Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

A comprehensive review on the capability of graphene quantum dots-based/involved platforms for the detection of inorganic ions

Prashant Dubey
Centre of Material Sciences, Institute of Interdisciplinary Studies (IIDS), University of Allahabad, Prayagraj 211002, Uttar Pradesh, India. E-mail: pdubey@allduniv.ac.in; pdubey.au@gmail.com

Received 10th July 2025 , Accepted 5th November 2025

First published on 24th November 2025


Abstract

Substantial contamination in the ecosystem (particularly, in waterbodies) due to the disposal of hazardous substances from superfluous industrial byproducts and other human activities is one of the most serious environmental issues. Although water is among the five basic elements required for all the living beings, it is continuously becoming unsafe and unhygienic for the purpose of drinking and household activities. The presence of heavy metal ions and many toxic anions (beyond their permissible concentration) significantly contributes to water pollution. Therefore, it is judicious to detect inorganic ions in order to avoid adverse situations related to the human health and ecological imbalance. Owing to the easy affordability and unique properties of zero-dimensional graphene quantum dots (GQDs), particularly, high/tunable fluorescence, electronic conductivity, electrocatalytic activity, chemiluminescence (CL) and electrochemiluminescence (ECL), GQDs-based/involved platforms are potentially deployed for the detection of inorganic ions with reasonable selectivity and sensitivity. The versatility of these sensing probes includes the possibility of detection through fluorimetric, colorimetric, electrochemical, CL, and ECL techniques. This review aims to comprehensively inspect the deployment of GQDs-based/involved sensors for the recognition of inorganic ions, highlighting different sensing approaches with the development of performance metrics and providing insights into their underlying mechanisms. Furthermore, this collective outlook on GQDs-based/involved sensors for inorganic ions may help to identify shortcomings in the existing knowledge and influence/inspire new research with better prospects.


image file: d5ra04935k-p1.tif

Prashant Dubey

Prashant Dubey obtained his PhD from the Indian Institute of Technology, Kanpur, India (2006; Department of Chemistry) under the supervision of Prof. S. Sarkar. He has also worked as a Postdoctoral Researcher with Prof. C. J. Lee, Korea University, South Korea. Currently, he is engaged in his role as an Associate Professor in the Centre of Material Sciences, University of Allahabad, Prayagraj, India. He has received various fellowships (CSIR, DST, SSL, INSA, and IASc) and an opportunity to work with Prof. C. N. R. Rao (Bharat Ratna) in his ongoing academic journey. His research interest includes the exploration of 0D/1D/2D nanomaterials for energy, environmental, and biological applications.


1. Introduction

Industrialization has helped to meet the demands of modern civilization; however, it has contributed to the release of a plethora of toxic chemicals and byproducts into the environment, especially in aquatic ecosystems and soil. Heavy metals/metal ions (HMs/HMIs) are some of the toxic substances, which are serious environmental pollutants that cause health hazards to living beings.1–4 The chemical production industry, mining & metallurgical industry, and electroplating engineering have significantly contributed to the discharge of various HMs/HMIs into the environment.5 These HMs/HMIs enter in the living organisms predominantly by the means of oral route, apart from skin contact and respiration.6 Among the metallic contaminants in the natural environment, toxic HMIs such as mercuric ion (Hg2+), lead ion (Pb2+), cadmium ion (Cd2+), arsenic ion (As3+) and chromium ion (Cr6+/Cr3+) have been regarded as the foremost detrimental substances to the living beings and human health.1–4 For instance, inorganic and organic mercury (Hg) compounds exhibit neuro-/geno-/immuno- and cardio-toxicity as well as cause damage to the reproductive system, promote cancer and diseases related to pregnancy, and induce damages to pulmonary and renal organs.7 Human exposure to lead (Pb) is responsible for damage to the kidneys, brain, reproductive organs, bones, and neurological system.8 The toxic effect of Cd2+ towards extensive damages at the tissue and cellular levels is induced by oxidative stress, calcium ion (Ca2+)-signal disruption, and cellular-signal interference.9 Chronic exposure to trivalent As3+ induces toxic effects in the liver, kidneys, reproductive organs, and cardiovascular system, which further manifest to cause cancer.10 Aside from the carcinogenic nature of Cr6+, it has been recognized as a budding neuro-toxicant owing to its deleterious effect on the human brain.11 Even though sodium ion (Na+), potassium ion (K+), Ca2+, magnesium ion (Mg2+), cupric ion (Cu2+), cobalt ion (Co2+), zinc ion (Zn2+), and ferrous/ferric ions (Fe2+/Fe3+)12,13 metallic cations and phosphate (PO43−),14 pyrophosphate (P2O74−),15 sulfide (S2−),16 nitrite (NO2),17 iodide (I),18 and thiocyanate (SCN)19 anions are essential for biological activities, they can be a threat beyond their permissible limit. There are proper guidelines set by the World Health Organization (WHO) and Environmental Protection Agency (EPA) for the permissible limits of many of the inorganic contaminants in drinking water (Table 1).20,21
Table 1 Various HMs and inorganic anions with their permissible limits according to the WHO and EPA20,21
HMs/inorganic anions Permissible limit in µg L−1
WHO EPA
Ag 100
Al 900
As 10 10
Cd 3 5
Cr (total) 50 100
Cu 2000 1300
CN 2000 200
F 1500 4000
Hg 6 2
Ni 70
NO2 3000 1000
Pb 10 10
U 30 30


Therefore, the assurance of the level of these contaminants in water by proactive and accurate monitoring is crucial to avoid escalating situations and further health risks. Various spectroscopic techniques such as atomic absorption spectroscopy (AAS),22,23 inductively coupled plasma-optical emission spectroscopy (ICP-OES),24,25 inductively coupled plasma-mass spectroscopy (ICP-MS),26,27 X-ray fluorescence spectroscopy,28,29 and surface-enhanced Raman scattering spectroscopy30,31 have been utilized for the reliable and efficient detection of inorganic ions, particularly HMs. However, these analytical methods are often limited due to the requirement of expensive instrumentation, lengthy processes, complicated sampling protocols, and trained personnel. In contrast, fluorimetric (FL), colorimetric (COL), electrochemical (EC), chemiluminescence (CL), and electrochemiluminescence (ECL) techniques offer the promising and reliable detection of inorganic ions due to their simple operation, short time, low cost, user-friendliness, good precision, and in situ analytical capability. However, although semiconductor quantum dots,32,33 metallic nanoparticles (NPs),34–36 metal oxides,37–39 and organic molecules40–42 have been employed as functional materials to sense inorganic ions, they possess inherent toxicity, structural instability, and environmental concerns. Hence, the development of sensing platforms with environmentally benign, cost-effective, non-toxic, and easily accessible functional materials is of extreme significance and has become a major research trend in the past few years.

Graphene quantum dots (GQDs) can be referred to as highly crystalline zero-dimensional carbon nanostructures with predominant sp2 hybridized carbon frameworks consisting of one or few graphene layers, graphitic in-plane lattice spacing of 0.18 to 0.24 nm, graphitic interlayer spacing of 0.334 nm, and lateral sizes commonly below 10 nm.43 Since GQDs were first fabricated in 2008,44 different top-down and bottom-up synthetic pathways have been actively explored for the easy and cost-effective production of GQDs (discussed in Section 2). The obtained GQDs usually contain various functional groups at their edges or on the defect sites of their basal planes, which along with their surface/edge states and/or size-based quantum confinement effect generate a strong and stable photoluminescence (PL).45,46 Tailoring the physiochemical and photo-physical properties of GQDs via surface-functionalization, heteroatom-doping, and structural defects have shown immense possibility to tune the electronic structures, optical properties, and chemical reactivity in modified-GQDs.47–49 Additionally, GQDs/modified-GQDs are chemically stable,50 possess a high quantum yield (QY),51 act as electron transporters,52 dispersible/soluble in aqueous medium,53 low-toxic,54 biodegradable, and biologically compatible.55 Consequently, GQDs/modified-GQDs have been shown to be suitable for a diverse range of applications including chemo-sensing,56 biosensing and bioimaging,57 energy storage-conversion,58 photodynamic and photo-thermal therapies,59 drug delivery,60 electrocatalysis,61 and light-emitting diodes.62 The promise of GQDs/modified-GQDs in the selective and sensitive detection of almost all types of metal ions (MIs) and various anions via FL, COL, EC, CL, and ECL methods is one of the highly explored areas of research, which enables the identification of environmental contaminants in aqueous medium as well as in biological samples and living cells.

In the literature, the HMI sensing applications of GQDs-based systems have been reviewed by some researchers, whereas the recognition of inorganic anions is often neglected or briefly addressed. For instance, Zhou et al. (2016)63 reviewed the GQDs synthesized via various top-down and bottom-up routes for the fluorescence-based detection of inorganic ions, organic molecules, and biomolecules. The optical detection of HMIs using graphene, graphene oxide (GO), GQDs, and doped-GQDs was summarized by Zhang et al. (2018).64 Li et al. (2019)65 broadly summarized the sensing applications of GQD- and carbon quantum dot (CQD)-based nanomaterials via FL, CL, ECL, and EC methods with specific examples of HMIs. The optical sensing applications of GQDs-based materials for toxic HMIs were further reviewed by Anas et al. in 2019.66 Li et al. (2021)48 emphasized the fluorescence-based detection of HMIs, along with the other analytes using doped-GQDs. Revesz et al. (2022)67 highlighted GQDs-based sensors for the detection of harmful contaminants such as HMIs, along with alkali and alkaline-earth MIs and discussed the various mechanisms involved in the sensing operation. In a recent review article, Wu et al. (2025)68 emphasized the property regulation of GQDs by heteroatom-doping and surface/edge modification, and furthermore their impact on the turn-off and turn-on based fluorescence sensing of various analytes including specific examples of HMIs and anions. Another recent review by Saisree S. et al. (2025)69 was dedicated to GQDs-based materials for the EC sensing of toxic HMIs, particularly, Cd2+, Pb2+, and Hg2+ along with the mechanistic details in the detection process.

Notably, single-/dual-heteroatom doped-GQDs including nitrogen-doped GQDs (N-GQDs), sulfur-doped GQDs (S-GQDs), boron-doped GQDs (B-GQDs), nitrogen/sulfur-co-doped GQDs (N,S-GQDs), boron/nitrogen-co-doped GQDs (B,N-GQDs), boron/phosphorus co-doped GQDs (B,P-GQDs), and nitrogen/phosphorus-co-doped GQDs (N,P-GQDs) have been encountered for the fabrication of GQDs-based sensing systems, and furthermore their application to detect inorganic ions at various levels of selectivity/sensitivity. Additionally, the introduction of specific functional groups in the GQDs moiety has shown their relevance to selectively interact with particular inorganic ions, resulting in considerable sensitivity in the detection operation. Moreover, GQDs/modified-GQDs also serve as key components, along with the other functional segments to build GQDs involved sensory architectures for the probing of inorganic ions. Obviously, the detection of inorganic ions (particularly, toxic species in ionic form and biocompatible ions) in aqueous medium and living bodies through GQDs-based/involved simple-effective sensing probes (by utilizing various sensing methods) is an effective strategy. Owing to the continuous progress of this research field, the rationality of the current review is to provide a comprehensive, in-depth, and systematic overview of the GQDs-based/involved platforms utilized in the identification of target analytes, particularly HMIs, biologically important alkali and alkaline-earth MIs, rare-earth MIs, radioactive MIs, and inorganic anions. A brief summary at the end of the discussion for each inorganic ion will provide an understanding, comparison, and identification of suitable nanoprobe/sensing methodologies. Before approaching the sensing aptitudes of GQDs-based/involved systems, this review provides a thorough discussion of the various top-down and bottom-up routes for the synthesis of GQDs/doped-GQDs, their functionalization (covalent and non-covalent) strategies, and relevant properties such as physiochemical, optical, CL, ECL, and EC characteristics. Finally, a collective summary in the form of conclusion and challenges/future prospects of GQDs-based/involved systems is presented in terms of their synthesis, property modulation, and loopholes/improvement of their sensing metrics. We believe that this review article will provide comprehensive information about GQDs-based/involved inorganic ion sensors to identify the research progress in one platform. Moreover, a balanced discussion about the advantages/disadvantages may result in a critical judgement on the capability of GQDs-based/involved systems in the analytical detection of inorganic ions. Consequently, the discussion may be complemented with the additional knowledge and impact from new ideas among scientists engaged in the area of inorganic ion analysis in environmental bodies and living systems.

2. Synthesis of GQDs/doped-GQDs

Miniaturization of the large-sized graphitic carbon material through successive cutting via chemical/physical means (top-down) or fusion of small organic molecules through condensation/carbonization (bottom-up) may result in the production of GQDs/doped-GQDs. Although GQDs or their doped-counterparts can be effectively as well as controllably synthesized using both approaches, their physical characteristics such as crystallinity, size distribution, surface functionality, colloidal stability, presence of hetero-elements, and emission characteristics entirely depend on the starting precursors and synthetic conditions. The first approach for the fabrication of GQDs is based on the top-down strategy, in which graphene crystallites are cleaved into the desired geometry even up to 10 nm small GQDs through the oxygen plasma etching process.44

2.1. Top-down approach

A large variety of readily available bulk carbonaceous precursors such as carbon black,70 graphite,71 carbon nano-onions (CNOs),72 graphene sheets (GSs),73 fullerene,74 carbon nanotubes (CNTs),75 carbon shoot,76 carbon fibers,77 GO,78 pyrolyzed biomass,79 and coal80 can undergo dissociation by means of chemical, hydrothermal (HT), solvothermal (ST), microwave (MW), electrolysis, ultrasonication or physical treatment to produce GQDs/doped-GQDs. Oxidative cutting of the bulk precursors using strong acid oxidants such as H2SO4/HNO3,71,72,77 HNO3,81,82 H3PO4,73 and HClO4 (ref. 83) at elevated temperature is the mainstream top-down condition for the synthesis of GQDs, affording various functional groups (–OH, –COOH, and epoxides). For example, oxidative cutting of carbon fiber in concentrated H2SO4/HNO3 (3[thin space (1/6-em)]:[thin space (1/6-em)]1 volume ratio) under ultrasonication (1 h), followed by heating (85 °C, 24 h) produced GQDs with an average size of 2.45 nm.77 Oxidative cleavage of brewery spent grain (BSG)-derived reduced GO (synthesized via the calcination of BSG and ferrocene at 300 °C, 45 min) in the presence of H2SO4/KMnO4 followed by ultrasonication (100 W, 1 h) resulted in GQDs with a size distribution of 10–35 nm.79 However, although the oxidative treatment route is widely implemented in the synthesis of GQDs, it is difficult to upscale and is also environmental unfriendly due to the production of harmful gases and inorganic salts during the whole process, along with residual corrosive acid and/or other by-products.

Therefore, the acid-free oxidative cleavage of carbonaceous precursors with Fenton reagent (H2O2/Fe3+),84 H2O2,80 alkaline H2O2,85 oxone,74 KMnO4,86 KO2,87 NaClO,88 etc. has been successfully applied to synthesize GQDs/doped-GQDs. For example, Lyu et al.84 reported the gram-scale synthesis of oxidized-GQDs (60% product yield) by treating GO powder with H2O2/FeCl3 in an autoclave at 180 °C (8 h). The dissociation of H2O2 (induced by Fe3+/Fe2+ catalysis) produced hydroxyl (˙OH) radicals, which attacked the defective carbons of GO to break into well-crystalline GQDs (average size: ∼3.7 nm). Pre-treated coal (600 °C, 1 h, argon (Ar)) was refluxed with H2O2 (80 °C, 10 h) followed by HT treatment in the presence of HF (120 °C, 12 h) to obtain dual-passivated fluorine (F)/nitrogen co-doped GQDs.80 Oxone-assisted opening of C60 molecules under ST conditions enabled the acid-free synthesis of blue-emitting GQDs (QY: 23.5%), which were further modified with 2,3-diaminonaphthalene (DAN), resulting in the formation of orange-emissive DAN–GQDs with a QY as high as 52.4%.74 The one-step acid-free oxidative cutting of GO sheets with KMnO4 under ultrasonication and MW (400 W, 90 °C, 30 min) irradiation afforded GQDs (average size/QY: 2 nm/23.8%) with a product yield up to 81%.86 Besides oxidative cutting, the amine,89 amine-hydrazine,90 hydrazine,91 NH3,76 dimethylformamide (DMF),92 N-methyl-2-pyrrolidone (NMP),93 etc. driven reduction/reductive cutting of oxidized-carbon/bulk carbon also yielded doped- or undoped-GQDs. For example, ST treatment of graphite in NMP solvent (300 °C, 24 h) facilitated simultaneous exfoliation and scission operation, resulting in 1–2 layered N-GQDs.93 Alternatively, an HT-treated GO dispersion (180 °C, 24 h) was simply tip-sonicated (100 W, 1 h), resulting in GQDs with an average size as small as ∼1.53 nm. Notably, the average size of GQDs was significantly reduced after tip sonication (average size before sonication was ∼15.7 nm). Moreover, the property of GQDs was further engineered with the inclusion of extra defects by Ar-plasma treatment.78

EC synthesis of GQDs is another top-down approach that holds a promise to control the degree of oxidation/cleavage of the carbon precursor by applying an electric potential under ambient conditions both in non-aqueous94,95 or aqueous96,97 electrolytes without involving toxic oxidizing/reducing agents. In this method, the bulk precursor is applied either as the working electrode or dispersed in a solvent. For example, the EC exfoliation of carbon fibers (anode) in an ionic liquid (IL, 1-butyl-3-methylimidazolium tetrafluoroborate, BMIMBF4) electrolyte resulted in the formation of blue-emitting GQDs. Electrical stress under a high applied voltage (6 V) favoured the intercalation of BF4 within the layers/edge sites of the carbon fibers to induce corrosion, and eventually the formation of IL-functionalized GQDs. Moreover, by adding 15/30% H2O in IL electrolyte, the obtained GQDs showed green-/yellow-emission.94 Qiang et al.97 demonstrated a facile electrochemical trimming to fragment a large GO nanosheet dispersion into graphene nanoribbons, graphene nanosheets (GNSs), and GQDs just by tuning the reaction time to 2, 3, and 5 h, respectively. Here, ˙OH and oxygen (˙O2) radicals from the high voltage electrolysis of water (30 V) get intercalated-adsorbed on the fragile portions of GO sheets, accompanied by the disintegration of sheets into smaller fragments.

The application of laser or pulsed laser has also demonstrated to etch bigger-size carbon materials into small size GQDs within a short duration at room temperature.98,99 This method follows a one-step environmentally benign process by avoiding harmful chemicals and tedious post-purification protocols. Kang et al.99 applied pulsed laser exposure (Q-switch Nd:YAG, λ = 355 nm, 30 min, room temperature, air) to graphite flakes (dispersed in ethanol/diethylenetriamine (DETA)), which fragmented into N-GQDs. Interestingly, the N-GQDs synthesized without sonication had a much better QY (9.1%) in comparison to sonication-assisted laser irradiation (QY: 4.2%), which is attributed to the effective incorporation of nitrogen element. Plasma-plume induced by the laser irradiation of cavitation bubbles thermally decomposed the starting materials (in the form of carbon clusters/nitrogen molecules), which evaporated-condensed to produce N-doped GNS aggregates, and further fragmentation by pulsed laser into small N-GQDs.

Purely mechanical tailoring of bulk pristine materials into GQDs is also a neat and clean synthetic approach.100–102 For example, 44.6 wt% product yield of GQDs from multi-walled CNTs (MWCNTs) was achieved by combining silica-assisted ball milling and sonication-based exfoliation, centrifugation, and filtration, which is inspiring.101

2.2. Bottom-up approach

Synthesis of GQDs/doped-GQDs via the bottom-up approach involves pyrolysis,103 HT/ST treatment,104,105 MW-assisted carbonization,106 MW-assisted HT (MW-HT) treatment,107 electron-beam irradiation,108 ultraviolet (UV) irradiation,109 solution-phase condensation,110 or direct current (DC) microplasma treatment111 of small precursor molecules. Among them, pyrolysis is one of the straightforward synthetic strategies, in which citric acid (CA),112 trisodium citrate (TSC),113 glucose,114 L-glutamic acid (GA),115 GA/aspartic acid,116 etc. get thermally decomposed-carbonized at high temperature (above their melting point), resulting in the formation of GQDs or doped-GQDs. For example, solid CA was heated at 200 °C for 30 min under stirring to transform into an orange liquid, which was subsequently dissolved in NaOH solution, followed by pH adjustment to 8.0 and centrifugation to isolate GQDs with an average diameter of ∼2.2 nm.112 Doped-GQDs were also synthesized through the pyrolysis method by adding other compounds such as urea,117 urea-ammonia sulphate,118 glutathione (GSH),119 and thiourea (TU)120 in CA. For example, N-GQDs and N,S-GQDs were synthesized via the infrared (IR)-assisted pyrolysis of CA-urea and CA-urea-ammonia sulphate, respectively, at 260 °C for 10 min. Here, induction-based transfer of heat energy to the precursor via electromagnetic radiation is beneficial to achieve a homogeneous and rapid heating process at a ramping rate of 30 °C min−1.118

The HT and ST synthesis of GQDs or their doped-counterparts using appropriate starting molecular precursors in water and organic solvents, respectively, and subsequent heating under inherent vapour pressure are other facile synthetic approaches, which have been widely employed in the literature. For instance, xylan dissolved in an aqueous solution of NaOH/urea was treated under HT condition (240 °C, 24 h) to obtain single-layered (sl-) N-GQDs (sl-N-GQDs) with an average size/QY of 3.2 nm/23.8%. Attachment of NaOH hydrates to the xylan chain via hydrogen-bonding followed by wrapping with urea hydrates led to the growth of a channel inclusion compound. Furthermore, hydrolysis-carbonization of xylan, exfoliation with the assistance of NH3 and CO2 (generated from the decomposition of urea), and incorporation of nitrogen-containing radicals during the HT process eventually generated sl-N-GQDs.121 The one-step ST treatment of gallic acid in absolute ethanol at 160 °C (6 h) produced green-fluorescent GQDs with a mean diameter of 10.1 nm.122

The MW-enabled carbonization of organic precursor/biomass extract provides a straightforward, quick, and homogeneous heating process to achieve GQDs/doped-GQDs in a few minutes. Hsieh et al.123 developed a solid-phase MW-assisted synthetic route (2.45 GHz, 720 W, ≤180 °C, 15 min) to synthesize N-GQDs and B,N-GQDs with a product yield up to 45.1 wt% using CA as the carbon source and urea/boric acid (H3BO3) as the nitrogen/boron source. The in situ nitrogen-doping is distributed in the form of pyrrolic/pyridinic/graphitic nitrogen and amide functional groups, while B4C/BCO2 bonding types (within GQDs structure or decorated on the skeleton) are assigned to the boron-configuration. An aqueous solution of Mangifera indica leaf extract was heated in an MW oven (10 min) to yield red-emissive GQDs under UV irradiation with a QY of 45%.124

The bottom-up approach via the MW-HT technique by combining MW and HT features is advantageous for the rapid, energy saving, uniform, and efficient preparation of GQDs. Contrary to MW synthesis, where the precursor is irradiated by MW under atmospheric pressure, the MW-HT method relies on the MW-based heating of the starting material in a MW-transparent sealed vessel. As a result, a high temperature is achieved in a short period of time due to the creation of a pressurized environment.125 For example, dielectric heating of a 1,3,6-trinitropyrene (TNP)/0.3 M NaOH solution in a confined glass vessel for 3 min under MW irradiation (reaction temperature: 200 °C) resulted in bright yellow-luminescent GQDs (under UV light).107

The microplasma technique has been effectively used under ambient conditions without involving harsh chemicals or reaction environment to synthesize GQDs/doped-GQDs. One-dimensional gaseous discharge within the small depth of the plasma–liquid interface can produce reactive species (radicals, ions, electrons, etc.) with a high energy density for the nucleation and growth of GQDs from the starting precursor.111 The plasma electrochemical synthesis of N-GQDs was demonstrated under ambient conditions using chitosan as the sole precursor. The reaction was performed under DC discharge flowing Ar (discharge current: 5 mA) for 1 h to accumulate N-GQDs (average size: 3.9 nm) at few mm below the plasma–liquid interface. Based on the experimental observations, it was deduced that plasma-generated ˙OH initially cleaves the glycosidic bonds of the long-chain chitosan to generate aldehyde- and carboxylic-containing species, which subsequently reassembled into an aromatic structure to grow N-GQDs with the involvement of solvated electrons (generated by plasma).126

Solution-phase chemistry can allow step-wise chemical reactions for the synthesis of GQDs from small organic molecules. For instance, D-glucose is catalytically (acetic acid as the catalyst) transformed into the Amadori product in the presence of hexadecylamine (HDA), followed by spontaneous dehydrolysis in the solution-phase, resulting in the formation of single-crystalline hexagonal-shaped GQDs with low oxygen defects.110 Ochi et al.127 demonstrated the mass-scale synthesis of blue-green fluorescent GQDs (average size: ∼1.4 nm) in an open atmosphere by air-flow reflux heating of phloroglucinol-Na3PO4·12H2O in 1,2-pentanediol (180 °C, 6 h), followed by dialysis and silica-gel chromatographic purification. Here, Na3PO4·12H2O acted as the base catalyst to promote the dehydration–condensation reaction during the synthesis process, resulting in an exceptionally high product yield of 99.4%. Moreover, after silica-gel purification, the QY of GQDs (in ethanol) increased from 54% to 75%. Experimental results revealed that the attachment of 1,2-pentanediol at the edges of GQDs effectively suppressed their aggregation and concentration-induced quenching, and therefore a high QY was achieved.

3. Functionalization of GQDs/doped-GQDs

Various functional moieties can be introduced in GQDs during their synthesis process. For example, HT treatment of CA and polyethyleneimine (PEI) resulted in PEI-functionalized N-GQDs.128 When CA is HT-treated in the presence of taurine, sulfonic acid group-functionalized GQDs are obtained with high water solubility (3.6 mg mL−1).129 Furthermore, GQDs/doped-GQDs synthesized via the top-down or bottom up approach inherently contain aromatic domains and a range of oxygen-containing functional groups on their surface/edges, which open the possibility to carry out post-modification through various covalent and non-covalent chemistries. The condensation reaction between –COOH and –NH2 groups in the presence of 1-ethyl-3-(3-dimethylaminopropyl)-carbodiimide (EDC)/N-hydroxysuccinimide (NHS) system or only with EDC through carbodiimide coupling chemistry can produce an amide linkage between two building blocks and has been extensively utilized in the covalent-functionalization of GQDs. For instance, GQDs synthesized from graphite flakes were coupled with 2,6-diaminopyridine (DAP) via EDC/NHS-based coupling reaction, which showed bluish-green emission and a higher QY (13.4%) in comparison to bare GQDs (4.7%).130 Recently, GQDs were activated with EDC/N-hydroxysulfosuccinimide sodium salt (sulfo-NHS) and coupled with dopamine (DA) to produce DA-functionalized GQDs (DA–GQDs).131 Amidation reaction can also be conducted under basic conditions without the involvement of EDC/NHS to functionalize GQDs. For example, the formation of imine bonds after the reaction between –NH2 groups of N-GQDs and –COOH groups of pamoic acid (PA) under alkaline conditions confirmed the covalent attachment of PA to the surface of N-GQDs.132

The acid chloride formation route relies on the transformation of –COOH groups into highly reactive acid chloride, which can react with amine/alcohol group-containing moieties to generate amide/ester bonding. The activation of the –COOH groups present on the GQDs in the form of acyl chloride, followed by N-(rhodamine B) lactam-ethylenediamine (RBD) substitution through the amide linkage in the presence of triethylamine (TEA) resulted in the formation of RBD–GQDs.133 Conversely, hydroxyl-functionalized 3,4-ethylenedioxythiophene (EDOT) was condensed with acid chlorides at the edge of GQDs (generated by the transformation of –COOH with oxalyl chloride) via ester linkage in the presence of TEA/4-dimethylaminopyridine (DMAP), resulting in EDOT–GQDs.134

GQDs can also be functionalized through esterification reaction between the –COOH and –OH groups present on the starting reactants. For instance, the N,N′-dicyclohexylcarbodiimide (DCC)/DMAP-induced coupling reaction between the –OH groups of GQDs and –COOH groups of the reversible addition–fragmentation chain transfer agent (RAFT) resulted in RAFT–GQDs, which were further integrated with a block copolymer (BCP) to produce multicolor emitting BCP–GQDs.135 Conversely, the –COOH groups of GQDs are activated with DCC/DMAP, followed by esterification reaction with the –OH moiety of dimercaprol (DMC) to synthesize DMC–GQDs.103 Consecutive Steglich-esterification followed by reductive-esterification condition occurred between pristine GQDs and the 4,4′-(1,2-diphenylethene-1,2-diyl)diphenol (TPE-DOH) rotor molecule to synthesize edge-functionalized TPE–GQDs. The presence of ester (–C(=O)OC–) and ether (–C–O–C–) groups in TPE–GQDs indicated the successful substitution of –COOH/C[double bond, length as m-dash]O groups at the edge of GQDs via esterification/reductive–esterification reaction. Moreover, the existence of four phenyl groups in TPE–GQDs effectively maximize the steric hindrance to inhibit aggregation-induced quenching (AIQ) and result in aggregation-induced emission (AIE) characteristics with a QY as high as 16.8% in the solid state.136

Some other covalent-functionalization strategies for the modification of GQDs are as follows: polyvinyl alcohol (PVA) is grafted on the GQDs surface via Friedel–Crafts alkylation reaction.137 Propargyl bromide is reacted with the –OH groups of GQDs to introduce C[triple bond, length as m-dash]C triple bonds at the periphery of GQDs, which subsequently cross-linked with azide-functionalized poly(ethylene oxide) via Cu+-catalyzed click chemistry to result in poly(ethylene glycol) (PEG)-functionalized GQDs (PEG–GQDs).138 Recently, the click reaction between the thiol (–SH) groups of cysteine (Cys) and C[double bond, length as m-dash]C double bonds of GQDs in the presence of azobisisobutyronitrile initiator afforded Cys–GQDs.139

Secondary weak interactions such as π–π, ionic, hydrogen-bonding, and van der Waals interactions may offer a simple and rapid protocol for the non-covalent-functionalization of GQDs. Due to the inherent aromatic nature of GQDs, they can conjugate with suitable counterparts via π–π stacking. For instance, π–π stacking between GQDs and peptide (PEP)-functionalized AuNPs resulted in the formation of an AuNPs–PEP@GQDs nanoconjugate.140 π–π interaction between OH-GQDs and PPy-Br dye resulted in the formation of a ratiometric conjugate for sensing application.141 The negative surface charge of GQDs arising from –COOH/–OH groups can allow them to interact with positively charged moieties through ionic or electrostatic interaction. For example, negatively charged GQDs are passivated with xylan and chitosan oligosaccharide via electrostatic interactions for sensing and bioimaging applications.142

Various functional groups on GQDs can also assist hydrogen-bonding interaction during post-modification or incorporation in other matrices. For instance, the strong hydrogen-bonding ability of GQDs with the –OH groups of cellulose resulted in the formation of a stable GQDs/cellulose membrane with open structure and high water permeability.143 Electro-polymerization of aniline in the presence of N-GQDs generated an N-GQDs/polyaniline (N-GQDs/PANI) nanocomposite for the non-invasive detection of glucose, where N-GQDs are electrostatically (preferably hydrogen-bonding) bonded with polymer chains.144

van der Waals interaction-based non-covalent functionalization of GQDs is less common in the literature. A short range van der Waals interaction between amine-functionalized GQDs (Am–GQDs) and few-layer MoS2 sheets by the simple mixing of two components may be a representative example, where Am-GQD/MoS2 heterostructures were probed for Foster-type energy transfer from Am–GQDs to MoS2 layers, and consequently the quenching of the fluorescence of Am–GQDs. Shifting of the Fermi level of Am–GQDs towards the conduction band in the van der Waals stacked heterostructures further validated the charge transfer-based quenching mechanism.145

4. Properties of GQDs/modified-GQDs

4.1. Physiochemical characteristics

GQDs are basically anisotropic sub-domains of graphene with lateral dimensions below 10 nm and thickness of a few nanometres. The high level of crystallinity in GQDs make them different from CQDs due to the presence of predominant sp2 carbon domains composed of one or few layers of graphene, while CQDs are less crystalline quasi-spherical particles with a mixture of sp2 and sp3 hybridized carbon.146 Moreover, the structure and chemical features of GQDs can be modified by heteroatom-doping and surface-functionalization.146 Transmission electron microscopy (TEM) and high-resolution TEM (HRTEM) are routinely applied to understand the morphology, size distribution, and structural features of GQDs/modified-GQDs. For example, the lateral size/crystallinity of N-GQDs (synthesized from the single precursor ortho-phenylenediamine (o-PDA), QY: 80%) measured from TEM/HRTEM images (Fig. 1a/1b) depicted the size distribution (average size)/lattice spacing of 1–8 nm (3.8 nm)/0.24 nm (corresponding to (110) diffraction plane), respectively. Moreover, fast Fourier transform (FFT) analysis further confirmed the six-fold symmetry and good crystallinity of the synthesized N-GQDs (Fig. 1c).147 Besides, atomic force microscopy (AFM) clearly indicates the number of layers in the GQDs according to the height analysis. The three-dimensional (3D) AFM image and corresponding topographic heights of B,N-GQDs showed nearly uniform heights of ∼0.5 nm (Fig. 1d), indicating the single-layer characteristic of B,N-GQDs.148 Alternatively, the thickness of 1–1.5 nm measured from the AFM topographic image of N-GQDs illustrates 3–5 layers of graphene (Fig. 1e and f).149
image file: d5ra04935k-f1.tif
Fig. 1 TEM image with size distribution bar-diagram (a), HRTEM image (b), and FFT pattern (c) of N-GQDs. Reproduced/adapted from ref. 147 with permission from The Royal Society of Chemistry, 2019. (d) 3D AFM image and the corresponding height pattern of B,N-GQDs. Reproduced/adapted from ref. 148 with permission from The Royal Society of Chemistry, 2017. Topographical AFM image (e) and corresponding thickness–diameter graph (f) of N-GQDs. Reprinted from ref. 149, copyright 2025, with permission from Elsevier.

The crystallinity and disordered characteristics of GQDs/modified-GQDs can also be judged by Raman spectroscopy. For instance, N-GQDs synthesized from a CA/NH4OH mixture exhibited the typical Raman peaks of a carbon material at ∼1340 and ∼1600 cm−1, corresponding to the D and G bands, respectively (Fig. 2a). Moreover, the large ID/IG peak intensity ratio of 0.99 (Fig. 2a) indicated the presence of significantly high defect levels in the N-GQDs due to the incorporation of nitrogen atoms in their lattice.150 The powder X-ray diffraction (PXRD) pattern of GQDs/modified-GQDs generally shows a peak in the 2θ range of 20–25° due to the presence of a graphitic structure. For example, MW-synthesized GQDs using Azadirachta indica (neem) leaf extract exhibited slightly broad peak at the 2θ value of ∼21.12° (Fig. 2b), indicating the graphitic structure of GQDs with a small content of amorphous carbon.151 Fourier transform IR (FTIR) and X-ray photoelectron spectroscopy (XPS) are effective tools for the identification of various functional groups and elemental compositions present in GQDs/modified-GQDs. For instance, the FTIR spectra of alkali lignin (AL) and GQDs derived from AL are shown in Fig. 2c. The presence of peaks at 3443/1660/1415 cm−1 indicated –OH/–COOH/C–N-enriched GQDs with the successful doping of nitrogen element. The peaks at 1590 (C[double bond, length as m-dash]C vibration) and 1049/827 cm−1 (C–H vibration) are due to the aromatic domains of GQDs. Moreover, the insignificant peak at 1190 cm−1 (due to C–O–C stretching vibration) in GQDs compared to AL suggested the breaking of the ether bond and formation of oxygen functionalities during the synthesis process.152


image file: d5ra04935k-f2.tif
Fig. 2 (a) Raman spectrum of N-GQDs showing D and G bands and the ID/IG intensity ratio. Reprinted (adapted) with permission from ref. 150, copyright 2024, the American Chemical Society. (b) PXRD profile of GQDs showing a prominent peak at a 2θ value of 21.1°. Reproduced/adapted from ref. 151 with permission from The Royal Society of Chemistry, 2025. (c) FTIR spectra of AL and GQDs synthesized from AL. Reprinted from ref. 152, copyright 2021, with permission from Elsevier.

The full scan XPS spectra of five types of bioresource-derived GQDs indicated the presence of C (283.9 eV), O (530 eV), N (398 eV), and S (167.9 eV) elements, along with the Na element (adsorbed/bonded with GQDs during synthesis because of the involvement of NaOH electrolyte, Fig. 3a). Moreover, the high-resolution XPS results of lignin-derived L-GQDs further confirmed the existence of various bonding features corresponding to C 1s (C[double bond, length as m-dash]C: 284.4 eV, C–N/C–S: 285.4 eV, C–O: 286.2 eV, C[double bond, length as m-dash]O: 287.3 eV, and COOH: 288.1 eV; Fig. 3b), S 2p (thiophene: 165.5 eV, SOx: 167.7 eV, and sulfone bridge: 168.9 eV; Fig. 3c), N 1s (adsorbed N: 397.4 eV, amino nitrogen: 399.3 eV, and pyrrolic nitrogen: 399.9 eV; Fig. 3d), and O 1s (C–OH: 530.9 eV, –COOH: 532.1 eV, C–O–C: 533.5 eV, and O–Na: 535.6 eV; Fig. 3e), complementing the successful synthesis of N,S-GQDs with 74.2/21.2/1.9/2.8 at% of carbon/oxygen/nitrogen/sulfur elements, respectively.111


image file: d5ra04935k-f3.tif
Fig. 3 (a) XPS survey scans of five bioresource-synthesized GQDs. High-resolution XPS patterns of lignin-derived L-GQDs corresponding to (b) C 1s, (c) S 2p, (d) N 1s, and (e) O 1s elements. Reprinted (adapted) with permission from ref. 111, copyright 2022, the American Chemical Society.

4.2. Optical characteristics

The UV-visible absorption spectrum of GQDs/modified GQDs generally features strong absorption below 300 nm (UV region) due to the π–π* electronic transitions from their conjugated domain and extended tails towards the visible/near IR (NIR) region (sometimes with shoulder peaks) due to the n–π* transitions from their surface functional groups.120,121,153 For instance, the absorption spectra of N-GQDs, S-GQDs, and B-GQDs exhibited characteristic absorption peaks corresponding to the π–π* and n–π* transitions at different positions (Fig. 4a). The peak corresponding to π–π* transition is more prominent in N-GQDs (234 nm) in comparison to S-GQDs (206 nm) and B-GQDs (240 nm), suggesting a relatively high amount of sp2 hybridized moieties (C[double bond, length as m-dash]C backbone) in N-GQDs due to the electron-donating ability of the nitrogen element and less amount of oxygen-containing functional groups.153
image file: d5ra04935k-f4.tif
Fig. 4 (a) UV-visible absorption spectra of N-GQDs, S-GQDs, and B-GQDs. Reprinted from ref. 153, copyright 2024, with permission from Elsevier. (b) Fluorescence spectra of FA,His,Ser–B,P-GQDs according to different λex and their 3D mapping. Reproduced/adapted from ref. 166 with permission from The Royal Society of Chemistry, 2024. (c) Fluorescence spectra of GQDs at λex values from 300 to 400 nm, showing their EDPL characteristic. Reprinted from ref. 167, copyright 2023, with permission from Elsevier. UCPL spectra within the λex value of 700–730 nm (d) and possible mechanism (e) of Er-GQDs. Reprinted from ref. 171, copyright 2020, with permission from Elsevier.

Room temperature PL is one of the most attractive features of GQDs/modified-GQDs for sensing and other fluorescence-related applications, which usually originates from their non-zero band gap structures due to their confined size effect, chemical doping, and surface passivation/functional groups. These parameters can be effectively tuned to obtain GQDs/modified-GQDs with emission features ranging from deep UV,154 to blue,155 green,156 yellow,157 orange,74 red,158 and NIR.159 Apart from their pronounced quantum confinement effect, the abundant edge states and functional groups of GQDs/modified-GQDs play a vital role in dictating their PL properties. Yan et al.160 demonstrated a gradual narrowing in the band gap of GQDs and their corresponding fluorescence color (green to red) using two different functionalization strategies, as follows: (i) lowering the position of π* orbitals by enlarging the π conjugation system in GQDs through covalent-functionalization with poly-aromatic compounds and (ii) creating an n orbital between the π and π* orbitals of GQDs via conjugation with electron-donating functionalities. This precise band gap tailoring from 2.40 eV to 2.05/1.95/1.91/1.88 eV (approach i) and 2.40 eV to 2.08/2.02/1.94 eV (approach ii) opened the possibility to engineer GQDs with different emission-wavelength (λem) characteristics. The degree of surface oxidation also influences the optical properties of GQDs. It was observed that by increasing the degree of oxidation in GQDs, their maximum emission peak shifted towards the higher wavelength side due to the presence of more surface defects.161,162 Among the three types of N-GQDs (N-GQDs-1, N-GQDs-2, and N-GQDs-3) synthesized via the DC microplasma method, the λem of N-GQDs-3 (532 nm) is significantly red-shifted with an enhanced peak intensity in comparison to N-GQDs-1 (459 nm) and N-GQDs-2 (∼468 nm), enabling emission tunable engineering by varying the heteroatom-doping configurations and surface functionalities. XPS results revealed that N-GQDs-3 possessed a larger amount of pyrrolic nitrogen along with an exclusive pyridinic nitrogen configuration, which intensified the electron density, and therefore the band gap narrowing caused a significant red-shift in their λem. Moreover, the high pyrrolic nitrogen content in N-GQDs-3 effectively minimized the emissive traps to result a high QY of 30.1% (QYs of N-GQDs-1 and N-GQDs-2: 4.68% and 1.74%).163

Both excitation-independent PL (EIPL)164–166 and excitation-dependent PL (EDPL)120,121,149,167 characteristics are observed in GQDs/modified-GQDs. For example, the appearance of a single emission peak (∼550 nm) with a variation in the excitation wavelength (λex: 300–550 nm) from the functionalized/doped FA,His,Ser–B,P-GQDs (FA: folic acid, His: histidine, and Ser: serine) indicated EIPL behaviour (Fig. 4b) due to the optimized band structure with predominantly single fluorescence center (n–π* transition).166 Alternatively, the GQDs synthesized from spent tea leaves showed a gradual red shift in λem with λex in the range of 300–400 nm (Fig. 4c), which is attributed to the different size effect and presence of oxygen-containing functional groups on the surface of GQDs.167 Although the exact mechanism is unclear, the EDPL with multicolor-emission features of GQDs-system is frequently explained by the quantum size effect and surface/edge states.120,149,167 GQDs/modified-GQDs may also exert more than one emission peaks at a single λex, which is advantageous for ratiometric fluorescence-based analytical applications. Experimental and theoretical results suggested that the triple emission peaks (599/640/710 nm at λex: 460–640 nm, EIPL behaviour) of red-fluorescent N-GQDs arise from the pyrrolic/pyridinic/amino nitrogen types of emissive states, while graphitic nitrogen in their structure is responsible for their good QY (35%).168

The multi-photon activation process may generate an anti-Stokes luminescence (shorter λem than λex) in the form of up-conversion PL (UCPL), which is generally governed by energy transfer, excited-state absorption, and photon avalanche mechanism.169,170 The UCPL spectra of erbium (Er)-doped GQDs (Er-GQDs) under λex of 700–730 nm are shown in Fig. 4d, which depicted emission peaks in the range of 437–442 nm. The UCPL phenomenon is explained based on the triplet–triplet annihilation (TTA) mechanism, which involved three stages, as follows: (i) excitation of sensitizer (GQDs) to singlet excited state (1ES), followed by intersystem crossing (ISC) to triplet state (3ES), (ii) triplet-type energy transfer (TTET) from GQDs 3ES to annihilator/acceptor (here Er3+) triplet state (3A), and (iii) recombination of two triplet states into one as a relaxation phenomenon to the ground state and the other to the singlet excitation state of an annihilator (1A), which actually generates emission (Fig. 4e).171

A high QY (ratio of emitted photons with respect to adsorbed photons during radiation-induced process) of GQDs/modified-GQDs is a direct reflection of their high PL intensity, which is relevant to their fluorescence-based application and can be achieved by suitable elemental-doping and surface-functionalization/passivation.

4.2.1. Effect of heteroatom doping. The electronegativity of nitrogen/boron (3.04/2.04) is quite different from carbon (2.55), which in turn effectively alters the electronic nature of doped-GQDs. Nitrogen doping in GQDs is found to be beneficial for a significant enhancement in their QY due to their chemical/electronic structure modulation. For instance, an ultrahigh QY of 99% with 98 nm Stokes shift was achieved in N-GQDs (25.91% nitrogen content, synthesized via thermal treatment of GO and PEI).172 A better QY of N-GQDs (19.1%, band gap: 0.249 eV) compared to bare GQDs (1.7%, band gap: 2.826 eV) is explained based on the smaller band gap and nitrogen atom-centered electronic transitions in the N-GQDs.173 The high QY of HT-synthesized N-GQDs (80.31%) is ascribed to the conversion of their non-radiative centers (–COOH groups) to radiative electron–hole recombination moieties (C–N/–C[double bond, length as m-dash]N).153 The intense quantum confinement and edge effects in the crystalline B-GQDs (4.8% boron content in the form of BC2O and BCO2) caused localized electron–hole pair generation for the optimum band gaps and an impressive QY of 22.7%.174 The electronegativity of sulfur (2.58) is quite close to carbon (2.55). As a result, the charge-transfer in the C–S bonds is expected to be low, and therefore S-GQDs have low QYs (10.6/10.2%).175,176 Doping of the phosphorus element in the carbon framework may also modulate the optical property and EC activity of the doped-counterpart. Besides inducing polarization in the P–C bond, coupling between the 2p of carbon and 3p orbitals of phosphorus (in PC3 configuration) may promote a near-Fermi level electron density, resulting in a lower band gap and better electron transfer activity.177 Phosphorus-doped GQDs (P-GQDs) synthesized through the ST method (precursor: glucose and triphenylphosphine) at 180 °C, 210 °C, and 240 °C showed QYs as high as 26.2%, 37.66%, and 41.84%, respectively. XPS results revealed that P-GQDs synthesized at 240 °C contain a higher amount of phosphorus-element (4.82 at%) with predominant PC3-structure for effective polarization and electron redistribution in comparison to that prepared at 180 °C (4.19 at% phosphorus) and 210 °C (4.57 at% phosphorus), which are mainly composed of PO4 and PO3 bond structures.177

Apart from single heteroatom-doping, dual-elemental doping in the form of B,N-GQDs, boron/sulfur co-doped GQDs (B,S-GQDs), N,P-GQDs, and N,S-GQDs has also employed to improve the optical properties of GQDs. The advantage of boron/nitrogen co-doping (0.9/7.5%) in B,N-GQDs can be revealed by their high QY (75%) in comparison to N-GQDs (71%) and B-GQDs (23%).178 B,S-GQDs synthesized via the pyrolysis of CA, H3BO3, and 3-mercaptopropionic acid showed much a higher QY (19.8%) in comparison to undoped GQDs (7.5%), which is ascribed to the modulation of their electronic structure and passivation of non-radiative recombination sites.179 The one-pot ST treatment of resorcinol and phosphonitrilic chloride trimer yielded green-emitting N,P-GQDs (nitrogen/phosphorus content: 3.05/1.81%) with a QY as high as 58.2%.180 The concurrent incorporation of nitrogen and sulfur in GQDs is also beneficial for improving their optical properties. For example, HT treatment of GA in the presence of urea and 1-octanethiol afforded blue-emissive N,S-GQDs with a QY as high as 70%.181

The incorporation of MIs in GQDs can also expand the scope of the doping strategy through a synergistic effect to enhance their fluorescence signal and QY. For example, manganese ion (Mn2+)-bonded B,N-GQDs showed a much higher QY (30.52%) compared to B,N-GQDs (20.12%), which is ascribed to the confinement effect between the surface functionality and Mn2+ to produce a uniform shape/size.182 Neodymium (Nd)-doped N-GQDs (Nd,N-GQDs; ∼1 at% Nd) synthesized via the MW method showed NIR fluorescence and QY up to 62%.183 Terbium (Tb)-doping in the GQDs resulted in almost twice the emission intensity at 452 nm in comparison to the bare GQDs due to the suppression of non-radiative recombination sites, and therefore a high QY (52%).184 Recently, Fe3+-chelation and nitrogen-doping simultaneously improved the PL intensity of Fe,N-GQDs and the QY was as high as 67%.185

4.2.2. Effect of functionalization. Non-covalently or covalently modified GQDs also showed better QY compared to bare GQDs. For example, non-covalent attachment of poly-L-lysine (PLL) on the surface of GQDs resulted in PLL@GQDs, which showed a significantly higher QY (41.33%) in comparison to bare GQDs (18.89%).186 An exceptionally high QY of 99.8% was achieved by the in situ attachment of D-penicillamine (DPA) over GQDs to minimize structural defects and enhance the quantum size effect.187 Benefitting from the large electron-donating circumstances by the covalent attachment of pentaethylenehexamine (PEHA) and DPA on GQDs, the resultant co-functionalized PEHA,DPA–GQDs (amine/hydroxyl/carboxyl groups rich) exhibited strong fluorescence (λem: 450 nm) and a QY as high as 90.91%.188 The time-dependent density functional theory (DFT) calculation revealed that sp3-type functional groups (O, OH or F) on the surface of GQDs can significantly improve the PL intensity as well as QY of functionalized GQDs due to the restriction of excited carriers on the graphitic layers and enlargement of the transition dipole moment during radiative recombination.189

4.3. CL and ECL characteristics

Besides PL, CL and electro-generated CL (ECL) are other intriguing features of GQDs/modified-GQDs for analytical purpose. CL emissions originate when a substance absorbs chemical energy (produced during chemical reaction) to reach the excited state, and then returns to the ground state. A variety of chemical initiators such as H2O2,190 permanganate-sulfite,191 and cerium ions (Ce4+),192 along with GQDs/modified-GQDs can produce strong CL signals. Three types of GQDs (bare GQDs, N-GQDs, and N,S-GQDs) utilized the exothermic energy released by the chemical reaction between bis(2,4,5-trichloro-6-carbopentoxyphenyl) oxalate and H2O2 in the presence of a base catalyst (sodium salicylate) to produce yellowish-white, green, and blue CL emissions. Among them, N,S-GQDs showed 5/2.5-fold higher CL efficiency than bare GQDs/N-GQDs, demonstrating that dual-doping is favorable to create efficient intrinsic-emissive surface states for substantial CL signal.193

ECL is a light-emitting phenomenon due to the electron transfer reaction of electrochemically generated radical species to form an excited state, and subsequent emission process, while returning back to the ground state.105,194 The ion-annihilation (excited state formation via electron transfer between the cation and anion radicals of luminophore itself) and co-reactant mediated (formation of radical species from co-reactant to react with luminophore and form excited state) routes are implemented to generate ECL systems; however, the latter is more common and can produce a strong ECL signal.81,105 The first observation of ECL from N-GQDs (greenish-yellow luminescence) in the presence of 0.1 M K2S2O8 co-reactant was reported as far back as early 2012.195 A schematic of NIR-ECL generation through the GQDs/K2S2O8 co-reactant system is shown in Fig. 5. By applying a negative potential, peroxodisulfate (S2O82−) and GQDs get reduced to the corresponding radical anions nearby the cathode surface. Thereafter, the second reduction of the GQD˙ radical anion to GQD2− dianion and its subsequent reaction with neutral GQDs generated two GQDs˙ radical anions. At the same time, the sulphate radical anion (SO4˙) is produced from the S2O83−˙ radical anion by the loss of the SO42− anion, which is further reduced by the GQD˙ radical anion to generate excited-state GQDs* and SO42− anion. Furthermore, the relaxation of GQDs* to the surface excited-state GQDsS*, followed by returning to the ground state resulted in the emission of a strong NIR-ECL signal.105


image file: d5ra04935k-f5.tif
Fig. 5 Schematic of the possible mechanism for the generation of NIR-ECL from the GQDs/K2S2O8 system. Reprinted (adapted) with permission from ref. 105, copyright 2021, the American Chemical Society.

4.4. EC characteristics

The large surface-to-volume ratio, abundant active sites, reliable electrical conductivity, and rapid electron transfer ability of GQDs/modified-GQDs qualify them as electrode modifiers for amplified EC signals. Both GQDs and modified-GQDs have been used to enhance the EC activity. For instance, a significant reduction in the electron transfer resistance with a GQDs/Nafion modified glassy carbon electrode (GCE) (GQDs/Nafion@GCE) in comparison to Nafion@GCE indicated favorable electron transfer kinetics in the GQD/Nafion electrode material, and therefore better EC activity.196 Around 10% nitrogen-doping and aromatic structure of N-GQDs provided surplus available electrons for enhanced conductivity and electrocatalytic activity.197 Covalently functionalized DMC–GQDs showed a better current response compared with bare GQDs, which is attributed to the active surface area and fast charge transfer rate of modified-GQDs.103 Dual-doping is also beneficial to enhance the conductivity and electrocatalytic activity of N,S-GQDs. This is possible due to the incorporation of bonding and anti-bonding sulfur orbitals between the bonding orbital of carbon and anti-bonding orbital of nitrogen, which improved the electron availability and narrowing of the band gap structure in the N,S-GQDs (band gap of N,S-GQDs/N-GQDs: 0.98/1.3 eV).198

5. Diverse applications of GQDs-based systems

Carbon-based dots are one of the preferable alternatives, especially where metal-containing QDs are employed in a specific application.199 The effective quantum confinement, good crystallinity, hydrophilicity, high surface-to-volume ratio, chemical stability, favourable electronic structures, facile charge transportation, and biocompatibility of GQDs permit them to be used in biological, optoelectronic, agricultural, environmental, and energy-related applications. Additionally, the application prospects/outcomes of GQDs can be further broadened by tuning their intrinsic properties through heteroatom-doping and post-functionalization. The compositing/heterostructuring of GQDs/modified-GQDs with other active counterparts is also a noteworthy route to improve their performance metrics. The wide range of applications of GQDs-based systems can be identified from the following representative examples: GQDs synthesized from neem extract showed bioactivity with better antibacterial and antioxidant performances compared to the starting extract.151 Nd,N-GQDs and thulium (Tm)-doped N-GQDs have shown potential as contrast agents for dual-mode biomedical imaging (ultrasound and NIR fluorescence) applications. Intravenously injected doped-GQDs (particularly, Nd,N-GQDs) in mice and animal organs significantly enhanced the response of both modes to achieve precise, accurate, and sensitive imaging for diagnostic and monitoring during therapeutic treatment.159 Due to the light-induced reactive oxygen species generation ability of GODs, GQDs with 132 conjugated carbon atoms (single-molecule) were applied for successful cancer eradication via photodynamic therapy (both in vitro and in vivo).200 Blue- and green-emissive GQDs were coupled with poly(sodium 4-styrenesulfonate) fibers to construct a nanocomposite with colour-tuneable and white light emission activity.201 Biocompatible GQDs were employed for boosting sustainable agricultural activity and showed a significant enhancement in nitrogen fixation (471.7% activities with respect to the control of Azotobacter vinelandii), along with an increase in the nitrogen content in soil.202 The nanocompositing of GQDs with graphitic carbon nitride (g-C3N4) resulted in an efficient photocatalyst for the degradation of Rhodamine B dye (optimum degradation efficiency: 95.2% within 2 h, catalyst concentration: 35 mg l−1).203 N-GQDs NiAl LDH/TiO2 (LDH = layered double hydroxide) heterostructures with 10 wt% N-GQDs loading exhibited efficient hydrogen production capability (1332.11 µmol g−1 h−1) via their photocatalytic activity under solar irradiation. Here, N-GQDs played a crucial role in mediating facile charge passage between NiAl LDH and TiO2 to delay electron–hole recombination.204 N-GQDs pillared Ta4C3Tx MXene exhibited a significant improvement in gravimetric capacitance (701 F g−1; bare MXene: 100 F g−1) and applicability to fabricate an asymmetric supercapacitor with specific capacity, energy density, power density, and cyclic stability of 110C g−1, 55 Wh kg−1, 9000 W kg−1, and 20[thin space (1/6-em)]000 cycles, respectively.205

Sensing of pollutants, bio-related species, and other items is another fascinating area of research, which is continuously progressing with the aim of improving the performance, utility and cost of existing probes, and applicability of user friendly detection techniques. Owing to the advantage of abundant functional groups and edge sites in GQDs/modified-GQDs, along with size-based confinement origin and other intriguing features, GQD-based systems have been extensively explored as a low cost and effective platform for the targeting of various analytes. The structural/compositional characteristics of GQDs and modified-GQDs make them suitable to interact with target substances selectively and respond accordingly. GQDs-based FL sensors function based on the quenching or enhancement of their fluorescence response when they contact the analyte. A colour change in the probe solution in the presence of analyte constitutes a COL sensor, which provides an opportunity for the visual monitoring and quantification of analytes using UV-visible absorbance spectra. Changes in the CL and ECL signals (generated from GQDs-based platforms) can also be monitored to quantify various analytes. The EC sensing strategy refers to analyzing substances through the electrode surface-confined charge transfer phenomenon, and consequently changes in the current/voltage response. Voltammetry-based methods including cyclic voltammetry (CV), differential pulse voltammetry (DPV), square wave voltammetry (SWV), and anodic stripping voltammetry (ASV) have been potentially used during the EC detection process.206 Some recent sensing attributes of GQDs-based systems are as follows: GQDs (band gap engineered from 3.3 to 1.9 eV by varying the graphitic core size of GQDs) were used as a photo-sensitizer to tune the dynamics of carrier transfer in GQDs-decorated In2O3. Consequently, the optimum GQDs (7 nm)/In2O3 sensor system showed efficient monitoring of environmentally hazardous gas NO2 under the visible light (blue)-activated photocatalytic sensing strategy with a recognizable response (Rg/Ra) of 97.1 at 1 ppm and rapid response/recovery time of 136/100 s.207 N-GQDs were coupled with a Prussian blue (PB) analogue-containing PB layer to assemble a wearable biosensor for the EC detection of H2O2 with sensitivity as high as 221.29 ± 1.77 µA mM−1 cm−2. Additionally, the immobilization of glucose oxidase on the conjugated composite resulted in a nanocomposite assembly for the selective monitoring of glucose with high sensitivity (90.49 ± 1.08 µA mM−1 cm−2).208 An enhancement in the CL response of luminol-H2O2 by incorporating N-GQDs, and its subsequent suppression in the presence of tetrabromobisphenol A (TBBPA; an environmental contaminant) were integrated with an SiO2@TBBPA MIP (MIP = molecular imprinted polymer) assembly to detect TBBPA with a detection limit of 0.032 nM.209 Coating of Co-modified exfoliated zirconium phosphate on the functionalized GQDs (His–GQDs) resulted in a synergistic electrocatalyst for the EC detection of methyl parathion (a toxic pesticide) with a detection limit of 10 nM and sensitivity up to 0.85 mA µM−1.210

Various metals in ionic form and many of inorganic anions are potential pollutants in the environment and living organisms. Conversely, alkali/alkaline-earth MIs are biologically important to regulate and control metabolic cycles. Therefore, the identification of inorganic ions with good sensitivity, selectivity, and easily implemented detection methods using a facile probe is an obvious environmental and biological concern. Moreover, the deployment of sensors and sensing strategies for reliable, on-site, real sample/water-body, and track-ability detection in living systems is also vitally important. Here, we discuss various GQDs-based and GQDs involved systems for the detection of inorganic ions, employing various sensing approaches.

6. GQDs-based/involved sensors in the detection of HMIs and other MIs

6.1. Fe3+

6.1.1. Doped- and undoped GQDs. In an earlier report (2013), green-emitting graphitic N-GQDs with abundant oxygen-containing functionalities specifically interacted with Fe3+ (high binding affinity with phenolic –OH groups) for fluorescence quenching to achieve a low limit of detection (LOD, Table 2).211 Subsequently, single- and dual-heteroatom doped-GQDs, MIs doped-GQDs, and undoped GQDs have been extensively explored for the assay of Fe3+ (Tables 2 and S1). Among the single-heteroatom doped-GQDs, N-GQDs are one of the preferred choices for the FL turn-off based detection of Fe3+ (Tables 2 and S1). For instance, bottom-up-synthesized N-GQDs (15.7% nitrogen) showed better sensitivity (Table 2) compared to the undoped GQDs (linear range (LR): 1–594 µM), indicating the nitrogen element-induced modification of the chemical-electronic structure for effective complexation between N-GQDs and Fe3+. However, due to the significant quenching effect of Hg2+/Cu2+ on the N-GQDs, masking agents were required to circumvent their interference.212 N-GQDs (bluish-green fluorescence) synthesized via the ST method showed the possibility for the selective detection of Fe3+ in a wide LR (Table 2), which is explained based on their strong affinity with surface-bound –NH2/–COOH groups to promote electron transfer between them (rapid quenching within 30 s). However, mechanistic details were beyond the scope of that investigation.213 Li et al.191 first employed N-GQDs (9.25% nitrogen) to sense Fe3+ via the CL method. The N-GQDs catalyzed the KMnO4–Na2S redox reaction to measurably enhance CL signal (20-fold) by the involvement of ˙OH radical. Importantly, undoped GQDs did not showed this improvement in the CL response, which indicates the crucial role of nitrogen-doping (particularly, pyridinic configuration) for catalyzing the CL reaction. With the addition of Fe3+, the CL signal gradually decreased due to the chelation effect. Although this probe showed selectivity in the presence of other interfering MIs within the concentration of Fe3+ (maximum: 1 µM), the high concentrations of ionic contaminants in real samples limit its wide applicability. The utility of biomass (Marigold)-derived N-GQDs in the selective recognition of Fe3+ with satisfactory sensitivity (Table 2) and tracking of Fe3+ in HeLa cells (human epithelial cancer cells), along with its quantification in real water specimens opens a sustainable possibility for the probing of Fe3+.214
Table 2 GQDs, modified-GQDs, and GQDs involved with other counterparts for Fe3+ and Fe2+ sensing applicationa
GQDs-based sensor Synthesis conditions Size range/average sizeb (nm) QY (%) Sensing process LR (µM) LOD (µM) Ref.
a SPTS: sodium p-toluenesulfonate, TBAP: tetrabutylammonium perchlorate, DMSO: dimethylsulfoxide, DCM: dichloromethane.b Measured from TEM.c Size range/average size of GQDs involved system.d Size range/average size of GQDs/doped-GQDs used with other counterparts.e Absolute QY.f QY after Fe3+ binding.g LR in µg mL−1.h LR/LOD measured from 358 nm emission peak.i LR/LOD measured from 408 nm emission peak.j LR/LOD measured from 358 nm excitation peak.k LR/LOD measured from down-conversion PL.l LR/LOD measured from UCPL.m Dynamic concentration range and corresponding LOD.n LOD predicted from machine learning algorithm.o Analytical ability in real water/biological fluid/supplement samples.p Analytical ability in living cells.q Paper-based sensing capability.
Fe3+
Doped-/undoped GQDs
N-GQDs Carbonization of pyrene with HNO3:H2SO4 (1[thin space (1/6-em)]:[thin space (1/6-em)]3) under reflux (95 °C, 48 h); centrifugation; HT with hydrazine hydrate/25 wt% NH3 (180 °C, 24 h); centrifugation; dialysis 5–10/5.5 11.7 FL, turn-off 0.5–20 0.005 211
N-GQDs Pyrolysis (CA, 200 °C, 30 min); dissolved in 10 mg per mL NaOH & pH adjusted to 8.0; HT treatment with 30% hydrazine (180 °C, 12 h); centrifugation 2.2–5.3/3.8 23.3 FL, turn-off 1–1105 0.09 212o
N-GQDs ST (GSH/AgNO3 in ethylene glycol, 200 °C, 12 h); centrifugation 1–5/2.5 FL, turn-off 50–2000 0.07 213o
N-GQDs HT (CA/urea in water, 200 °C, 6 h); filtration; dialysis 3.5–6/4.8 CL, turn-off 0.01–1 0.004 191o
N-GQDs Pyrolysis (Marigold granules, 1000 °C, 5 h, Ar); acid oxidation with HNO3:H2SO4 (1[thin space (1/6-em)]:[thin space (1/6-em)]3) under reflux (90 °C, 5 h); filtration; pH adjusted to 7.0; dialysis; drying; HT (obtained powder in ethylenediamine solution, 200 °C, 10 h) 1.5–4.5/3.2 7.84 FL, turn-off 0–20, 200–667 0.0411, 0.5 214o,p
N-GQDs ST (GO in DMF, 200 °C, 5 h); centrifugation; filtration 1.1–5.3/3.17 14.32e FL, turn-off 0–34 0.00238 165o,q
N-GQDs HT (Bamboo fiber powder/urea in water, 200 °C, 8 h); filtration; dialysis 2–20/5 40.36 FL, turn-off 1–1000 0.034 215p
N-GQDs HT (aspartic acid/urea in water, 180 °C, 8 h); centrifugation 1–4/2.22   FL, turn-off 100–600g 216
Mg,N-GQDs HT (aspartic acid/urea/MgCl2·6H2O in water, 180 °C, 8 h); centrifugation 0.8–2/1.31 " 150–450g "
S-GQDs Electrolysis of graphite rod in 0.1 M SPTS aqueous solution, 3 h; filtration, dialysis 2–4/3 10.6 FL, turn-off 0.01–0.70 0.0042 175o
B-GQDs Electrolysis of graphite rod in 0.1 M borax aqueous solution, 2 h; filtration, dialysis 3–6/4.5 5.2 FL, turn-off 0.01–100 0.005 217o
N,S-GQDs HT (acid hydrotrope fractionation of Miscanthus/p-amino-benzene sulfonic acid monosodium salt in water, 200 °C, 12 h); filtration; dialysis —/4.05 20.2 FL, turn-off 0–10.6h, 10.6–900h 0.00141h 218
" 0–10.6i, 10.6–800i 0.00231i
" 0–10.6j, 10.6–1000j 0.00209j
Er-GQDs HT (Lactose/Er(NO3)3·5H2O in water, 200 °C, 4 h); filtration; dialysis 2–8/4.7 18 FL, turn-off 0.01–1k, 1–120k 0.0028k 171o
" 0.1–20l, 20–200l 0.028l
GQDs HT (Rice husk powder in water, 150 °C, 5 h); filtration; centrifugation 2.5–5.5/3.9 8.8 FL, turn-off 0–300 0.0058 220
GQDs-1 HT (2 mg per mL GO in water, pH adjusted to 9.5, 130 °C, 10 h); filtration; freeze drying 4–8/5.8 6 FL, turn-off 1–8.75 0.136 221
GQDs-2 HT (2 mg per mL GO in water, pH adjusted to 8.0, 175 °C, 10 h); filtration; freeze drying " 8.9 " 1–75 1.36
[thin space (1/6-em)]
Functionalized GQDs
RBD–N-GQDs Electrolysis of graphite rod in 0.01 M TBAP/DMSO, 3 h; centrifugation & drying; acid oxidation with HNO3:H2SO4 (1[thin space (1/6-em)]:[thin space (1/6-em)]3) under reflux (100 °C, 24 h); pH adjusted to 7.0; dialysis; covalently modified with RBD 3.5–6.5/5 43f FL, turn-on 0–1 0.02 133p
DA–GQDs Pyrolysis (CA, 200 °C, 25 min); mixed in 10 mg per mL NaOH solution & pH adjusted to 7.0; covalently modified with DA 2–9/4.5 10.2 FL, turn-off 0.02–1.5 0.0076 222
DPA–GQDs HT (CA/DPA in water, 200 °C, 2.5 h); dissolved in water; dialysis 1–9/4.7 99.8 FL, turn-off 4–1800 1.2 187o
Am–GQDs Carbonization (pre-oxidized Asphalt, 900 °C, 1 h, He); acid oxidation with HNO3:H2SO4 (1[thin space (1/6-em)]:[thin space (1/6-em)]2) under ultrasonication (1 h) & reflux (100 °C, 23 h); diluted with water & pH adjusted to 7.0 by NH3; HT (180 °C, 6 h); dialysis 2–3.6/2.3 13.8 FL, turn-off 0–50 5.1 × 10−4 164o
N-GQDs@xylan Liquid phase exfoliation of graphite flake in NMP/0.1 g NaOH under bath/probe ultrasonication (8 h/4 h); dialysis; filtration; non-covalently modified with 5% xylan under HT (180 °C, 5 h) 1–3/1.97 36.63 FL, turn-off 0–75 0.0928 142
Arg,Ser–B-GQDs Pyrolysis (CA/Arg/Ser/H3BO3, 160 °C, 4 h); diluted with water; centrifugation; dialysis 1.0–12/4.8 40.12 FL, turn-off 0–50 0.075 223o
[thin space (1/6-em)]
GQDs involved with other counterparts
LS/GQDs Pyrolysis (CA·H2O, 200 °C, 15 min); treated with 10 mg per mL NaOH containing 20 µL LS under stirring (2 h); pH adjusted to 7.0; dialysis 390–800/590c; 2–8/—d 23.3 FL, turn-off 0.005–500 0.0005 224o
GQDs/PVA@ PETP HT (glucose/NH3 in water, 200 °C, 8 h); dialysis; non-covalently modified with PVA & coated on PETP film 8–17/15.5d FL, turn-off 0–30m 0.1m 225o,q
AuNPs@N-GQDs ST (GO in DMF, 250 °C, 5 h); centrifugation; in situ decorated with AuNPs; centrifugation 12.5–33/23.4c 12.3 FL, turn-off 0.1–0.75; 0.001–10m 0.03; 0.001m 226o
GQDs/CNCmod HT (GO in water, pH adjusted to 9.5, 135 °C, 10 h); filtration; dialysis; incorporated with CNCmod & solvent casted over PETP substrate <10/—d FL, turn-off 1 × 10−9–2 × 10−6 8 × 10−10 227
AuNPs@N-GQDs ST (GO in DMF, 200 °C, 5 h); centrifugation; in situ decoration with AuNPs 10–40/17c; 2–8/4.6d Optical, turn-on 0.001n 228o
GQDs-Au-Ni micromotor GQDs solution purchased from ACS materials; electrochemical template deposition of GQDs layer followed by Au and Ni layers on Ag-coated polycarbonate membrane; etching of Ag; removal of membrane by dissolving in DCM Solid FL, turn-off 1 × 10−6–10 7.0 229
    Magnetic, speed reduction " 9.0
GQDs-Au-Ni@ SPCE GQDs-Au-Ni casted over SPCE electrode EC, DPV " 6.0
[thin space (1/6-em)]
Fe2+
N,S,I-GQDs Pyrolysis (CA, 230 °C, 5 min); pyrolysis (melted CA/garlic extract/KI/KIO3, 230 °C, 5 min); mixed in 0.25 M NaOH solution 2.36–3.78/— 45 FL, turn-off 0.36–3.6, 3.6–17.98 0.4, 1.16 231o
GQDs MW (Mangifera indica leaf residue in water); dispersion in ethanol, centrifugation, filtration & drying; MW (slurry in water, 10 min); drying 1–13/7.1 45 FL, turn-off 0–2.5 4.07 124
FA,His,Ser–B,P-GQDs Pyrolysis (CA/FA/His/Ser/H3BO3/H3PO4, 160 °C, 4 h); dissolved in water; centrifugation; dialysis 2–12/4.6 60.2 FL, turn-off 0.01–50 0.0042 166o


The Fe3+ probing capability of N-GQDs via a quenching-based FL process is illustrated by theoretical calculations. DFT calculations showed a significant increase in the band gap of N-GQDs in the presence of Fe3+ (1.072 eV, band gap of N-GQDs: 0.249 eV, Fig. 6a and b), which illustrates the inhibition of the active sites in the nanoprobe at a higher energy level, and therefore electron transfer from N-GQDs to Fe3+ via chelation kinetics, resulting in fluorescence quenching-based Fe3+ detection (Fig. 6c).173


image file: d5ra04935k-f6.tif
Fig. 6 DFT calculation-based energy level pictures of N-GQDs (a) and N-GQDs along with Fe3+ (b). (c) Schematic of Fe3+ detection via an electron transfer-based fluorescence quenching process. Reprinted from ref. 173, copyright 2022, with permission from Elsevier.

ST-synthesized N-GQDs with an excitation-independent cyan colour emission (λem: 525 nm at 360 nm λex) showed a low LOD of 2.38 nM (LR: 0–34 µM) in the detection of Fe3+ through the synergistic effect of –OH group-driven coordination, static quenching effect (SQE), and inner filter effect (IFE). The N-GQDs could also detect Fe3+ in mouse serum/human urine (biological samples) with good recoveries/relative standard deviations (RSDs) (98.1–104.6/0.12–2.71%) and considerable inter-day/intra-day precision. Furthermore, the portable sensors (hydrogel kit and flexible film; stable up to 1 month under 4 °C storage conditions), conveniently fabricated by immobilizing N-GQDs in a PVA matrix, exhibited a visual as well as on-site detection capability for Fe3+. A gradual decrease in the cyan fluorescence of the hydrogel kit with an increase in Fe3+ concentration (0–34 µM) and subsequent recovery with adenosine triphosphate (ATP, 0–10 µM) can be seen in Fig. 7a and b, respectively. Fig. 7c shows the flexibility of the prepared membrane device (without obvious marks after multiple folding) and accurate colour visibility under UV light (cyan fluorescence) and in the presence of Fe3+ (quenched fluorescence)/ATP (recovered fluorescence). Moreover, the ‘AND’ logic gate of the portable sensor was correctly executed in the sensing operation by utilizing FL/COL dual readout to achieve good accuracy (Fig. 7d).165 Subsequently, Khan et al.215 demonstrated the use of bamboo fiber (biomass)-derived N-GQDs (QY: 40.36%) for the selective detection of Fe3+ with improved sensitivity compared to previous biomass-synthesized N-GQDs (Tables 2 and S1). The coordination between Fe3+ and oxygen-containing functionalities (preferably –OH groups) on the surface of N-GQDs facilitated electron transfer from N-GQDs to Fe3+, and therefore weakening of the inherent photo-induced electron transfer (PET) process to quench the fluorescence of N-GQDs. The involvement of SQE via the formation of a non-fluorescent ground-state complex is evidenced by the high KSV calculated from the Stern–Volmer plot (1.31 × 104 M−1). Although the fluorescence quenching is relatively greater with Fe3+ (∼54%) than Hg2+ (∼26%), the interference from Hg2+ in real water samples cannot be avoided. The blue luminescence of recently synthesized N-GQDs and magnesium (Mg)-doped GQDs (Mg,N-GQDs) diminished due to the coordination of Fe3+ with –OH and –NH2 functional groups present on the doped-GQDs (Fig. 8a). The incorporation of MIs in the N-GQDs did not improve their sensitivity for the detection of Fe3+ (Table 2) and the effect of doping on their selectivity is unclear. The various energy levels and associated electronic transitions before-after the addition of Fe3+ in N-GQDs and Mg,N-GQDs are shown in Fig. 8b and c, respectively, which depict the passage of photo-excited electrons from the doped-GQDs to partially filled Fe3+ orbitals to inhibit the routine radiative process and fluorescence signal.216


image file: d5ra04935k-f7.tif
Fig. 7 Digital images of a hydrogel kit, showing a gradual fluorescence quenching with 0 to 34 µM concentrations of Fe3+ (a) and recovery of fluorescence with 0 to 10 µM concentrations of ATP (b) under a 365 nm UV light. (c) Digital pictures of the membrane under UV light showing flexibility/cyan fluorescence (first two) and turn-off-on response (last two) with 34 µM Fe3+ and subsequent addition of 10 µM ATP. (d) Truth table using the input from Fe3+ and ATP and corresponding “AND” type logic scheme. Reproduced/adapted from ref. 165 with permission from The Royal Society of Chemistry, 2023.

image file: d5ra04935k-f8.tif
Fig. 8 (a) Fluorescence quenching of doped-GQDs after interaction with Fe3+. Existing energy levels and electron transitions in (b) N-GQDs and N-GQDs + Fe3+ and (c) Mg,N-GQDs and Mg,N-GQDs + Fe3+. Reprinted from ref. 216, copyright 2025, with permission from Elsevier.

Single-heteroatom doped S-GQDs for Fe3+ sensing has rarely been reported in the literature (Tables 2 and S1). Blue-green fluorescent S-GQDs (4.25% sulfur) were employed in early years (2014) for the selective sensing of Fe3+ with regeneration ability after quenching operation using ethylenediamine tetraacetic acid (EDTA); however, the narrow LR (0.01–0.7 µM) of this probe limits its wider applicability. Interestingly, the N-GQDs, B-GQDs, and undoped GQDs tested in this study did not exhibit a significant decrease in fluorescence even at 0.7 µM Fe3+, indicating the importance of sulfur in GQDs to promote the coordination of phenolic –OH groups with Fe3+.175 Chen et al.217 first employed B-GQDs (∼3.2% boron) for the selective estimation of Fe3+ with considerable sensitivity (Table 2) and EDTA-induced regenerative characteristics. The quenching-based FL detection is driven by the strong adsorption ability of Fe3+ on the surface of B-GQDs and subsequent energy transfer between them. Subsequently, newly synthesized B-GQDs via the bottom-up method showed a wider LR but with the compromise of higher LOD (Table S1).

N,S-GQDs are found to be suitable fluorescent probes among the two types of dual doped-GQDs (N,S-GQDs and N,P-GQDs) tested for the sensing of Fe3+ (Tables 2 and S1). For instance, valorization of Miscanthus biorefinery waste in the form of N,S-GQDs (nitrogen/sulfur content: 2.53/2.83 wt%) was used for the selective as well as tri-channel sensitive detection of Fe3+. The emission peaks of N,S-GQDs centred at 358/408 nm and excitation peak at 358 nm provided a tri-channel FL platform to sense Fe3+ with high sensitivity and improved precision (LODs (LRs): 1.41 nM (0–900 µM)/2.31 nM (0–800 µM) and 2.09 nM (0–1000 µM)). The XPS, FTIR, and time-resolved PL (lifetime (τ) changed from 11.95 to 9.98 ns after adding Fe3+) results inferred a collision-type dynamic quenching effect (DQE) between the probe and analyte rather than the common SQE in the detection of Fe3+. Noticeably, the sensitivity and tri-channel-based accuracy of this biomass-derived probe for the quantification of Fe3+ surpassed the sensing performance of single-/dual-heteroatom doped-GQDs via the FL method (Tables 2 and S1); however, potential interference from Cr2O72− was detected during the selectivity test.218

The incorporation of MIs in the structure of GQDs to improve their optical properties and selectively detect Fe3+ has also been reported in the literature (Tables 2 and S1). For example, a rare-earth (Er) inclusion in the form of Er-GQDs (1.8 at% Er-doping) with down-conversion/UCPL characteristics at λex of 360/730 nm showed a good Fe3+ sensing performance (Table 2). Interestingly, the UCPL-based detection of Fe3+ showed a larger LOD (28 nM) in comparison to the down-conversion-based detection process (2.8 nM), which is attributed to the weak fluorescence intensity and relatively low quenching response in the up-conversion domain. Moreover, the detection of Fe3+ in human serum also validated the response of the sensor in both down-conversion/up-conversion domains with LODs of 11.2/336 nM, justifying its good analytical performance towards the higher concentration side of Fe3+ in the biological sample.171

The analytical ability of bare GQDs for the detection of Fe3+ is also revealed in Tables 2 and S1. Zhu et al.219 presented insight into the selectivity of GQDs (containing phenolic –OH groups) towards Fe3+ via the formation of GQDs-aggregates under acidic conditions (pH: 3.5). Due to the extremely lower Ksp (solubility-product constant) of Fe(OH)3 (2.8 × 10−39) in comparison to Cu(OH)2/Ni(OH)2/Co(OH)2 (2.2 × 10−20/5.0 × 10−16/2.3 × 10−16) at a lower pH, the formation of Fe(OH)3 induced the aggregation of GQDs, resulting in fluorescence quenching. The potentiality of biomass-derived GQDs (from rice husk powder) in the fluorescence quenching-based detection of Fe3+ can be appreciated by their satisfactory sensing performance (Table 2).220 Two HT conditions for the scissoring of GO (Table 2) showed feasibility for the synthesis of GQDs with different amounts of –COOH/–OH functional groups, and correspondingly varying sensing performance levels. It was observed that high –COOH-containing GQDs-1 resulted in a lower LOD (0.136 µM), while GQDs-2 with a higher QY (8.9%) and –OH groups at their edge were advantageous for a wider LR (1–75 µM, LOD: 1.36 µM) in the turn-off based detection of Fe3+. Here, –COOH groups are considered hard binding sites for Fe3+ (hard HMI) according to the hard-soft acid-base (HSAB) theory, and therefore a lower LOD in the case of GQDs-1.221 Later, biomass-based GQDs and F-rich GQDs were further utilized to sense Fe3+ but with an inferior performance (Table S1). Noticeably, spent tea-derived GQDs with the involvement of oxone oxidant in the synthesis process showed a lower LOD for Fe3+ detection rather than without the addition of acid-oxidant in the ethanol-assisted single-step ST synthesis (Table S1), which may be related to the structure and different surface states of GQDs under the two synthesis conditions.

6.1.2. Functionalized GQDs. The Fe3+ sensing results of GQDs/doped-GQDs functionalized with various chemical moieties are summarized in Tables 2 and S1. For instance, covalently modified RBD–N-GQDs (nitrogen content: 7.08%) were used to selectively detect Fe3+ via an unusual fluorescence enhancement process (Table 2). The greenish-yellow emission (λem: 520 nm) of GQDs was red-shifted to yellow luminescence (λem: 550 nm) after RBD functionalization. When Fe3+ was added to the system, a new/strong orange-red emission (λem: 580 nm) was observed due to the spirolactam ring opening of RBD, and consequently the QY increased to 43%. The 580 nm emission peak progressively increased to trace Fe3+ and the probe showed biomedical applicability (track Fe3+ in HeLa and pancreatic cancer stem cells (CSCs) via bright orange-red fluorescence); however, its narrow LR and interference from aluminium ions (Al3+) are some of its limitations.133 Subsequently, DA–GQDs were successfully applied for the nanomolar-level detection of Fe3+ due to the strong coordination ability of DA with Fe3+, followed by the oxidation of the catechol moiety to ortho-semiquinone. However, although DA–GQDs showed a low Fe3+ LOD (7.6 nM), the nanoprobe was irreversible in nature (could not be regenerated after the addition of EDTA) and only applicable in a lower detection range.222 A wide LR of 4–1800 µM in the fluorescence quenching-based quantification of Fe3+ from DPA-functionalized GQDs (DPA–GQDs, QY as high as 99.8%) was achieved, which could also quantify Fe3+ in iron supplement oral liquids (recovery: 98.2–102.5%) without measurable interference from Fe2+ (2000-fold concentration). The authors proposed that the functional groups on DPA–GQDs favour the formation of a stable octahedral complex with Fe3+ rather than unstable tetrahedral complexation with Fe2+.187 Later, naturally available high-softening point asphalt precursor-derived Am–GQDs (yellow-emissive) with abundant amide/amino groups exhibited a considerable sensing performance for Fe3+ with a very low LOD (0.51 nM), which is probably the lowest LOD reported thus far among the undoped/doped-/functionalized GQDs-based sensors for Fe3+ via the FL method (Tables 2 and S1). This probe also showed a reversible binding affinity with Fe3+ (released after adding EDTA) and promising applicability in real river water samples (LR: 0–90 µM and recovery: 95–105%).164

Cai et al.142 described the simple ultrasonication-assisted exfoliation of graphite flakes in NaOH/NMP solution to produce N-GQDs (QY: 19.12%), which were non-covalently passivated by hydrophilic saccharide (xylan) to improve their solubility/stability in aqueous medium and QY up to 36.63%. The resultant N-GQDs@xylan nanoprobe showed almost no interference from other cations/anions (Fig. 9a) and a satisfactory sensing performance for Fe3+ (LOD/LR: 92.8 nM/0–75 µM, Fig. 9b). The quenching mechanism in the detection of Fe3+ is ascribed to the combined effect of IFE (excitation-emission of N-GQDs@xylan overlaps with the absorption of Fe3+, Fig. 9c) and charge transfer (insignificant change in τ after the addition of Fe3+, Fig. 9d) between the fluorophore and analyte.


image file: d5ra04935k-f9.tif
Fig. 9 (a) Fluorescence quenching response of N-GQDs@xylan with Fe3+ without much effect from the presence of various ions. (b) Fluorescence spectra of N-GQDs@xylan in the presence of Fe3+ (0 to 150 µM), showing a gradual decrease in fluorescence intensity and a linear plot of 409 nm fluorescence intensity with respect to Fe3+ concentration (inset). (c) Excitation-emission spectra of N-GQDs@xylan, showing an overlap with the UV-visible absorption spectrum of Fe3+ rather than Fe2+. (d) Time-resolved fluorescence decay profiles of N-GQDs, N-GQDs@xylan, and N-GQDs@xylan in the presence of Fe3+. Reprinted (adapted) with permission from ref. 142, copyright 2023, the American Chemical Society. Linear calibration plot of the fluorescence intensity ratio (I/I0) vs. concentration of Fe3+ (e) and maximum quenching efficiency with Fe3+ among the tested MIs (f) using GQDs/CNCmod as a fluorophore. Reprinted (adapted) with permission from ref. 227, copyright 2023, the American Chemical Society.

Recently, Ye et al.223 employed dual-emissive (λem: 460/555 nm at λex: 370/480 nm) arginine (Arg) and Ser-functionalized B-GQDs (Arg,Ser–B-GQDs, QY: 40.12%) for the quantification of Fe3+ via the gradual weakening of their 555 nm yellow-emission peak. Interestingly, the LOD measured at 370 nm UV excitation is much higher (12[thin space (1/6-em)]400 nM) than the 480 nm visible excitation (75 nM), highlighting the good sensitivity of this probe in the visible light-induced sensing process. Alterations in the electronic structure of the probe after boron-doping narrowed its bandgap to improve its visible light absorption and dual functionality, further synergising its optical properties. Its relatively high selectivity (especially, against Fe2+) is ascribed to the strong coordination ability of Fe3+ with the oxygen/nitrogen-containing functional groups present on the surface of the probe to form a stable octahedral structure. Additionally, the constructed probe exhibited satisfactory reproducibility (1.7% RSD after 50 successive cycles), long-term stability (2.1% RSD after 6 weeks), and applicability in iron-fortified beverage samples (recoveries and RSDs: 99.4–100.8% and 1.2–2.5%).

6.1.3. GQDs involved with other counterparts. GQDs/modified-GQDs are effectively encountered with other counterparts to construct sensing platforms for Fe3+ (Tables 2 and S1). For instance, a core/shell hybrid of lignin sulfonate (LS)/GQDs (LS/GQDs) showed a satisfactory performance in the FL-based sensing of Fe3+ (Table 2). Here, the π-rich and sulphur-containing LS molecules on the GQDs favoured a 4-fold higher fluorescence intensity in comparison to bare GQDs, in addition to their chelating ability with Fe3+ to achieve good sensitivity/selectivity.224

GQDs or N-GQDs were non-covalently passivated with PVA and coated on a polyethylene terephthalate (PETP) film to fabricate a test paper-based convenient platform (stable and low cost) for the online detection of Fe3+ or Hg2+. The effective diffusion of HMIs in the fabricated kit showed a quick (<2 min) and real-time detection avenue; however, the kit is still not suitable for trace-level quantification. The visual fluorescence response of the test paper (simultaneously coated with N-GQDs/PVA (Fig. 10A(a)) and GQDs/PVA (Fig. 10A(b)) in real drinking water can be seen in Fig. 10B, which was separately or simultaneously quenched when 5 µM Fe3+ (Fig. 10C), 5 µM Hg2+ (Fig. 10D) or 5 µM Fe3+ and Hg2+ (Fig. 10E) was added, indicating the simultaneous and rapid detection of both HMIs in the real samples.225 The local optical field/edge functional groups in N-GQDs are enhanced/modified in the AuNPs@N-GQDs heterostructure to significantly improve its fluorescence intensity (∼12.1-times higher than N-GQDs). The interfacial and strong coupling of the plasmonic AuNPs with N-GQDs favoured an enhancement in electron density on N-GQDs to develop an approach for the fabrication of highly fluorescent nanoprobes. As a result, the heterostructure probe exhibited a low LOD of 1 nM for Fe3+ by applying the unconventional Langmuir adsorption law and non-radiative charge transfer dynamics in the entire detection range (0.001–10 µM), justifying the high sensitivity of the GQDs involved heterostructures (sensitivity of N-GQDs and N-GQDs/AuNPs mixture: 100 and 1000 nM, respectively). Their EDTA-triggered reversibility is also advantageous for multi-times sensing activity.226 Subsequently, a self-standing modified cellulose nanocrystal (CNCmod) thin film-hosted GQDs optochemical sensor (GQDs/CNCmod) was applied for the trace-level detection of Fe3+ (LR/LOD: 0.001–2/0.0008 pM, Fig. 9e), surpassing the Fe3+ detection limits by all other FL sensors (Tables 2 and S1). The coherent interaction between the sensing probe and Fe3+, resulting in the quenching phenomenon, can be ascribed to the high KSV values of 6.7 × 10−14/5.8 × 10−10 M−1 at Fe3+ concentrations of 0.001/2.0 pM. Moreover, the approximately double the fluorescence quenching with Fe3+ in comparison to other HMIs (Al3+, Cd2+, Co2+, and Cu2+, Fig. 9f) justified the appropriate selectivity of the sensor device. However, although this probe achieved a high level of sensitivity in the detection of Fe3+, its fabrication process is very specific and involves complicated steps.227


image file: d5ra04935k-f10.tif
Fig. 10 UV light illuminated digital photographs of N-GQDs (a) and GQDs (b) coated test-paper (A) and test-paper dipped in real drinking water (B). Different fluorescence quenching responses of test-paper dipped in drinking water containing 5 µM Fe3+ (C), 5 µM Hg2+ (D), and 5 µM Fe3+ and Hg2+ (E). Reproduced/adapted from ref. 225 with permission from The Royal Society of Chemistry, 2018.

Recently, Das et al.228 integrated machine learning (ML) with a solid-state photodetector to develop a suitable algorithm for the optimization of the Fe3+ sensing performance (experimental data for ML-based prediction is obtained from AuNPs@N-GQDs heterostructure). After optimization of the operating wavelength, the ML-trained model showed nearly 100% selectivity and nanomolar-level sensitivity (LOD: 1 nM) for Fe3+. Moreover, the constructed solid-state sensor could respond to Fe3+ in real-world samples (river water) at the lowest concentration of 10 nM. Apart from its robust stability (experimentally as well as ML-predicted), the strong affinity between Fe3+ and the probe facilitated the formation of Fe–O bonds and light-induced chare transfer for the current response (increasing/decreasing trends of dark/light current with an increase in the concentration of Fe3+) in the sensing operation, which was validated by the experimental and ML-based heatmap analyses. In another recent report, a GQDs-Au-Ni tubular micromotor exhibited a decreasing trend in its solid-state fluorescence intensity and speed (under a magnetic field) with an increase in the concentration of Fe3+ to quantify Fe3+ but with poor sensitivity (Table 2). Moreover, GQDs-Au-Ni@SPCE (SPCE: screen-printed carbon electrode) was also tested for the EC detection of Fe3+ (LOD: 6 µM).229

Summary: According to the above discussion, we can infer that turn-off based FL detection is common for Fe3+. The Fe3+ detection capability of N,S-GQDs is superior to that of single-heteroatom doped-GQDs and other dual-element doped alternatives. Among the single-hetero-element doped GQDs, N-GQDs are preferable probes due to their easy synthesis and good sensing performance; however, B-GQDs can also effectively sense Fe3+ via the FL method. Moreover, additional functional groups existing in functionalized GQDs (particularly, amide and amino) can selectively interact with Fe3+ and achieve a low LOD of 0.51 nM. The selectivity of GQDs/modified-GQDs with Fe3+ originates from its stable octahedral complexation with their functional groups. Non-covalent functionalization of N-GQDs with biocompatible xylan can also construct a good FL probe for Fe3+. Heteroatom-doping as well as incorporating multiple functional groups in GQDs can build a visible light-driven sensor for good Fe3+ sensing activity. Specifically, the incorporation of boron and covalent functionalization with multiple amino acids in GQDs can result in intense dual emission (blue and yellow) and satisfactory sensitivity/selectivity in the detection of Fe3+ using low energy visible light excitation. The excellent selectivity and high sensitivity of composite/heterostructure systems composed of GQDs/modified-GQDs involved cannot be ignored. For instance, LS/GQDs core–shell composites and AuNPs@N-GQDs heterostructures may be representative platforms for effective Fe3+ sensing. Furthermore, the new development of ML-based predication of their sensing metrics is noticeable and opens a new direction to identify contaminants with minimum experimental efforts.

6.2. Fe2+

Saenwong et al.230 first demonstrated the indirect speciation of Fe2+ in the presence of H2O2 via Fenton reaction to produce Fe3+ for the fluorescence quenching of GSH–GQDs (Table S1). Subsequently, the direct detection of Fe2+ was demonstrated using doped-/undoped/functionalized GQDs (Tables 2 and S1). For example, doped N,S,I-GQDs (nitrogen/sulfur/iodine content: 1.41/0.41/0.85%) were the first fluorescent nanoprobe that could directly determine Fe2+ via an AIQ-based mechanism (Table 2). However, EDTA and AgNO3 are used as a general and Fe3+-specific masking agent during the detection process, respectively.231 A purely biogenic (Mangifera indica leaves) and MW-assisted green approach yielded undoped GQDs (QY: 45%) for the fluorescence quenching-based detection of Fe2+ but with low sensitivity compared to previous results (Table 2).124

Recently, a synergistic effect of multi-functionality and dual-heteroatom doping in the FA,His,Ser–B,P-GQDs fluorophore resulted in an intense yellow-emission in aqueous solution (λem: 550 nm at 490 nm λex and QY: 60.2%) and significant yellow-emission even at a high concentration (6 mg mL−1) or in the solid state. The visible light-driven (490 nm excitation) sensing aptitude of Fe2+ with this nanoprobe in the presence of ortho-phenanthroline (Phen) showed a gradual decrease in fluorescence intensity (Fig. 11a), following a wide LR (0.01–50 µM, R2 = 0.991, Fig. 11b) and a low LOD of 4.2 nM. Based on the UV-visible and PL analyses (significant overlap between the absorbance of orange-red Fe–Phen complex and emission of fluorophore), it is proposed that efficient energy transfer between the generated complex (Fe–Phen) and fluorophore is responsible for fluorescence quenching. The specific complexation between Fe2+ and Phen rather than the other tested MIs and anions confirmed the good selectivity of this probe. Moreover, the detection of Fe2+ in human urine samples exhibited a good recovery in the range of 95.4–102.3%.166


image file: d5ra04935k-f11.tif
Fig. 11 Fluorescence spectra of FA,His,Ser–B,P-GQDs with the addition of 0 to 100 µM concentration of Fe2+ (a) and linear plot of Fp (final to initial intensity ratio) vs. Fe2+ concentration (0 to 50 µM) (b). Reproduced/adapted from ref. 166 with permission from The Royal Society of Chemistry, 2024.

Summary: The direct detection of Fe2+ in aqueous solution has recently witnessed new developments where the turn-off based FL detection of Fe2+ using multiple functional groups/dual-heteroatoms containing GQDs achieved the best performance.

6.3. Hg2+

6.3.1. Undoped and doped-GQDs. The application of a GQDs-based FL sensor in the sensing of Hg2+ was first demonstrated in early 2013 (Table S2). Subsequently, undoped GQDs synthesized from various precursors/conditions were further used for the purpose of Hg2+ recognition (Tables 3 and S2). For instance, a low LOD of 0.439 nM in the quenching-based detection of Hg2+ and analytical applicability in real water samples were achieved from CA precursor-derived GQDs, but the probe is limited to a lower concentration range (Table 3) and its selectivity was only tested with the same concentrations of different MIs/anions as Hg2+ in aqueous solution.232
Table 3 GQDs, modified-GQDs, and GQDs involved with other counterparts for Hg2+ sensing applicationa
GQDs-based sensor Synthesis conditions Size range/average sizeb (nm) QY (%) Sensing process LR (µM) LOD (µM) Ref.
a L-DOPA: 3,4-dihydroxy-L-phenylalanine, MSA: methanesulfonic acid, and EAA: ethyl acetoacetate.b Measured from TEM.c Size range/average size measured from dynamic light scattering.d Size range/average size of GQDs used with other counterparts.e Dynamic concentration range and corresponding LOD.f LR in ppm.g LR/LOD in µg L−1.h LR/LOD of paper-based sensor.i Analytical ability in real water/biological fluid samples.j Analytical ability in living cells.k Visual detection capability.
Undoped/doped-GQDs
GQDs Pyrolysis (CA, 200 °C, 30 min); dissolved in 10 mg per mL NaOH solution and pH adjusted to 8.0 7–11/— 15.4 FL, turn-off 0.001–0.05, 0.12–2 0.000439 232i
OH-rich GQDs Pyrolysis (CA, 200 °C, 25 min); mixed in 1% NaOH solution; centrifugation; dialysis 0.5–3/1.5 50 FL, turn-off 0–20 0.00987 233i,j
GQDs HT (Furfural derived CBDA-2 in NH4OH/H2O solution, 200 °C, 12 h); dialysis; centrifugation 4–7/— 45 FL, turn-off 10–100e 2.5e 235
Oxygen-rich N-GQDs Pyrolysis (CA/L-DOPA, 230 °C, 40 min); dissolved in water and pH adjusted to 7.0; dialysis 4–25/12.5 18 FL, turn-off 0.04–3 0.0086 236i
N-GQDs HT (nitrogen-doped GO in water, pH adjusted to 8.0, 200 °C, 12 h); filtration 3–6.4/— FL, ratiometric 0.002–0.2 0.00018 237j
COL 0.002–0.2 0.00032 237
N-GQDs@ITO IR-assisted pyrolysis (CA/urea, 250 °C, 10 min); dispersed in water; centrifugation; drop coated on ITO glass 3–6/4.5 EC, CV 0.05–0.25 0.05 117
N-GQDs Ar/DC microplasma treatment of chitosan in 50 mM CH3COOH electrolyte (pH: 4.44), 1 h; purification 4–9/6.39 30.1 FL, turn-off 0.5–60, 60–100 0.0479 163i
N-GQDs HT (Spent tea powder in water, 250 °C, 12 h); filtration; dialysis 0.9–2.5/1.6 22 FL, turn-off 0.1–0.5 0.004 149
N,S-GQDs IR-assisted pyrolysis (CA/urea/ammonia sulfate, 260 °C, 10 min); dispersed in water; centrifugation 1.5–6/3–5 25.5 FL, turn-off 0.01–10f 0.05 118
N,S-GQDs Ar/DC microplasma treatment of chitosan in 35 mM MSA or 0.1 M NH4OH aqueous solution, 1 h; purification 2.2–5.9/4.2 1.7 FL, turn-off 1–10, 10–40 0.0074 111
N-GQDs Ar/DC microplasma treatment of lignin in 35 mM MSA or 0.1 M NH4OH aqueous solution, 1 h; purification 2–4.8/3.1 1.0 " 1–20, 20–50 0.0685
N,S-GQDs Pyrolysis (CA/Cys, 160 °C, 5 min); dissolved in water; dialysis —/3.2 FL, turn-off 0.5–100g 0.33g 238i
0.1–10g,h 0.048g,h
B,N-GQDs Pyrolysis (CA/urea/H3BO3, 200 °C, 2 h); dissolved in water; centrifugation; filtration; dialysis 1.5–3.5/2 17.16 FL, turn-off 0–4 0.0043 239i
Mn,N,S-GQDs Acid oxidation of lignosulfonic acid sodium salt with HNO3 under ultrasonication (12 h); filtration; HT (obtained filtrate/2 wt% MnCl2 in water, 200 °C, 12 h); filtration; centrifugation 6–13/∼10 31.6 FL, turn-off 0.001–0.1, 0.2–1 0.00056 240i
[thin space (1/6-em)]
Functionalized GQDs
Val–GQDs Pyrolysis (CA/Val, 200 °C, 2.5 h); mixed in 1 M NaOH; pH adjusted to 7.0; dialysis 1–4/3 28.07 FL, turn-off 0.0008–1 0.0004 241i
FA–GQDs HT (maleic acid/FA in water, 180 °C, 2.5 h); dissolved in water and pH adjusted to 7.0; dialysis 2–8/5.2 FL, turn-off 0.000005–2 0.0000017 242i
PEHA,DPA–GQDs HT (CA/PEHA in water, 200 °C, 1.5 h); HT (obtained mixture/DPA, 200 °C, 2 h); diluted with water; dialysis 1–6/3.16 90.91 FL, turn-off 0.0001–200 0.000046 188i,j
DMC–GQDs@GCE Pyrolysis (CA, 200 °C, 30 min); dissolved in 10 mg per mL NaOH solution & pH adjusted to 7.0; covalently modified with DMC; electro-deposition on GCE 8–14/— EC, DPASV 1 × 10−6–15 × 10−6 0.26 × 10−6 103i
[thin space (1/6-em)]
GQDs involved with other counterparts
TH–GQDs HT (GO/TH in water, pH adjusted to 8.0, 180 °C, 12 h); filtration; dialysis 2–6/— 42 FL, turn-off 0.0005–0.05 0.00015 243i
GQDs/TH–ZnPc Non-covalent conjugation of GQDs with TH-ZnPc —/20c 9.0 FL, turn-off-on 0.0001–0.02 0.00005
TH–GQDs/TH–ZnPc Non-covalent conjugation of TH-GQDs with TH-ZnPc —/31c 3.0 " 0.005–0.05 0.0247
ZnNCs–N-GQDs/Au @GCE HT (TSC·2H2O/urea/Zn-DTT suspension in water, 160 °C, 8 h); solid washed with ethanol and dispersed in water; drop casted on AuNPs-coated GCE —/5d ECL, turn-off 0.00001–1000 3 × 10−6 244i,k
COL, turn-off 0.0001–100 33 × 10−6
Am–GQDs/PTH@ GCE Pyrolysis (CA, 175 °C, 30 min); mixed in aqueous ammonia solution; dialysis; deposited on GCE along with the electro-polymerization of thionine <10/5d EC, CV 1 × 10−6–1 0.6 × 10−6 245i,j
GQDs/Ce-ZnONFs@ GCE Liquid-phase exfoliation of GO/H2O in NaOH/EAA suspension under probe sonication (3 h); filtration; dialysis; in situ-immobilized in Ce-ZnONFs; drop-casted on GCE EC, DPV 0.1–100 0.267 246
GQDs/Gemini surfactant droplets Pyrolysis (CA, 150 °C, 12 min); mixed in 5 mg per mL NaOH solution; pH adjusted to 7.0; non-covalently conjugated with Gemini surfactant 2.75–4.75/4.3d FL, turn-off 0.1–0.5 0.0305 247i


Later, OH-rich GQDs, which were synthesized under similar conditions/precursors, showed a smaller size/high QY compared to previous GQDs (Table 3). As a result, this probe achieved the selective determination of Hg2+ in the presence of 500-/1000-times higher concentration of Fe3+/other interfering MIs with an extended LR and acceptable LOD (Table 3). The authors proposed that the quenching of the fluorescence of GQDs is due to their formation of a complex with Hg2+ and subsequent reduction of Hg2+ into Hg+ and metallic Hg via electron transfer from GQDs to Hg2+,233 which is validated in another report (Fig. 12) by spectroscopic, microscopic, and DPV results.234 The first application of biomass (Psidium guajava leaves)-derived red-fluorescent GQDs (maximum intensity λem at 673 nm) for the sensing of Hg2+ showed inferior sensitivity, as can be revealed in Table S2; however, the utility of biogenic precursors to achieve high wavelength-emissive GQDs (with the advantage of brightness and minimizing auto fluorescence in biological media) and applicability for HMI detection opens a sustainable and ecofriendly research direction. Subsequently, NIR-fluorescent GQDs (two emission peaks at λem: 440 and 850 nm by 310 nm λex) with a fairly high QY (45%) were synthesized from a biomass (furfural)-generated organic compound (cis-cyclobutane-1,2-dicarboxylic acid, CBDA-2, Table 3) for the quenching-based detection of Hg2+ with an acceptable performance (Table 3); however, selectivity is an issue with this probe (Fe3+, Fe2+, and Cu2+ are potential fluorescence quenchers at the higher concentration of 100 µM). Moreover, NIR-emitting GQDs are beneficial for biological application (tested for bioimaging) due to the minimum scattering of emissive light, small background effect, and high penetration capability in biological tissue.235


image file: d5ra04935k-f12.tif
Fig. 12 Schematic of the proposed mechanism, where Hg2+ is reduced to Hg+/Hg at pH 7.0 and Hg at pH 13 by coordination with GQDs. Reprinted from ref. 234, copyright 2020, with permission from Elsevier.

The relevance of nitrogen-doping in GQDs and use as an Hg2+ sensor started with an oxygen-rich N-GQDs FL probe, showing satisfactory sensitivity (Table 3); however, this probe required masking chemicals such as triethanolamine (TEtA) and sodium hexametaphosphate (SHMP) to circumvent the significant quenching arising from Pb2+/Cd2+/Cu2+/nickel ion (Ni2+)/Fe3+.236 Subsequently, various reports confirmed the applicability of N-GQDs in the field of Hg2+ sensing (Tables 3 and S2). For example, Peng et al.237 obtained low LODs (0.18/0.32 nM) for Hg2+ via FL/COL dual-mode sensing methods with linearities on the smaller concentration side (Table 3). The metalloporphyrin (MnIIITMPyP; TMPyP = 5,10,15,20-tetrakis(1-methyl-4-pyridinio)porphyrin) formation mechanism (accelerated by Hg2+ and N-GQDs) occurred via the ability of the large Hg2+ to deform the TMPyP nucleus, followed by the backside incorporation of small Mn2+ into it through N-GQDs as a carrier. As a result, the fluorescence signal of N-GQDs and TMPyP was enhanced and suppressed with IFE-based mechanism (ratiometric design), while original (422 nm)/red-shifted (462 nm) absorbance of TMPyP/MnIIITMPyP decreased/increased (basis for COL method), respectively. Moreover, this strategy could be successfully applied for the ratiometric monitoring of intracellular Hg2+ in A549 cells (human lung cancer cells). An N-GQDs-modified indium tin oxide (ITO) glass electrode was utilized for the EC detection of Hg2+ with a satisfactory result (Table 3); however, the coverage of intermediate Hg states and subsequent metallic Hg clustering on N-GQDs@ITO limited the sensitive detection of Hg2+ at low concentrations.117 The allowable LOD of 47.9 nM along with broad LR up to 100 µM in Hg2+ sensing by N-GQDs (synthesized from green precursor (chitosan) using Ar/DC plasma treatment) is noticeable (Table 3) but this probe is also sensitive to Cu2+, and thus requires a masking/chelating ligand to avoid its interferance.163 Recently, the gram-scale synthesis of N-GQDs (QY: 22%, production of 1.3 g in one batch) was shown to be possible via the HT carbonization of spent tea powder without involving an extra nitrogen source. The amino- and nitro-rich N-GQDs (nitrogen content: 8.1%) showed appropriate binding affinity with Hg2+ (soft acid-base interaction) to promote non-radiative processes and fluorescence suppression through DQE. Consequently, this biogenic platform achieved the trace-level FL detection of Hg2+ (LOD: 4 nM) but with the limitation of probing in a narrow concentration range (Table 3) and some perturbations from Pb2+/Cd2+.149

The Hg2+ sensing ability of dual-heteroatom doped-GQDs, such as N,S-GQDs, B,N-GQDs, and N,P-GQDs has also been reported in the literature (Tables 3 and S2). Among them, N,S-GQDs exhibited a sub-nanomolar level detection possibility in a low concentration window for Hg2+ (Table S2). The high sensitivity of N,S-GQDs (KSV: 0.22 l mg−1) rather than N-GQDs (KSV: 0.052 l mg−1) in the detection of Hg2+ is attributed to the presence of C–SOx–C sulphone bridges and other sulfur-doping configurations (C–S–C, C–SH), apart from the nitrogen/oxygen-containing functional groups in N,S-GQDs. As a result, the adjustment of the local electronic state, Fermi level, and creation of new energy states (from defect sites) in N,S-GQDs efficiently promoted their affinity towards Hg2+.118

Kurniawan et al.111 reported an energy-efficient DC microplasma-based method for the sustainable conversion of bio-resources (CA and saccharides such as fructose, chitosan, lignin, cellulose, and starch) into heteroatom doped-GQDs or undoped GQDs for the probing of environmental contaminants, including Hg2+, Cu2+, and 4-nitrophenol (4-NP) (Fig. 13a). Among them, chitosan-derived N,S-GQDs (nitrogen/sulfur content: 7.3/0.9%) showed notable sensitivity (LOD: 7.4 nM) in the turn-off based detection of Hg2+, while N-GQDs derived from lignin were found to be less sensitive for Hg2+ (Table 3). A recent report on N,S-GQDs (green-fluorescent) demonstrated that they not only selectively detected Hg2+ in the solution phase but also could be applied to construct a paper-based analytical device (PAD), and furthermore exhibited relatively high sensitivity (Table 3). The simple and cost effective construction of a biodegradable sensing device via the modification of PAD with polycyclic aromatic hydrocarbons (PAHs), followed the integration with N,S-GQDs and coordination interaction with Hg2+ is shown in Fig. 13b. The PAHs-modified PAD effectively adhered to the doped-GODs in well-dispersed manner to improve their sensitivity. This report also confirmed the PET-based quenching mechanism during the Hg2+ detection process and solution/paper-based detection potential in spiked-water/fermented fish samples.238 The incorporation of a small amount of boron (0.59%) in addition to nitrogen (9.61%) in B,N-GQDs also enabled the selective and sensitive recognition of Hg2+; however, it is not as good as the N,S-GQDs probe for Hg2+ (Tables 3 and S2).239


image file: d5ra04935k-f13.tif
Fig. 13 (a) Transformation of various bio-resources into heteroatom doped-GQDs through Ar microplasma treatment for the fluorescence quenching-based detection of environmental contaminants (Hg2+, Cu2+, and 4-NP). Reprinted (adapted) with permission from ref. 111, copyright 2022, the American Chemical Society. (b) Construction of paper-based device for the detection of Hg2+. Reprinted from ref. 238, copyright 2025, with permission from Elsevier.

It is observed that Mn2+-incorporated single-/dual-heteroatom doped-GQDs can also be selective for fluorescence quenching in the presence of Hg2+ (Tables 3 and S2). For example, the QY of in situ doped N,S-GQDs (23%, synthesized from lignosulfonate biomass, Table 3) further improved up to 31.6% by incorporating Mn2+-dopant (0.24 at%) in their structure. The collective effect of multi-element doping and abundant defect sites created new energy/edge states between π and π* of carbon in the Mn,N,S-GQDs (Fig. 14a) rather than N,S-GQDs (Fig. 14b) for the enhancement of their PL property. As a result, Mn,N,S-GQDs showed better applicability towards the sub-nanomolar sensitive detection of Hg2+ (LOD: 0.56 nM) in comparison to N,S-GQDs (LOD: 7 nM) and reusability after recovering their quenched PL with EDTA.240


image file: d5ra04935k-f14.tif
Fig. 14 Band structures of Mn,N,S-GQDs (a) and N,S-GQDs (b), showing various possible transitions during their excitation and emission processes. Reprinted (adapted) with permission from ref. 240, copyright 2020, the American Chemical Society.
6.3.2. Functionalized GQDs. Functionalized GQDs are also credited for the recognition of Hg2+ even at the sub-nanomolar level of sensitivity via the FL or EC method (Tables 3 and S2). For example, valine (Val)-functionalized GQDs (Val–GQDs) showed a 14-fold higher sensitive response (88.2% quenching) than bare GQDs (6.2% quenching) in the presence of 2 µM Hg2+ due to their functional groups (nitrogen- and oxygen-containing)-assisted strong interactions/complexation with Hg2+, and furthermore good sensitivity in the lower concentration side (Table 3).241 Li et al.242 designed FA–GQDs (FA groups covalently attached at the edges of GQDs; a large amount of graphitic and amine nitrogen compared to pyridinic nitrogen), which showed an LOD of 0.0017 nM again in the lower concentration range (LR: 0.005–2 µM), surpassing the LODs of all FL turn-off-based Hg2+ detection processes (Tables 3 and S2). Subsequently, dual-functionalized PEHA,DPA–GQDs (QY: 90.91%; amine groups dominated along with some sulfur element) were employed in the FL-based highly sensitive detection of Hg2+ (LR/LOD: 0.0001–200 µM/0.046 nM) without significant interference from other ions. The predominant nitrogen/oxygen-containing functionalities at the edges constitute a favourable configuration on the probe to effectively/selectively coordinate with Hg2+ and result fluorescence quenching.188

Hg2+ could also be sensed by the EC method using functionalized GQDs. For instance, the specific EC detection of trace-level Hg2+ with a DMC–GQDs-modified GCE electrode (DMC–GQDs@GCE) using differential pulse ASV (DPASV) measurement showed a minimum detection limit of 0.26 pM from the linear calibration plot. However, a high degree of sensitivity was achieved in this report via the pre-concentration and pre-reduction of Hg2+ at the active electrode surface. Here, the thiol functional groups in the DMC ligand and high surface area of the functionalized GQDs provided distinct and abundant complexation sites for Hg2+ during the EC operation. Besides its satisfactory repeatability, reproducibility, and stability, DMC–GQDs@GCE is also applicable for measuring the Hg2+ concentrations in tap and river water.103

6.3.3. GQDs involved with other counterparts. The construction of sensor probes via the utilization of GQDs/modified-GQDs with other functional moieties and their application for the sensing of Hg2+ have also been reported (Tables 3 and S2). For instance, a turn-off-on based sensing strategy was demonstrated for the low-level detection of Hg2+ using a conjugate system containing GQDs or thymine (TH)-functionalized GQDs (TH–GQDs) and TH-functionalized zinc phthalocyanin (ZnPc) (TH–ZnPc). Fluorescence quenching occurred due to the π–π interaction between GQDs or TH–GQDs and TH–ZnPc (through Förster resonance energy transfer (FRET)), which again were regenerated after the addition of Hg2+ due to the formation of TH–Hg2+–TH base pairs and interaction inhibition. It was observed that the non-covalent conjugation of bare GQDs with TH–ZnPc (GQDs/TH–ZnPc) possessed a significantly higher sensitivity (LOD: 0.05 nM) compared to TH–GQDs-conjugated TH–ZnPc (TH–GQDs/TH–ZnPc) or TH–GQDs, highlighting a turn-off-on approach for better sensitivity than the turn-off route (Table 3).243 The ECL detection of Hg2+ with high sensitivity is possible with systems containing GQDs, along with Hg2+-specific DNA aptamers and other components (Table S2); however, the construction of this type of platform is very complex, expensive, and requires precise experimental conditions.

Therefore, Wu et al.244 developed a simple zinc dithiothreitol (Zn-DTT) nanocrystals (NCs)-connected N-GQDs composite luminophore (ZnNCs–N-GQDs) for the sensitive ECL detection of Hg2+. The HT-based synthesis of ZnNCs–N-GQDs, their deposition over an Au-coated GCE, and ECL signal response are shown in Fig. 15a. The chelating ability of Hg2+ with the free S–H groups of ZnNCs turned the ECL signal off, which is validated by the decrease in absorbance after quenching (Fig. 15b). Consequently, this sensor showed an acceptable performance for the detection of Hg2+ (wide LR/low LOD: 0.01–1[thin space (1/6-em)]000[thin space (1/6-em)]000 nM/3 pM), which is much better than GQDs involved aptamer sensors (Table S2). This sensor probe also showed COL detection possibility for Hg2+ with reasonably good sensitivity (Table 3). Moreover, the visual detection capability (brown-coloured probe solution turned into a colourless supernatant with brown precipitate) and monitoring of Hg2+ in tap/lake water samples (recoveries: 96–105%) are practical attributes of this probe.


image file: d5ra04935k-f15.tif
Fig. 15 (a) Synthesis of ZnNCs–N-GQDs through HT method, and steps involved in the fabrication of ECL sensor along with the on/off signal response. (b) On/off mechanism in the absence/presence of Hg2+ and corresponding UV-visible spectra. Reproduced/adapted from ref. 244 with permission from The Royal Society of Chemistry, 2022.

The EC detection of Hg2+ with high sensitivity (LOD: 0.6 pM) in the low concentration range (LR: 1 pM–1 µM) using an Am–GQDs/poly(thionine) (Am–GQDs/PTH) nanocomposite is noticeable. However, the modified electrode is very specific to PTH deposition cycles and the EC process involved complicated steps, where the Cu-catalyzed Fenton-like reaction first increased the cathodic peak current. Then, Cu2+ is consumed by TU (CuTU2+ complex formation) to inhibit Fenton-like reaction/EC response, followed by the displacement of Cu+ from the complex in the presence of Hg2+ to regain the current response and readout to quantify Hg2+.245 Later, Qi et al.246 applied a GQDs/Ce–ZnONFs (NFs = nanofibers) hybrid as an electrode material for the DPV-based EC detection of Hg2+ in a wide LR of 0.1–100 µM (LOD: 267 nM). Although the redox process during the sensing operation is based on Ce/Zn, the functional groups of GQDs in the hybrid material improve the adsorption of Hg2+ on the electrode surface by creating sufficient oxygen vacancies/affinity and facilitate redox reactions for a selective and sensitive EC response.

Recently, a positively charged Gemini surfactant (zeta potential: + 55.9 mV) and negatively charged GQDs (zeta potential: −25.2 mV) were self-assembled in aqueous medium via electrostatic interaction to construct blue-fluorescent droplets (Fig. 16a). The droplet nanoprobe was found to be highly selective towards Hg2+ and exhibited a fluorescence quenching response on the progressive addition of Hg2+ with good sensitivities both in standard aqueous solution (LOD: 30.5 nM) and in spiked-tap water (LOD: 75.2 nM). The effective binding affinity between the luminescent droplets and Hg2+ was justified by the high KSV constant value (4.633 × 106 M−1), and furthermore by analyzing the confocal microscopic images, leading to diminished inherent blue fluorescence from the droplets (Fig. 16b) in the presence of 10 µM Hg2+ (Fig. 16c).247


image file: d5ra04935k-f16.tif
Fig. 16 (a) Schematic of the formation of blue-luminescent droplets via the self-assembly of GQDs and Gemini surfactant. Confocal microscopic images of droplets, showing blue fluorescence in the absence of Hg2+ (b) and quenched fluorescence in the presence of 10 µM Hg2+ (c). Reprinted (adapted) with permission from ref. 247, copyright 2025, the American Chemical Society.

Summary: The achievement of high selectivity/sensitivity with ratiometric design (N-GQDs/Hg2+-mediated metalloporphyrin formation), N-GQDs/N,S-GQDs (bio-precursor (chitosan) and energy-efficient microplasma derived), and Mn2+-synergised N,S-GQDs is highlighted in the execution of doped-GQDs for the identification of Hg2+. B,N-GQDs may also be potential candidates for the FL quenching-based sensitive detection of Hg2+. Moreover, the highly sensitive detection of Hg2+ using functionalized GQDs (specifically, PEHA and DPA dual-functionalized GQDs) via the FL method and achievement of picomolar-level sensitivity via the EC method (using DMC–GQDs) are noticeable. It is also observed that the compositing strategy of GQDs/modified-GQDs with other counterparts can be one of the suitable strategies by which Hg2+ can be detected at trace levels via ECL and EC approaches. Specifically, the picomolar-level ECL detection capability in a wide concentration range using a ZnNCs–N-GQDs modified electrode is recognizable.

6.4. Cu2+

6.4.1. Functionalized GQDs. The first report of Cu2+ sensing was achieved with functionalized GQDs (Am–GQDs; Am group introduced after the synthesis of GQDs, LOD: 6.9 nM, Table S3). Later, better sensitivity (LOD: 5.6 nM) was achieved for Cu2+ using Am–GQDs (Am group introduced during the synthesis of GQDs via the bottom-up approach, Table 4), indicating the advantage of incorporating Am groups during the growth stage of GQDs. However, the probing of Cu2+ by Am–GQDs followed linearity in a narrow concentration range.248
Table 4 GQDs, modified-GQDs, and GQDs involved with other counterparts for Cu2+ sensing applicationa
GQDs-based sensor Synthesis conditions Size range/average sizeb (nm) QY (%) Sensing process LR (µM) LOD (µM) Ref.
a LSV: linear sweep voltammetry.b Measured from TEM.c Size range/average size of GQDs used with other counterparts.d QY with respect to methylene blue dye.e LR/LOD in µg L−1.f Analytical ability in real water/vegetable/Thai recipe/serum samples.g Visual sensing capability.h Analytical ability in mouse cells/living cells/tumor cells.
Functionalized GQDs
Am–GQDs HT (glucose/NH3/H2O2 in water, 150 °C, 4 h); filtration; dialysis 1–7/4.34 32.8 FL, turn-off 0.01–0.1 0.0056 248
Am–GQDs Purchased from Suzhou Carbon-rich Graphene Technology Co. Ltd 10–35/21.3 FL, turn-off 0–80 1 249
[thin space (1/6-em)]
Undoped/doped-GQDs
sl-GQDs Acid oxidation of CNOs with 5 M HNO3 under reflux (95 °C, 4 h); pH adjusted to 7.0; dialysis; portion left in dialysis bag 2–6/3.1   FL, turn-off 0.02–0.2 0.02 250
GQDs@GCE Chemical oxidation of GO with 30 wt% H2O2/O3 under ultrasonication (3 h); reaction terminated by N2 purging (15 min); drop-casted on GCE 2–13/— EC, DPASV 50–650e 0.0003 251
GQDs/DNPC Pyrolysis (CA, 200 °C); mixed in 10 mg per mL NaOH solution and pH adjusted to 7.0 2.5–5.5/— FL, turn-off-on 0.01–10 0.0045 252f
GQDs Ar/DC microplasma treatment of starch in 0.1 M NaOH aqueous solution, 1 h; filtration 1.5–7/3.6 21.1 FL, turn-off 0.5–25 0.5 253f
GQDs Pyrolysis (CA, 200 °C, 30 min); mixed in NaOH solution and pH adjusted to 8.0; centrifugation —/2.2 55 FL, turn-off 0.01–0.5 0.0025 112
GQDs Pyrolysis (CA, 200 °C, 45 min); mixed in 10 mg per mL NaOH solution; filtered and pH adjusted to 7.0; aged for 3 days at 4 °C and mixed in ethanol; centrifugation 1.2–3.6/2.15 FL, turn-off 0.04–2 0.04 254f,g
N-GQDs Ar/DC microplasma treatment of chitosan in 35 mM HNO3 electrolyte (pH: 2.69), 1 h; purification 2.5–6.5/4.36 1.74 FL, turn-off 0.5–10, 10–100 0.1465 163f
N-GQDs/Paper Pyrolysis (CA/urea, 200 °C, 15 min); mixed in 10 mg per mL NaOH solution and pH adjusted to 7.0; drop-coated on paper strip 1.9–3.1/2.46 ECL, turn-on 0.01–1000 0.18 255f
N-GQDs ST (Styrofoam in acetone/NH3/H2O2 mixture, 225 °C, 1.5 h) 5–7/— 187d COL, turn-on 0–1 × 105 256f,g
N-GQDs HT (nitronaphthalene/p-aminobenzoic acid in water, 180 °C, 12 h); filtration 1–5/2.1 29.75 FL, turn-off 0–10 257h
[thin space (1/6-em)]
GQDs involved with other counterparts
Fe3O4@Chitosan–GQDs Pyrolysis (CA, 200 °C, 5 min); mixed in 0.25 M NaOH solution; immobilized on Fe3O4 NPs/Chitosan —/9c ICP-OES 0.05–1500e 0.015e 258f
Am–GQDs/SeNPs Pyrolysis (CA, 175 °C, 30 min); ultrasonically mixed in aqueous NH3; dialysis —/5c FL, turn-off-on 0.001–10 0.0004 259f,h
CdS/AuNPs/GQDs @ITO GQDs purchased from ACS material, USA; in situ incorporated in AuNPs and CdS nanorods; casted on ITO glass <5/—c PEC, LSV 0.0001–0.29 0.00227 260
GQDs//g-C3N4NSs/MWCNTs@GCE HT (CA/TU in water, 180 °C, 6 h); dispersed in water; centrifugation; dialysis; g-C3N4NSs and MWCNTs mixture is drop-coated on GCE ECL, ratiometric 0.0005–1 0.00037 261f
GQDs/TPPS (1[thin space (1/6-em)]:[thin space (1/6-em)]9) @SPCE MW (GA/triethylenetetramine in water, 300 W, 225 °C, 5 min); non-covalently modified with TPPS 0.5–6.5/—c EC, SWV 0–6, 6–13 0.172 262f


Recently, Ren et al.249 again applied Am–GQDs in the sensing of Cu2+, which showed a broad LR of 0–80 µM in the quenching process. Fig. 17a–c show the binding affinity of Cu2+ with –OH (populated with electron pair), 5σ molecular orbital of C[double bond, length as m-dash]O, and –NH2 (having electron pair) surface/edge groups of Am–GQDs, respectively. As a result, the photo-excited electrons in Am–GQDs are transferred to the empty 3d orbital of Cu2+ and inhibit radiative recombination for fluorescence quenching (Fig. 17d). However, this report did not test the quenching effect of the probe in the presence of other interfering ionic species and its applicability towards real sample analyses.


image file: d5ra04935k-f17.tif
Fig. 17 Coordination interactions of Cu2+ with hydroxyl (a), carbonyl (b), and amino (c) functionalities of Am–GQDs. (d) Schematic illustration of the electron transfer-based quenching process in the detection of Cu2+. Reprinted from ref. 249, copyright 2024, with permission from Elsevier. (e) Formation of DNPC and Cu2+-induced catalytic transformation of Cys to L-cystine for the inhibition of DNPC. (f) Fluorescence quenching of GQDs in the presence of DNPC. Reprinted from ref. 252, copyright 2019, with permission from Elsevier. (g) Schematic of the Cu2+ sensing strategy via ˙OH radical-induced disruption of the GQDs structure. Reprinted from ref. 254, copyright 2024, with permission from Elsevier.
6.4.2. Undoped- and doped-GQDs. The involvement of bare GQDs in the detection of Cu2+ can be revealed in Tables 4 and S3. For instance, blue-emitting sl-GQDs (remained in dialysis bag) with a high content of oxygenated defects are selective towards Cu2+ due to their structural, compositional, and energetic difference in comparison to multilayered GQDs (extracted from dialysis bag, low oxygen content, UV-emitting, and selective for Fe3+). The presence of more surface defects in sl-GQDs can effectively attract Cu2+, and furthermore the occurrence of charge transfer (from sl-GQDs to Cu2+) facilitated fluorescence quenching.250 Wen et al.251 first used GQDs as an electrode material in the EC detection of Cu2+ (LOD: 0.3 nM). The authors proposed that abundant –COOH groups on the GQDs are beneficial to create more negative sites for Cu2+ adsorption and redox reactions in the detection process. However, the EC response of GQDs@GCE with other interfering ions was not investigated. Ding et al.252 developed a turn-off-on based sensing platform for Cu2+. They observed that 2,4-dinitrophenylcysteine (DNPC, a reaction product of Cys and 1-chloro-2,4-dinitrobenzene (CDNB), Fig. 17e) effectively quenched the fluorescence signal of GQDs (Fig. 17f). Once Cu2+ is added, Cys is catalytically oxidized into L-cystine, and therefore the fluorescence recovery due to the suppression of DNPC formation (Fig. 17e) facilitated the identification of Cu2+ with satisfactory sensitivity (LOD: 4.5 nM) and practical applicability in real water samples. However, the sensing protocol functioned under specific conditions and is time consuming (Cu2+-catalyzed oxidation of Cys occurs at 90 °C, incubation time of Cys: 30 min, reaction time between Cys and CDNB: 50 min). The charge transfer-based fluorescence quenching of the GQDs in the presence of Cu2+ was revealed through DFT/time-dependent DFT studies. It is inferred that the photo-excited electrons from GQDs are mainly located on the Cu side, inhibiting the usual relaxation process, and therefore the forbidden-emission activity of the excited electrons is responsible for the quenching phenomenon. Additionally, the XPS and UV-visible results suggested the formation of a Cu–GQDs complex and charge transfer between them as the factors responsible for the quenching operation. It was also observed that the Cu2+ detection sensitivity of microplasma-synthesized GQDs is much better than that of HT-synthesized GQDs (precursor for both synthesis routes: starch), which is attributed to their high crystallinity (less defects) and facile electron transfer characteristics.253

Recently, Guo et al.254 realized the dual-mode detection of Cu2+ using undoped GQDs. Although the FL-based LOD of Cu2+ (40 nM) is larger than that in some previous reports (Tables 4 and S3), this probe could visually detect Cu2+ (minimum dose: 10 µM) under 365 nm light irradiation through the fluorescence quenching effect. The blue-emission of GQDs (EIPL characteristic, λem: 480 nm) is effectively suppressed by the ˙OH radical-driven (produced from the Fenton-like reaction between Cu2+ and ascorbate) disruption of the structure of GQDs (Fig. 17g) to monitor the Cu2+ concentration in the solution. However, this probe was found to behave irreversibly (cannot be reused/regenerated after the process) due to its structural disintegration, and also suffered from interference from Fe2+/Fe3+ (required masking agent: TEtA/SHMP). Finally, this probe is applicable to quantify Cu2+ in spiked-lake water samples with 84.4–108% recovery and visibility under UV light irradiation.

N-GQDs derived from different precursors and experimental processes are exclusively used as turn-off type FL, COL, and ECL probes to sense Cu2+ (Tables 4 and S3). For example, N-GQDs synthesized via the DC microplasma irradiation of a chitosan solution under atmospheric pressure showed a satisfactory performance in the detection of Cu2+ (Table 4).163 Zhu et al.255 demonstrated a paper-based ECL sensor using N-GQDs (fabricated via the screen-printing technique) to trace the switch-on ECL signal with an increase in the concentration of Cu2+ (Table 4). Nitrogen-doping in the N-GQDs facilitated electron transfer with the co-reactant to generate a stable/strong ECL response. Furthermore, electron/free radical transfer is accelerated after the addition of Cu2+ to enhance the ECL signal, and subsequently Cu2+ quantification.

In a recent observation, the utility of waste (Styrofoam) for the synthesis of N-GQDs and their further applicability for the selective quantification of Cu2+ by simply observing changes in their colour (blue to green among 20 tested MIs, Fig. 18A and B) without the assistance of UV light are noticeable. The variations in electronic states or surface plasmon arising from the N-GQDs after their interaction with Cu2+ are likely the reason for the sharp and visual colour changes. The analyses of the images in gray-mode by varying the Cu2+ concentration provided a calibration plot to quantify Cu2+ from unknown solutions, and the probe also showed its functionality in the analysis of real river water.256 Another recently synthesized N-GQDs probe with yellow-emissive characteristics (λex/λem: 494/540 nm) showed fluorescence quenching-driven detection capability for Cu2+ (Table 4). This probe is nearly recyclable with an EDTA chelating agent; however, the LOD was not reported in this study. Moreover, the applicability of this probe is extended to detect Cu2+ in spiked-pesticide/dye wastewater as well as in biological mouse cells.257


image file: d5ra04935k-f18.tif
Fig. 18 0.1 M solutions of 20 different metal salts (A) and the corresponding changes in their colour after the addition of N-GQDs (B). Reprinted (adapted) with permission from ref. 256, copyright 2024, the American Chemical Society.
6.4.3. GQDs involved with other counterparts. There are various mixture/composite/heterostructure systems in which GQDs/modified-GQDs play a crucial role to detect Cu2+ at different levels of sensitivity (Tables 4 and S3). For instance, the trace-level detection of Cu2+ in Thai food samples was demonstrated via traditional ICP-OES (LOD: 0.015 µg l−1) using a magnetic nanocomposite (Fe3O4@chitosan–GQDs) as an adsorbent material for Cu2+ pre-concentration. However, although the magnetic probe is reusable up to seven cycles, the tedious adsorption–desorption steps for Cu2+ before injecting into costly the ICP-OES setup limits this sensing as a facile approach.258 PET between Am–GQDs and SeNPs resulted in fluorescence quenching, which again turned-on in the presence of Cu2+–ascorbic acid (AA) to achieve a low LOD (0.4 nM). However, the optimal fluorescence quenching-recovery was obtained after mixing SeNPs, AA, and Cu2+ with Am–GQDs for 2.5 h, showing a time-consuming sensing process. Moreover, the sensor platform not only detected Cu2+ in tap and lake water samples (recoveries: 98.7–103%) but also in HeLa and cisplatin-resistant tumor cells (less uptake of Cu2+ in the resistant tumor cells compared to tumor (HeLa) cells due to the down regulation of human copper transporter-1 expression).259 The visible light-induced photoelectrochemical (PEC) detection of Cu2+ using a ternary CdS/AuNPs/GQDs@ITO electrode system is also beneficial for good sensitivity (Table 4). It is observed that the loading of GQDs in CdS/AuNPs is advantageous to significantly improve the photocurrent density of ternary composites (2.3-fold) due to their high charge mobility. Additionally, the favourable alignment of the energy levels between CdS/AuNPs and GQDs inhibited charge recombination to promote the PEC performance. However, although the probe is very selective for Cu2+ over other tested MIs, it exhibited an interference feature with Fe2+.260 After that, Liu et al.261 developed a dual-potential ratiometric ECL platform using GQDs and g-C3N4 nanosheets (g-C3N4NSs)/MWCNTs luminophores for the selective detection of Cu2+ with improved accuracy and sensitivity. It was observed that the strong cathodic ECL from g-C3N4NSs/MWCNTs@GCE was progressively suppressed in the presence of Cu2+, while anodic ECL from the solution-phase GQDs remained intact. As a result, the cathodic/anodic ECL intensity ratio responded for the very sensitive detection of Cu2+ (LOD: 0.37 nM) in comparison to the non-ratiometric single probe without GQDs (LOD: 45 nM), justifying the importance of GQDs to improve the sensitivity in the detection process. The low RSDs (3%) for the measurement of Cu2+ from 5 independent electrodes in the ratiometric design rather than the non-ratiometric approach (RSDs: 7.2%) further justified the positive effect of ratiometric sensing. However, this system suffered from Fe3+ interference, which could be overcome by the chelation strategy (SHMP). The satisfactory detection of Cu2+ in wastewater samples is also possible using the constructed sensor. The applicability of a composite system (GQDs/TPPS (1[thin space (1/6-em)]:[thin space (1/6-em)]9); TPPS = 5,10,15,20-tetrakis(4-sulfonatophenyl)porphyrin; 1[thin space (1/6-em)]:[thin space (1/6-em)]9 implies mass ratio of GQDs and TPPS) was demonstrated recently to detect Cu2+ via the EC method; however, its sensitivity was found to be inferior compared to previously reported EC detection results (Table 4). Here, the nitrogen of TPPS provided coordination sites for Cu2+ binding and its subsequent detection.262

Summary: We can say that Am functional groups on the surface of GQDs are advantageous for the selective/sensitive detection of Cu2+. Furthermore, good sensitivity for the detection of Cu2+ can be achieved by bare GQDs via EC and chemical reaction-driven FL quenching-recovery processes. Microplasma-generated crystalline GQDs and CA-derived GQDs (pyrolysis method) are also competitive probes for the FL detection of Cu2+, even at the visual level under UV light. Among the doped-GQDs, N-GQDs have shown potential to interact with Cu2+ and sense via the FL, COL or ECL method. Specifically, N-GQDs-deposited paper could selectively/sensitively probe Cu2+ via the ECL strategy. The valorisation of waste into N-GQDs and the quantification of Cu2+ (under normal light) by simply analyzing smart-phone based images is another noticeable attempt. The advantage of GQDs/functionalized GQDs for the fabrication of composites/mixtures with other counterparts and their employment in the sensitive detection of Cu2+ cannot be underestimated. Specifically, the selective and low-quantity detection of Cu2+ with GQDs involved systems through the PEC sensing strategy and ratiometric ECL manner are noticeable.

6.5. Pb2+

The first demonstration of the turn-on-based FL detection of Pb2+ is related to a conjugated system containing 3,9-dithia-6-monoazaundecane (DMA)-functionalized GQDs (DMA–GQDs; sulfur-containing moiety) and tryptophan, which exhibited an LOD of 9 pM in the lower concentration region (Table 5). The Pb2+-induced formation of a rigid structure between the two conjugated components favoured energy transfer interaction for the enhancement of the fluorescence intensity. Moreover, the probe is applicable to quantify the Pb2+ concentration in rat brains.263 Subsequently, the ECL and EC detection of Pb2+ were demonstrated using GQDs-based systems, showing 10-fold lower LOD in EC detection compared to ECL technique (Table S4).
Table 5 GQDs, modified-GQDs, and GQDs involved with other counterparts for Pb2+ sensing application
GQDs-based sensor Synthesis conditions Size range/average sizea (nm) QY (%) Sensing process LR (µM) LOD (µM) Ref.
a Measured from TEM.b Size range/average size of GSH-GQDs used with other counterparts.c QY of GSH–GQDs.d Analytical ability in rat brain microdialysate/biological fluid/real water samples.
DMA–GQDs/Tryptophan HT (GO/DMA in water, 200 °C, 24 h); filtration 1–3/1.8 FL, turn-on 0.00001–0.001 9 × 10−6 263d
GO/PDDA/G5/PDDA/GSH–GQDs@quartz SAMs Pyrolysis (CA/GSH, 240 °C, 10 min); dissolved in water; chromatography; layer-by-layer deposition of GO, PDDA, G-rich DNA and GSH–GQDs on quartz substrate 6–10/—b 33.6c FL, turn-off 0.0024–0.012 0.0022 264d
AuCuNCs/N-GQDs@ GCE HT (PANI/2 M NaOH in water, 220 °C, 12 h); centrifugation; used as reducing agent to synthesize CuNCs/N-GQDs; Galvanic exchange process to replace some surface CuNCs with Au using AuCl3, 65 °C, 4 h; drop-casted on GCE 3–5.5/— EC, DPV 1 × 10−6–10, 20–1000 1 × 10−6 265d
Hyb-BNQDs/N-GQDs@ GCE HT (PANI/bulk boron nitride in water, two drops of 2 M NaOH, 220 °C, 24 h); filtration 5–9.9/— EC, DPV 1 × 10−6–100 1 × 10−6 266d
CuNCLs@N,S-GQDs@ GCE HT (PANI in 0.05 M H2SO4 aqueous solution, 220 °C, 12 h); used during synthesis of CuNCLs from CuSO4/GSH; centrifugation; drop-casted on GCE 3–5.5/— EC, DPV 1 × 10−6–50, 20–1000 1 × 10−6 267d


The employability of doped-GQDs, GQDs/functionalized GQDs along with Pb2+-specific DNA (aptamer-based sensor) and functionalized GQDs for the sensing of Pb2+ can be disclosed in Tables 5 and S4. Among the doped-GQDs, the fluorescence signal of S-GQDs was selectively quenched in the presence of Pb2+ and used for its quantification with satisfactory sensitivity (Table S4). A self-assembled multi-layer (SAM) device was constructed on a quartz substrate using GO/GSH–GQDs as an energy acceptor/energy donor and poly(diallydimethylammonium) chloride (PDDA)/G-rich ssDNA strand (G5) as a linker. The shortening of the distance between GO and GSH–GQDs due to the formation of a G-quadruplex (folded form of G5 DNA) in the presence of Pb2+ resulted in an enhancement in FRET between the energy acceptor and donor for the fluorescence quenching-based detection of Pb2+ with high sensitivity (Table 5) and applicability in real blood samples. However, the construction of this sensing platform required a complex procedure and expensive reagents, along with extra precaution during its sensing activity.264

To avoid the complicated/expensive fabrication of DNA-involved sensor systems, recently AuCuNCs/N-GQDs@GCE with temporal stability greater than one year was explored for the picomolar sensitive EC detection of Pb2+ (LOD: 1 pM), which explicitly surpassed the previously reported sensing performances (Tables 5 and S4). The significant current response with AuCuNCs/N-GQDs@GCE is attributed to the spontaneous reduction of Pb2+, which is facilitated by Au+ (basic electrolyte partially oxidizes Au metal to Au+) and electron rich N-GQDs. The effective Pb2+–Au+ interaction is responsible for the Pb2+ selectivity by AuCuNCs/N-GQDs. The low band gap (1.32 eV) and small charge transfer resistance (0.6 kΩ) of this electrode further support its appropriate electrocatalytic activity. Apart from its satisfactory reusability (97% retention of its current response after 50 washing cycles) and reproducibility (1.72% RSD for five different electrodes), the EC platform is also validated for the analysis of Pb2+-spiked environmental samples with recoveries/RSDs of 99–100.8%/<0.5%. Meanwhile, the consumption of expensive gold salt for the synthesis of the electrode materials and their degradation/oxidation during the sensing operation cannot be ignored.265

In another recent report, a boron nitride (BN) QDs/N-GQDs hybrid system was used as an electrode modifier (Hyb-BNQDs/N-GQDs@GCE) and applied for the low-level detection of Pb2+. The N-GQDs in the hybrid configuration significantly improved the electrical conductivity, while BNQDs are advantageous for chemical inertness and overall stability.266 Unlike the previous EC electrode system, where the precious Au component of AuCuNCs facilitated the pre-reduction of Pb2+,265 here the electrode material functioned without the requirement of Au+-induced pre-reduction, and also excluded the use of costly chemicals in the synthesis process to achieve nearly equivalent sensitivity. Fig. 19a illustrates the fabrication of the electrode and its current response in the presence of Pb2+. The DPV responses of Hyb-BNQDs/N-GQDs@GCE according to the concentration of Pb2+ are shown in Fig. 19b, which exhibited an LOD of 1 pM (Fig. 19c) and linearly fitted in the dynamic concentration range of 1 × 10−6–100 µM (Fig. 19d). The chronoamperometric analyses (Fig. 19e) clearly depicted the selective response of the electrode material with Pb2+ and small non-interfering current from Fe2+. The sensing platform is well reusable (≥95% retention after 50 washing cycles), reproducible (≤4.8% RSD in the current response from five independent electrodes), and stable (∼97% retention in current after 180 days) for practical use. Moreover, this detection platform maintained its performance for the sensing of Pb2+ in wastewater samples (containing multiple interfering species) with good recoveries (>95%) and RSDs (≤5%).266 More recently, the same research group again reported the EC detection capability for Pb2+ using N,S-GQDs anchored with Cu nanoclusters (NCLs) (CuNCLs@N,S-GQDs) as an active electrode component. This EC probe not only featured picomolar-level sensitivity (Table 5) and good specificity (higher binding affinity for Pb2+ with thiol groups on the active material) but was also structurally stable, even after 365 days of storage (non-bonded and conjugated π electrons in the aromatic structure of N,S-GQDs stabilized CuNCLs). Its current response in the DPV curve did not require a pre-reduction step and Pb2+ was spontaneously/directly reduced to metallic Pb during the detection process due to the efficient electron-donating ability of N,S-GQDs. Additionally, this probe showed 98.67–99.80% recovery of Hg2+ in real water samples, appropriate reusability, reproducibility, and good temporal stability.267


image file: d5ra04935k-f19.tif
Fig. 19 (a) Schematic showing the modification of GCE with Hyb-BNQDs/N-GQDs and its application for the DPV signal-based EC sensing of Pb2+. (b) DPV curve with increasing current response according to the Pb2+ concentration (1 pM–1 mM). (c) DPV current response with 10−11 and 10−12 M Pb2+, showing an LOD of 1 pM. (d) Plot of current vs. Pb2+ concentration (10−12 to 10−4 M). (e) Chronoamperometric profile showing negligible interference in the presence of other MIs. Reprinted (adapted) with permission from ref. 266, copyright 2024, the American Chemical Society.

Summary: The applicability of GQDs/modified-GQDs to quantify Pb2+ via the FL, EC, and ECL methods is known. The presence of sulfur in GQDs can be favorable to specifically interact with Pb2+. Selectivity and satisfactory sensitivity in the FL detection of Pb2+ have been achieved by S-GQDs. GQDs involved hybrid/heterostructures are some of the appropriate choices to realize picomolar-level sensitivity and acceptable selectivity for the EC recognition of Pb2+. Specifically, hybrids of two QDs (Hyb-BNQDs/N-GQDs) and CuNCLs@N,S-GQDs heterostructures are suitable electrode modifiers for the EC detection of Pb2+ with perceptible performances.

6.6. Cr6+/Cr3+

N-GQDs were first employed for the label-free selective-sensitive detection of Cr6+ (Table 6) and simultaneous quantification of total Cr content (Cr6+ and Cr3+; Cr3+ is oxidized to Cr6+ by a chemical treatment protocol) in lake/river/domestic/industrial water samples. The functional groups in N-GQDs (nitrogen/oxygen-containing) showed stronger binding affinity and rapid chelating characteristic with Cr6+ in comparison to bare GQDs, indicating the suitability of the nitrogen element in the GQDs for Cr6+ sensing application.268 Subsequently, the ECL, FL, and EC detection of Cr6+/Cr3+ became possible using undoped/doped-GQDs, functionalized GQDs, and GQDs involved with other counterparts (Tables 6 and S5). For example, the IFE and SQE-based Cr6+ sensing performance of undoped GQDs via turn-off fashion with an LOD of 3.7 nM in a broad LR (0.05–500 µM) is noticeable. According to the FTIR analyses, the –OH, –COO, and C–H groups present on the GQDs interact with Cr6+, resulting in fluorescence quenching; nevertheless, this report did not confirm the composition of GQDs using other reliable techniques such as XPS and TEM-based elemental analysis.269 The rapid sensing of Cr6+ via the EC method (<1 min) using a GQDs/PANI composite-modified SPCE (GQDs/PANI@SPCE) exhibited satisfactory sensitivity (Table 6). This electrode was found to analyze more than 90 samples of Cr6+ per hour (each sample volume: 0.5 mL) and practical applicability in mineral water and deteriorated Cr-plating specimens. A report attempted to prepare the working electrode and Cr6+ sensing via automated mode, but the setup for online monitoring is very specific, not user friendly, and could not detect Cr6+ at the trace level. Moreover, Fe3+ is a potential interfering HMI (the reduction peak of Fe3+ effectively overlaps with Cr6+ even at the concentration ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]1) for the probe.270
Table 6 GQDs, modified-GQDs, and GQDs involved with other counterparts for Cr6+ sensing applicationa
GQDs-based sensor Synthesis conditions Size range/average sizeb (nm) QY (%) Sensing process LR (µM) LOD (µM) Ref.
a SFA-LSV: stopped-flow analysis with linear sweep voltammetry.b Measured from TEM.c Measured from AFM.d LR/LOD in µg mL−1.e LR/LOD of N,S-GQDs containing paper-based sensor.f LR/LOD of Cr6+ during direct addition.g LR/LOD of Cr6+ from the oxidation of Cr3+.h Analytical ability in real water samples.
N-GQDs HT (CA/NH3, 200 °C, 10 h); diluted with water and excess NH3 removed by heating (100 °C, 1 h) —/6.4 18.6 FL, turn-off 0–140 0.04 268h
GQDs GQDs purchased from XFNANO 0.5–2.5/1.2c 5 FL, turn-off 0.05–500 0.0037 269h
GQDs/PANI@ SPCE Pyrolysis (CA, 200 °C, 30 min); mixed in 10 mg per mL NaOH solution and pH adjusted to 4.0; diluted with water; loaded with aniline monomer onto SPCE; electro-polymerization EC, SFA-LSV 0.1–10d 0.097d 270h
sl-N-GQDs HT (Xylan/urea/NaOH in water, 240 °C, 24 h); centrifugation; dialysis —/3.2 23.8 FL, turn-off 3–150 0.43 121
sl-N-GQDs/PAAm hydrogel@MC In situ immobilized in PAAm hydrogel and integrated with MC " 3–75 0.1 121h
N,S-GQDs Pyrolysis (CA/TU, 180 °C, 30 min); dispersed in water; centrifugation; dialysis 1.6–5.7/2.8 22 FL, turn-off 1–100 0.01 120h
0.5–10e 0.4e
CQDs@GQDs CQDs derived from Houttuynia cordata extract via HT (180 °C, 4 h); pyrolysis (CA/CQDs, 220–240 °C, 5 min); mixed in 0.25 M NaOH solution 1–5/2.7 15 FL, turn-off 0.005–0.1f 0.01576f 271h
" 0.005–0.1g 0.00759g
N-GQDs HT (soluble starch/Arg in water, 190 °C, 4 h); centrifugation 1.4–3.4/2.4 10.9 FL, turn-off 0–50 0.8 272h


A self-passivated non-aromatic xylan layer on the surface of GQDs provided sl-N-GQDs (1.38% nitrogen) for application in the selective and sensitive sensing of Cr6+ (Table 6). The chelating effect between Cr6+ and nitrogen/oxygen functional groups of sl-N-GQDs destroyed the passivation boundary to cause fluorescence quenching, which was extended to a point-of-care device (portable microfluidic chip (MC)) for the smart-phone-based on-site/visual monitoring of Cr6+. An sl-N-GQDs-impregnated polyacrylamide (PAAm) hydrogel (in situ strengthened with cellulose nanofiber) was integrated with MC for the fabrication of a portable device and subsequent image-specific Cr6+ quantification. A significant decrease in the gray-scale brightness (100% to 71.52%) in 9 s by injecting 3 µM Cr6+ solution can be seen in Fig. 20a, which showed an acceptable LR/LOD of 3–75/0.1 µM (Fig. 20b) in the detection process. Additionally, the reliability of the portable platform was confirmed using real lake water-spiked samples, which showed satisfactory recoveries/RSDs (97–104/3.4–4.6%). However, due to its structural destruction during Cr6+ interaction, the probe lacks complete fluorescence recovery (even with excess AA, which reduces Cr6+ to Cr3+) and cannot be reused.121 Bezuneh et al.120 employed N,S-GQDs for the turn-off based FL sensing of Cr6+, which exhibited better sensitivity compared to N-GQDs (Table 6). The solution-phase sensing capability of N,S-GQDs (LOD: 10 nM) was extended to a paper-based device by evaluating smart-phone-captured changes in the blue values (B0B) under UV light according to the Cr6+ concentration (0–200 µM, Fig. 20c), which is a simplified device attempt towards the on-site detection of Cr6+ in real tap/river water samples. Nevertheless, its detection sensitivity is inferior to previously reported MC-hydrogel devices (Table 6) and requires a longer incubation time (60 min) for acquiring a detectable blue value. In an observation, it was found that the LOD of Cr6+ (generated from the H2O2-induced oxidation of Cr3+) is lower (7.59 nM) than directly added Cr6+ (15.76 nM) in the low concentration domain when a CQDs@GQDs nanohybrid is used as a probe, showing the estimation capability of total Cr content in the water samples. This probe is applicable to achieving satisfactory Cr6+ recoveries in real water samples but the reason behind its improved sensitivity in the case of indirect Cr6+ detection is unclear.271


image file: d5ra04935k-f20.tif
Fig. 20 (a) Average gray-scale brightness profile after incorporating 3 µM Cr6+ solution into MC, showing a decreasing trend with time. (b) Decrease in gray-scale brightness (after 9 s) with different concentrations of Cr6+, showing linear fitting from 3 to 75 µM. Reprinted from ref. 121, copyright 2021, with permission from Elsevier. (c) Plot of the changes in blue value with respect to various concentrations of Cr6+ (0–200 µM), along with the corresponding fluorescence images under UV irradiation (inset). Reprinted (adapted) with permission from ref. 120, copyright 2023, the American Chemical Society. (d) Excitation-emission spectra of N-GQDs, showing overlap with the absorption spectrum of Cr6+. (e) Bar diagram of zeta potentials for N-GQDs, Cr6+, and N-GQDs in the presence of Cr6+. (f) PL decay curves showing τ value of N-GQDs without and with different concentrations of Cr6+. Reproduced/adapted from ref. 272 with permission from The Royal Society of Chemistry, 2024.

Recently, Ni et al.272 employed HT-synthesized N-GQDs (using bio-precursor: starch as a carbon source and Arg as a nitrogen source) with a high nitrogen content (11.98%) for the detection of Cr6+ but with an inferior performance compared to previously reported N-GQDs or N,S-GQDs (Tables 6 and S5). Based on UV-visible (overlap of Cr6+ absorbance with the excitation/emission spectra of N-GQDs, Fig. 20d), zeta potential (negative value similar to N-GQDs after the incorporation of Cr6+, Fig. 20e), and time-resolved spectroscopy (insignificant change in τ after Cr6+ addition, Fig. 20f) results, the quenching of fluorescence is attributed to the IFE and SQE mechanism. