 91 (Supplement C): 291–297.
  199. 199 Dong, W., Zhou, S., Dong, Y. et al. (2015). The preparation of ethylenediamine‐modified fluorescent carbon dots and their use in imaging of cells. Luminescence 30 (6): 867–871.
  200. 200 Dong, W., Dong, Y., Wang, Y. et al. (2013). Labeling of human hepatocellular carcinoma cells by hexamethylene diamine modified fluorescent carbon dots. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 116 (Supplement C): 209–213.
  201. 201 Huang, Y., Zhou, J., Feng, H. et al. (2016). A dual‐channel fluorescent chemosensor for discriminative detection of glutathione based on functionalized carbon quantum dots. Biosensors and Bioelectronics 86: 748–755.
  202. 202 Matai, I., Sachdev, A., and Gopinath, P. (2015). Self‐assembled hybrids of fluorescent carbon dots and PAMAM dendrimers for epirubicin delivery and intracellular imaging. ACS Applied Materials & Interfaces 7 (21): 11423–11435.
  203. 203 Dorcéna, C.J., Olesik, K.M., Wetta, O.G. et al. (2013). Characterization and toxicity of carbon dots‐poly(lactic‐co‐glycolic acid). nanocomposites for biomedical imaging. Nano Life 03 (01): 1340002.
  204. 204 Zou, Y., Yan, F., Zheng, T. et al. (2015). Highly luminescent organosilane‐functionalized carbon dots as a nanosensor for sensitive and selective detection of quercetin in aqueous solution. Talanta 135: 145–148.
  205. 205 Wang, W., Kim, T., Yan, Z. et al. (2015). Carbon dots functionalized by organosilane with double‐sided anchoring for nanomolar Hg2+ detection. Journal of Colloid and Interface Science 437 (Supplement C): 28–34.
  206. 206 Wang, F., Xie, Z., Zhang, H. et al. (2011). Highly luminescent organosilane‐functionalized carbon dots. Advanced Functional Materials 21 (6): 1027–1031.
  207. 207 Wang, D., Liu, J., Chen, J.‐F. et al. (2016). Surface functionalization of carbon dots with polyhedral oligomeric silsesquioxane (POSS). for multifunctional applications. Advanced Materials Interfaces 3 (1): 1500439.
  208. 208 Wang, W.‐J., Hai, X., Mao, Q.‐X. et al. (2015). Polyhedral oligomeric silsesquioxane functionalized carbon dots for cell imaging. ACS Applied Materials & Interfaces 7 (30): 16609–16616.
  209. 209 Turner, A.P.F., Karube, I., and Wilson, G.S. (1987). Biosensors: Fundamentals and Applications , 770. Oxford: Oxford University Press.
  210. 210 Mohd Yazid, S., Chin, S., Pang, S. et al. (2013). Detection of Sn(II). ions via quenching of the fluorescence of carbon nanodots. Microchim Acta 180 (1–2): 137–143.
  211. 211 Xu, H., Yang, X., Li, G. et al. (2015). Green synthesis of fluorescent carbon dots for selective detection of tartrazine in food samples. Journal of Agricultural and Food Chemistry 63 (30): 6707–6714.
  212. 212 Xu, J., Sahu, S., Cao, L. et al. (2012). Efficient fluorescence quenching in carbon dots by surface‐doped metals – disruption of excited state redox processes and mechanistic implications. Langmuir 28 (46): 16141–16147.
  213. 213 Mohapatra, S., Sahu, S., Sinha, N. et al. (2015). Synthesis of a carbon‐dot‐based photoluminescent probe for selective and ultrasensitive detection of Hg2+ in water and living cells. Analyst 140 (4): 1221–1228.
  214. 214 Sapsford, K.E., Berti, L., and Medintz, I.L. (2006). Materials for fluorescence resonance energy transfer analysis: beyond traditional donor–acceptor combinations. Angewandte Chemie International Edition 45 (28): 4562–4589.
  215. 215 Campos, B.B., Contreras‐Cáceres, R., Bandosz, T.J. et al. (2017). Carbon dots coated with vitamin B12 as selective ratiometric nanosensor for phenolic carbofuran. Sensors and Actuators B: Chemical 239: 553–561.
  216. 216 Wang, Y., Jiang, K., Zhu, J. et al. (2015). A FRET‐based carbon dot‐MnO2 nanosheet architecture for glutathione sensing in human whole blood samples. Chemical Communications 51 (64): 12748–12751.
  217. 217 Dai, H., Shi, Y., Wang, Y. et al. (2014). A carbon dot based biosensor for melamine detection by fluorescence resonance energy transfer. Sensors and Actuators B: Chemical 202: 201–208.
  218. 218 Amjadi, M., Abolghasemi‐Fakhri, Z., and Hallaj, T. (2015). Carbon dots‐silver nanoparticles fluorescence resonance energy transfer system as a novel turn‐on fluorescent probe for selective determination of cysteine. Journal of Photochemistry and Photobiology A: Chemistry 309: 8–14.
  219. 219 Li, Q., Zhang, L., Li, J. et al. (2011). Nanomaterial‐amplified chemiluminescence systems and their applications in bioassays. TrAC Trends in Analytical Chemistry 30 (2): 401–413.
  220. 220 Lin, Z., Xue, W., Chen, H. et al. (2011). Peroxynitrous‐acid‐induced chemiluminescence of fluorescent carbon dots for nitrite sensing. Analytical Chemistry 83 (21): 8245–8251.
  221. 221 Lin, Z., Xue, W., Chen, H. et al. (2012). Classical oxidant induced chemiluminescence of fluorescent carbon dots. Chemical Communications 48 (7): 1051–1053.
  222. 222 Amjadi, M., Manzoori, J.L., Hallaj, T. et al. (2014). Direct chemiluminescence of carbon dots induced by potassium ferricyanide and its analytical application. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 122: 715–720.
  223. 223 Shi, J., Lu, C., Yan, D. et al. (2013). High selectivity sensing of cobalt in HepG2 cells based on necklace model microenvironment‐modulated carbon dot‐improved chemiluminescence in Fenton‐like system. Biosensors and Bioelectronics 45: 58–64.
  224. 224 Amjadi, M., Manzoori, J.L., Hallaj, T. et al. (2014). Strong enhancement of the chemiluminescence of the Cerium(IV)‐thiosulfate reaction by carbon dots, and its application to the sensitive determination of dopamine. Microchim Acta 181 (5): 671–677.
  225. 225 Maloy, J.T. (2004). ECL theory. In: Electrogenerated Chemiluminescence (ed. A.J. Bard), 101–162. CRC Press.
  226. 226 Carrara, S., Arcudi, F., Prato, M. et al. (2017). Amine‐rich nitrogen‐doped carbon nanodots as a platform for self‐enhancing electrochemiluminescence. Angewandte Chemie International Edition 56 (17): 4757–4761.
  227. 227 Long, Y.M., Bao, L., Zhao, J.Y. et al. (2014). Revealing carbon nanodots as coreactants of the anodic electrochemiluminescence of Ru(bpy)3 2+ . Analytical Chemistry 86 (15): 7224–7228.
  228. 228 Li, L., Yu, B., Zhang, X. et al. (2015). A novel electrochemiluminescence sensor based on Ru(bpy)3 2+ /N‐doped carbon nanodots system for the detection of bisphenol A. Analytica Chimica Acta 895: 104–111.
  229. 229 Xu, Z., Yu, J., and Liu, G. (2013). Fabrication of carbon quantum dots and their application for efficient detecting Ru(bpy)3 2+ in the solution. Sensors and Actuators B: Chemical 181: 209–214.
  230. 230 Li, L., Liu, D., Mao, H. et al. (2017). Multifunctional solid‐state electrochemiluminescence sensing platform based on poly(ethylenimine). capped N‐doped carbon dots as novel co‐reactant. Biosensors and Bioelectronics 89 (Pt 1): 489–495.
  231. 231 Zhang, P., Xue, Z., Luo, D. et al. (2014). Dual‐peak electrogenerated chemiluminescence of carbon dots for iron ions detection. Analytical Chemistry 86 (12): 5620–5623.
  232. 232 Stradiotto, N.R., Yamanaka, H., and Zanoni, M.V.B. (2003). Electrochemical sensors: a powerful tool in analytical chemistry. Journal of the Brazilian Chemical Society 14: 159–173.
  233. 233 Li, Y., Zhong, Y., Zhang, Y. et al. (2015). Carbon quantum dots/octahedral Cu2O nanocomposites for non‐enzymatic glucose and hydrogen peroxide amperometric sensor. Sensors and Actuators B: Chemical 206: 735–743.
  234. 234 Huang, Q., Lin, X., Lin, C. et al. (2015). A high performance electrochemical biosensor based on Cu2O‐carbon dots for selective and sensitive determination of dopamine in human serum. RSC Advances 5 (67): 54102–54108.
  235. 235 Huang, Q., Lin, X., Zhu, J.‐J. et al. (2017). Pd‐Au@carbon dots nanocomposite: facile synthesis and application as an ultrasensitive electrochemical biosensor for determination of colitoxin DNA in human serum. Biosensors and Bioelectronics 94: 507–512.
  236. 236 Gao, Q., Han, J., and Ma, Z. (2013). Polyamidoamine dendrimers‐capped carbon dots/Au nanocrystal nanocomposites and its application for electrochemical immunosensor. Biosensors and Bioelectronics 49: 323–328.
  237. 237 Hu, S., Huang, Q., Lin, Y. et al. (2014). Reduced graphene oxide‐carbon dots composite as an enhanced material for electrochemical determination of dopamine. Electrochimica Acta 130: 805–809.
  238. 238 Wang, Y., Wang, Z., Rui, Y. et al. (2015). Horseradish peroxidase immobilization on carbon nanodots/CoFe layered double hydroxides: direct electrochemistry and hydrogen peroxide sensing. Biosensors and Bioelectronics 64: 57–62.
  239. 239 Mamtani, R., Stern, P., Dawood, I. et al. (2011). Metals and disease: a global primary health care perspective. Journal of Toxicology 2011.
  240. 240 Li, H., Zhai, J., Tian, J. et al. (2011). Carbon nanoparticle for highly sensitive and selective fluorescent detection of Mercury(II). ion in aqueous solution. Biosensors and Bioelectronics 26 (12): 4656–4660.
  241. 241 Gonçalves, H., Jorge, P.A.S., Fernandes, J.R.A. et al. (2010). Hg(II). sensing based on functionalized carbon dots obtained by direct laser ablation. Sensors and Actuators B: Chemical 145 (2): 702–707.
  242. 242 Gonçalves, H.M.R., Duarte, A.J., and Esteves da Silva, J.C.G. (2010). Optical fiber sensor for Hg(II). based on carbon dots. Biosensors and Bioelectronics 26 (4): 1302–1306.
  243. 243 Liu, Y., Liu, C.Y., and Zhang, Z.Y. (2012). Synthesis of highly luminescent graphitized carbon dots and the application in the Hg2+ detection. Applied Surface Science 263: 481–485.
  244. 244 Gonçalves, H. and Esteves da Silva, J.G. (2010). Fluorescent carbon dots capped with PEG200 and mercaptosuccinic acid. Journal of Fluorescence 20 (5): 1023–1028.
  245. 245 Stricks, W. and Kolthoff, I.M. (1953). Reactions between mercuric mercury and cysteine and glutathione. Apparent dissociation constants, heats and entropies of formation of various forms of mercuric mercapto‐cysteine and ‐glutathione. Journal of the American Chemical Society 75 (22): 5673–5681.
  246. 246 Gonçalves, H.M.R., Duarte, A.J., Davis, F. et al. (2012). Layer‐by‐layer immobilization of carbon dots fluorescent nanomaterials on single optical fiber. Analytica Chimica Acta 735: 90–95.
  247. 247 Wee, S.S., Ng, Y.H., and Ng, S.M. (2013). Synthesis of fluorescent carbon dots via simple acid hydrolysis of bovine serum albumin and its potential as sensitive sensing probe for lead (II). ions. Talanta 116: 71–76.
  248. 248 Patidar, R., Rebary, B., Bhadu, G.R. et al. (2016). Fluorescent carbon nanoparticles as label‐free recognizer of Hg2+ and Fe3+ through effective fluorescence quenching in aqueous media. Journal of Luminescence 173: 243–249.
  249. 249 Li, X., Zhang, S., Kulinich, S.A. et al. (2014). Engineering surface states of carbon dots to achieve controllable luminescence for solid‐luminescent composites and sensitive Be2+ detection. Scientific Reports 4: 4976.
  250. 250 Chen, J., Li, Y., Lv, K. et al. (2016). Cyclam‐functionalized carbon dots sensor for sensitive and selective detection of copper(II). ion and sulfide anion in aqueous media and its imaging in live cells. Sensors and Actuators B: Chemical 224: 298–306.
  251. 251 Zhu, A., Qu, Q., Shao, X. et al. (2012). Carbon‐dot‐based dual‐emission nanohybrid produces a ratiometric fluorescent sensor for in vivo imaging of cellular copper ions. Angewandte Chemie 124 (29): 7297–7301.
  252. 252 Liu, Y., Zhao, Y., and Zhang, Y. (2014). One‐step green synthesized fluorescent carbon nanodots from bamboo leaves for copper(II). ion detection. Sensors and Actuators B: Chemical 196: 647–652.
  253. 253 Ye, Z., Tang, R., Wu, H. et al. (2014). Preparation of europium complex‐conjugated carbon dots for ratiometric fluorescence detection of copper(II). ions. New Journal of Chemistry 38 (12): 5721–5726.
  254. 254 Gedda, G., Lee, C.‐Y., Lin, Y.‐C. et al. (2016). Green synthesis of carbon dots from prawn shells for highly selective and sensitive detection of copper ions. Sensors and Actuators B: Chemical 224: 396–403.
  255. 255 Zhou, Y. and Ma, Z. (2016). A novel fluorescence enhanced route to detect copper(II). by click chemistry‐catalyzed connection of Au@SiO2 and carbon dots. Sensors and Actuators B: Chemical 233: 426–430.
  256. 256 Samadi‐Maybodi, A. and Rezaei, V. (2014). A new sol–gel optical sensor with nonporous structure for determination of trace zinc. Sensors and Actuators B: Chemical 199: 418–423.
  257. 257 Zhang, Z., Shi, Y., Pan, Y. et al. (2014). Quinoline derivative‐functionalized carbon dots as a fluorescent nanosensor for sensing and intracellular imaging of Zn2+ . Journal of Materials Chemistry B 2 (31): 5020–5027.
  258. 258 Gupta, A., Verma, N.C., Khan, S. et al. (2016). Carbon dots for naked eye colorimetric ultrasensitive arsenic and glutathione detection. Biosensors and Bioelectronics 81: 465–472.
  259. 259 Sharma, V., Saini, A.K., and Mobin, S.M. (2016). Multicolour fluorescent carbon nanoparticle probes for live cell imaging and dual palladium and mercury sensors. Journal of Materials Chemistry B 4 (14): 2466–2476.
  260. 260 Kim, Y., Jang, G., and Lee, T.S. (2015). New fluorescent metal‐ion detection using a paper‐based sensor strip containing tethered rhodamine carbon nanodots. ACS Applied Materials & Interfaces 7 (28): 15649–15657.
  261. 261 Barati, A., Shamsipur, M., and Abdollahi, H. (2016). Metal‐ion‐mediated fluorescent carbon dots for indirect detection of sulfide ions. Sensors and Actuators B: Chemical 230: 289–297.
  262. 262 Ray, P.D., Huang, B.‐W., and Tsuji, Y. (2012). Reactive oxygen species (ROS). homeostasis and redox regulation in cellular signaling. Cellular Signalling 24 (5): 981–990.
  263. 263 Simões, E.F.C., Leitão, J.M.M., and Esteves da Silva, J.C.G. (2017). Sulfur and nitrogen co‐doped carbon dots sensors for nitric oxide fluorescence quantification. Analytica Chimica Acta 960: 117–122.
  264. 264 Bhattacharya, S., Sarkar, R., Chakraborty, B. et al. (2017). Nitric oxide sensing through azo‐dye formation on carbon dots. ACS Sensors 2 (8): 1215–1224.
  265. 265 Simões, E.F.C., Esteves da Silva, J.C.G., and Leitão, J.M.M. (2015). Peroxynitrite and nitric oxide fluorescence sensing by ethylenediamine doped carbon dots. Sensors and Actuators B: Chemical 220: 1043–1049.
  266. 266 Simões, E.F.C., da Silva, J.C.G.E., and Leitão, J.M.M. (2014). Carbon dots from tryptophan doped glucose for peroxynitrite sensing. Analytica Chimica Acta 852: 174–180.
  267. 267 Gao, X., Ding, C., Zhu, A. et al. (2014). Carbon‐dot‐based ratiometric fluorescent probe for imaging and biosensing of superoxide anion in live cells. Analytical Chemistry 86 (14): 7071–7078.
  268. 268 Bhattacharya, S., Sarkar, R., Nandi, S. et al. (2017). Detection of reactive oxygen species by a carbon‐dot–ascorbic acid hydrogel. Analytical Chemistry 89 (1): 830–836.
  269. 269 Kudr, J., Richtera, L., Xhaxhiu, K. et al. (2017). Carbon dots based FRET for the detection of DNA damage. Biosensors and Bioelectronics 92: 133–139.
  270. 270 Godavarthi, S., Mohan Kumar, K., Vázquez Vélez, E. et al. (2017). Nitrogen doped carbon dots derived from Sargassum fluitans as fluorophore for DNA detection. Journal of Photochemistry and Photobiology B: Biology 172: 36–41.
  271. 271 Qaddare, S.H. and Salimi, A. (2017). Amplified fluorescent sensing of DNA using luminescent carbon dots and AuNPs/GO as a sensing platform: a novel coupling of FRET and DNA hybridization for homogeneous HIV‐1 gene detection at femtomolar level. Biosensors and Bioelectronics 89: 773–780.
  272. 272 Ye, Y.‐D., Xia, L., Xu, D.‐D. et al. (2016). DNA‐stabilized silver nanoclusters and carbon nanoparticles oxide: a sensitive platform for label‐free fluorescence turn‐on detection of HIV‐DNA sequences. Biosensors and Bioelectronics 85: 837–843.
  273. 273 Liang, S.‐S., Qi, L., Zhang, R.‐L. et al. (2017). Ratiometric fluorescence biosensor based on CdTe quantum and carbon dots for double strand DNA detection. Sensors and Actuators B: Chemical 244: 585–590.
  274. 274 Fong, J.F.Y., Chin, S.F., and Ng, S.M. (2016). A unique “turn‐on” fluorescence signalling strategy for highly specific detection of ascorbic acid using carbon dots as sensing probe. Biosensors and Bioelectronics 85: 844–885.
  275. 275 Li, L., Wang, C., Liu, K. et al. (2015). Hexagonal cobalt oxyhydroxide–carbon dots hybridized surface: high sensitive fluorescence turn‐on probe for monitoring of ascorbic acid in rat brain following brain ischemia. Analytical Chemistry 87 (6): 3404–3411.
  276. 276 Kong, W., Wu, D., Li, G. et al. (2017). A facile carbon dots based fluorescent probe for ultrasensitive detection of ascorbic acid in biological fluids via non‐oxidation reduction strategy. Talanta 165: 677–684.
  277. 277 Priyadarshini, E. and Rawat, K. (2017). Au@carbon dot nanoconjugates as a dual mode enzyme‐free sensing platform for cholesterol. Journal of Materials Chemistry B 5 (27): 5425–5432.
  278. 278 Bui, T.T. and Park, S.‐Y. (2016). A carbon dot‐hemoglobin complex‐based biosensor for cholesterol detection. Green Chemistry 18 (15): 4245–4253.
  279. 279 Fang, J., Xie, Z., Wallace, G. et al. (2017). Co‐deposition of carbon dots and reduced graphene oxide nanosheets on carbon‐fiber microelectrode surface for selective detection of dopamine. Applied Surface Science 412: 131–137.
  280. 280 Shahnawaz Khan, M., Bhaisare, M.L., Pandey, S. et al. (2015). Exploring the ability of water soluble carbon dots as matrix for detecting neurological disorders using MALDI‐TOF MS. International Journal of Mass Spectrometry 393: 25–33.
  281. 281 Xiang, X., Zhang, Z., Han, L. et al. (2017). Fluorescence switching sensor for sensitive detection of sinapine using carbon quantum dots. Sensors and Actuators B: Chemical 241: 482–488.
  282. 282 Li, N., Liu, T., Liu, S.G. et al. (2017). Visible and fluorescent detection of melamine in raw milk with one‐step synthesized silver nanoparticles using carbon dots as the reductant and stabilizer. Sensors and Actuators B: Chemical 248: 597–604.
  283. 283 Madrakian, T., Maleki, S., Gilak, S. et al. (2017). Turn‐off fluorescence of amino‐functionalized carbon quantum dots as effective fluorescent probes for determination of isotretinoin. Sensors and Actuators B: Chemical 247: 428–443.
  284. 284 Gowthaman, N.S.K., Sinduja, B., Karthikeyan, R. et al. (2017). Fabrication of nitrogen‐doped carbon dots for screening the purine metabolic disorder in human fluids. Biosensors and Bioelectronics 94: 30–38.
  285. 285 Guo, W., Pi, F., Zhang, H. et al. (2017). A novel molecularly imprinted electrochemical sensor modified with carbon dots, chitosan, gold nanoparticles for the determination of patulin. Biosensors and Bioelectronics 98: 299–304.
  286. 286 Niu, J. and Gao, H. (2014). Synthesis and drug detection performance of nitrogen‐doped carbon dots. Journal of Luminescence 149: 159–162.
  287. 287 Feng, T., Chua, H.J., and Zhao, Y. (2017). Reduction‐responsive carbon dots for real‐time ratiometric monitoring of anticancer prodrug activation in living cells. ACS Biomaterials Science & Engineering 3 (8): 1535–1541.
  288. 288 Jin, H., Gui, R., Sun, J. et al. (2018). Facilely self‐assembled magnetic nanoparticles/aptamer/carbon dots nanocomposites for highly sensitive up‐conversion fluorescence turn‐on detection of tetrodotoxin. Talanta 176: 277–283.
  289. 289 An, Z., Li, Z., He, Y. et al. (2017). Ratiometric luminescence detection of hydrazine with a carbon dots‐hemicyanine nanohybrid system. RSC Advances 7 (18): 10875–10880.
  290. 290 Wu, Y., Wei, P., Pengpumkiat, S. et al. (2015). Development of a carbon dot (C‐dot)‐linked Immunosorbent assay for the detection of human α‐fetoprotein. Analytical Chemistry 87 (16): 8510–8516.
  291. 291 Wang, D., Lin, B., Cao, Y. et al. (2016). A highly selective and sensitive fluorescence detection method of glyphosate based on an immune reaction strategy of carbon dot Labeled antibody and antigen magnetic beads. Journal of Agricultural and Food Chemistry 64 (30): 6042–6050.
  292. 292 Yang, W., Huang, T., Zhao, M. et al. (2017). High peroxidase‐like activity of iron and nitrogen co‐doped carbon dots and its application in immunosorbent assay. Talanta 164: 1–6.
  293. 293 Li, N.‐L., Jia, L.‐P., Ma, R.‐N. et al. (2017). A novel sandwiched electrochemiluminescence immunosensor for the detection of carcinoembryonic antigen based on carbon quantum dots and signal amplification. Biosensors and Bioelectronics 89: 453–460.
  294. 294 Stone, S.R., Betz, A., and Hofsteenge, J. (1991). Mechanistic studies on thrombin catalysis. Biochemistry 30 (41): 9841–9548.
  295. 295 Li, N., Liu, S.G., Zhu, Y.D. et al. (2017). Tuning gold nanoparticles growth via DNA and carbon dots for nucleic acid and protein detection. Sensors and Actuators B: Chemical 251 (Supplement C): 455–461.
  296. 296 Zhang, L., Cui, P., Zhang, B. et al. (2013). Aptamer‐based turn‐on detection of thrombin in biological fluids based on efficient phosphorescence energy transfer from Mn‐doped ZnS quantum dots to carbon nanodots. Chemistry (Weinheim an der Bergstrasse, Germany) 19 (28): 9242–9250.
  297. 297 Wolberg, A.S. and Campbell, R.A. (2008). Thrombin generation, fibrin clot formation and hemostasis. Transfusion and Apheresis Science 38 (1): 15–23.
  298. 298 Huang, S., Wang, L., Huang, C. et al. (2015). A carbon dots based fluorescent probe for selective and sensitive detection of hemoglobin. Sensors and Actuators B: Chemical 221: 1215–1222.
  299. 299 Barati, A., Shamsipur, M., and Abdollahi, H. (2015). Hemoglobin detection using carbon dots as a fluorescence probe. Biosensors and Bioelectronics 71: 470–475.
  300. 300 Bowers, G.N. Jr. and McComb, R.B. (1975). Measurement of total alkaline phosphatase activity in human serum. Clinical Chemistry 21 (13): 1988–1995.
  301. 301 Hausamen, T.U., Helger, R., Rick, W. et al. (1967). Optimal conditions for the determination of serum alkaline phosphatase by a new kinetic method. Clinica Chimica Acta 15 (2): 241–245.
  302. 302 Qian, Z., Chai, L., Tang, C. et al. (2015). Carbon quantum dots‐based recyclable real‐time fluorescence assay for alkaline phosphatase with adenosine triphosphate as substrate. Analytical Chemistry 87 (5): 2966–2673.
  303. 303 Qian, Z.S., Chai, L.J., Huang, Y.Y. et al. (2015). A real‐time fluorescent assay for the detection of alkaline phosphatase activity based on carbon quantum dots. Biosensors and Bioelectronics 68: 675–680.
  304. 304 Li, G., Fu, H., Chen, X. et al. (2016). Facile and sensitive fluorescence sensing of alkaline phosphatase activity with photoluminescent carbon dots based on inner filter effect. Analytical Chemistry 88 (5): 2720–2726.
  305. 305 Tang, C., Qian, Z., Huang, Y. et al. (2016). A fluorometric assay for alkaline phosphatase activity based on β‐cyclodextrin‐modified carbon quantum dots through host‐guest recognition. Biosensors and Bioelectronics 83: 274–280.
  306. 306 Hu, Y., Geng, X., Zhang, L. et al. (2017). Nitrogen‐doped carbon dots mediated fluorescent on‐off assay for rapid and highly sensitive pyrophosphate and alkaline phosphatase detection. Scientific Reports 7 (1): 5849.
  307. 307 Qian, Z., Chai, L., Zhou, Q. et al. (2015). Reversible fluorescent nanoswitch based on carbon quantum dots nanoassembly for real‐time acid phosphatase activity monitoring. Analytical Chemistry 87 (14): 7332–7339.
  308. 308 Kong, W., Wu, D., Xia, L. et al. (2017). Carbon dots for fluorescent detection of α‐glucosidase activity using enzyme activated inner filter effect and its application to anti‐diabetic drug discovery. Analytica Chimica Acta 973: 91–999.
  309. 309 Qian, Z., Chai, L., Tang, C. et al. (2016). A fluorometric assay for acetylcholinesterase activity and inhibitor screening with carbon quantum dots. Sensors and Actuators B: Chemical 222: 879–886.
  310. 310 Li, G., Kong, W., Zhao, M. et al. (2016). A fluorescence resonance energy transfer (FRET). based “turn‐on” nanofluorescence sensor using a nitrogen‐doped carbon dot‐hexagonal cobalt oxyhydroxide nanosheet architecture and application to α‐glucosidase inhibitor screening. Biosensors and Bioelectronics 79: 728–735.
  311. 311 Lu, S., Li, G., Lv, Z. et al. (2016). Facile and ultrasensitive fluorescence sensor platform for tumor invasive biomaker β‐glucuronidase detection and inhibitor evaluation with carbon quantum dots based on inner‐filter effect. Biosensors and Bioelectronics 85: 358–362.
  312. 312 Chai, L., Zhou, J., Feng, H. et al. (2015). Functionalized carbon quantum dots with dopamine for tyrosinase activity monitoring and inhibitor screening: in vitro and intracellular investigation. ACS Applied Materials & Interfaces 7 (42): 23564–23574.
  313. 313 Yang, W., Ni, J., Luo, F. et al. (2017). Cationic carbon dots for modification‐free detection of hyaluronidase via an electrostatic‐controlled ratiometric fluorescence assay. Analytical Chemistry 89 (16): 8384–8390.
  314. 314 Sidhu, J.S., Singh, A., Garg, N. et al. (2017). Carbon dot based, naphthalimide coupled FRET pair for highly selective ratiometric detection of thioredoxin reductase and cancer screening. ACS Applied Materials & Interfaces 9 (31): 25847–24556.
  315. 315 Swietach, P., Vaughan‐Jones, R.D., Harris, A.L. et al. (2014). The chemistry, physiology and pathology of pH in cancer. Philosophical Transactions of the Royal Society B: Biological Sciences 369 (1638): 20130099.
  316. 316 Pan, D., Zhang, J., Li, Z. et al. (2010). Observation of pH‐, solvent‐, spin‐, and excitation‐dependent blue photoluminescence from carbon nanoparticles. Chemical Communications 46 (21): 3681–3683.
  317. 317 Shen, L., Zhang, L., Chen, M. et al. (2013). The production of pH‐sensitive photoluminescent carbon nanoparticles by the carbonization of polyethylenimine and their use for bioimaging. Carbon 55: 343–349.
  318. 318 Kong, B., Zhu, A., Ding, C. et al. (2012). Carbon dot‐based inorganic–organic nanosystem for two‐photon imaging and biosensing of pH variation in living cells and tissues. Advanced Materials 24 (43): 5844–5848.
  319. 319 Shi, W., Li, X., and Ma, H. (2012). A tunable ratiometric pH sensor based on carbon nanodots for the quantitative measurement of the intracellular pH of whole cells. Angewandte Chemie International Edition 124 (26): 6538–6541.
  320. 320 Nie, H., Li, M., Li, Q. et al. (2014). Carbon dots with continuously tunable full‐color emission and their application in ratiometric pH sensing. Chemistry of Matererials 26 (10): 3104–3112.
  321. 321 Karnebogen, M., Singer, D., Kallerhoff, M. et al. (1993). Microcalorimetric investigations on isolated tumorous and non‐tumorous tissue samples. Thermochim Acta 229: 147–155.
  322. 322 Gota, C., Okabe, K., Funatsu, T. et al. (2009). Hydrophilic fluorescent nanogel thermometer for intracellular thermometry. Journal of the American Chemical Society 131 (8): 2766–2767.
  323. 323 Vetrone, F., Naccache, R., Zamarrón, A. et al. (2010). Temperature sensing using fluorescent nanothermometers. ACS Nano 4 (6): 3254–3258.
  324. 324 Wang, C., Lin, H., Xu, Z. et al. (2016). Tunable carbon‐dot‐based dual‐emission fluorescent nanohybrids for ratiometric optical thermometry in living cells. ACS Applied Materials & Interfaces 8 (10): 6621–6628.
  325. 325 Wei, L., Ma, Y., Shi, X. et al. (2017). Living cell intracellular temperature imaging with biocompatible dye‐conjugated carbon dots. Journal of Materials Chemistry B 5 (18): 3383–3390.
  326. 326 Wang, C., Jiang, K., Wu, Q. et al. (2016). Green synthesis of red‐emitting carbon nanodots as a novel “turn‐on” nanothermometer in living cells. Chemistry – A European Journal 22 (41): 14475–14479.
  327. 327 Kalytchuk, S., Poláková, K., Wang, Y. et al. (2017). Carbon dot nanothermometry: intracellular photoluminescence lifetime thermal sensing. ACS Nano 11 (2): 1432–1442.
  328. 328 Eftekhari, E., Wang, W., Li, X. et al. (2017). Picomolar reversible Hg(II). solid‐state sensor based on carbon dots in double heterostructure colloidal photonic crystals. Sensors and Actuators B: Chemical 240: 204–211.
  329. 329 Cunningham, B.T. and Zangar, R.C. (2012). Photonic crystal enhanced fluorescence for early breast cancer biomarker detection. Journal of Biophotonics 5 (8–9): 617–628.
  330. 330 Martinez, A.W., Phillips, S.T., Butte, M.J. et al. (2007). Patterned paper as a platform for inexpensive, low‐volume, portable bioassays. Angewandte Chemie International Edition 46 (8): 1318–1320.
  331. 331 Martinez, A.W., Phillips, S.T., Whitesides, G.M. et al. (2010). Diagnostics for the developing world: microfluidic paper‐based analytical devices. Analytical Chemistry 82 (1): 3–10.
  332. 332 Carrilho, E., Martinez, A.W., and Whitesides, G.M. (2009). Understanding wax printing: a simple micropatterning process for paper‐based microfluidics. Analytical Chemistry 81 (16): 7091–7095.
  333. 333 Zhang, M., Liu, H., Chen, L. et al. (2013). A disposable electrochemiluminescence device for ultrasensitive monitoring of K562 leukemia cells based on aptamers and ZnO@carbon quantum dots. Biosensors and Bioelectronics 49: 79–85.
  334. 334 Wang, Y., Wang, S., Ge, S. et al. (2013). Facile and sensitive paper‐based chemiluminescence DNA biosensor using carbon dots dotted nanoporous gold signal amplification label. Analytical Methods 5 (5): 1328–1336.
  335. 335 Yan, F., Zou, Y., Wang, M. et al. (2014). Highly photoluminescent carbon dots‐based fluorescent chemosensors for sensitive and selective detection of mercury ions and application of imaging in living cells. Sensors and Actuators B: Chemical 192: 488–495.
  336. 336 Zhang, H., Chen, Y., Liang, M. et al. (2014). Solid‐phase synthesis of highly fluorescent nitrogen‐doped carbon dots for sensitive and selective probing ferric ions in living cells. Analytical Chemistry 86 (19): 9846–9852.
  337. 337 Sun, X., Yang, S., Guo, M. et al. (2017). Reversible fluorescence probe based on N‐doped carbon dots for the determination of mercury ion and glutathione in waters and living cells. Analytical Sciences 33 (7): 761–767.
  338. 338 Jiang, K., Sun, S., Zhang, L. et al. (2015). Red, green, and blue luminescence by carbon dots: full‐color emission tuning and multicolor cellular imaging. Angewandte Chemie International Edition 54 (18): 5360–5363.
  339. 339 Zhang, L., Wang, D., Huang, H. et al. (2016). Preparation of gold–carbon dots and ratiometric fluorescence cellular imaging. ACS Applied Materials & Interfaces 8 (10): 6646–6655.
  340. 340 Kang, M.S., Singh, R.K., Kim, T.‐H. et al. (2017). Optical imaging and anticancer chemotherapy through carbon dot created hollow mesoporous silica nanoparticles. Acta Biomaterialia 55: 466–480.
  341. 341 Zhao, Q., Wang, S., Yang, Y. et al. (2017). Hyaluronic acid and carbon dots‐gated hollow mesoporous silica for redox and enzyme‐triggered targeted drug delivery and bioimaging. Materials Science and Engineering: C 78: 475–484.
  342. 342 Wang, Y., Cui, Y., Zhao, Y. et al. (2017). Fluorescent carbon dot‐gated multifunctional mesoporous silica nanocarriers for redox/enzyme dual‐responsive targeted and controlled drug delivery and real‐time bioimaging. European Journal of Pharmaceutics and Biopharmaceutics 117: 105–115.
  343. 343 Pan, L., He, Q., Liu, J. et al. (2012). Nuclear‐targeted drug delivery of TAT peptide‐conjugated monodisperse mesoporous silica nanoparticles. Journal of the American Chemical Society 134 (13): 5722–5725.
  344. 344 Creixell, M. and Peppas, N.A. (2012). Co‐delivery of siRNA and therapeutic agents using nanocarriers to overcome cancer resistance. Nano Today 7 (4): 367–379.
  345. 345 Kamkaew, A., Lim, S.H., Lee, H.B. et al. (2013). BODIPY dyes in photodynamic therapy. Chemical Society Reviews 42 (1): 77–88.
  346. 346 Zou, L., Wang, H., He, B. et al. (2016). Current approaches of photothermal therapy in treating cancer metastasis with nanotherapeutics. Theranostics 6 (6): 762–772.
  347. 347 Pandey, S., Gedda, G.R., Thakur, M. et al. (2017). Theranostic carbon dots ‘clathrate‐like’ nanostructures for targeted photo‐chemotherapy and bioimaging of cancer. Journal of Industrial and Engineering Chemistry 56: 62–73.
  348. 348 Zhang, M., Wang, W., Zhou, N. et al. (2017). Near‐infrared light triggered photo‐therapy, in combination with chemotherapy using magnetofluorescent carbon quantum dots for effective cancer treating. Carbon 118: 752–764.
  349. 349 Nandi, S., Bhunia, S.K., Zeiri, L. et al. (2017). Bifunctional carbon‐dot‐WS2 nanorods for photothermal therapy and cell imaging. Chemistry – A European Journal 23 (4): 963–969.
  350. 350 Beack, S., Kong, W.H., Jung, H.S. et al. (2015). Photodynamic therapy of melanoma skin cancer using carbon dot – chlorin e6 – hyaluronate conjugate. Acta Biomaterialia 26: 295–305.
..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset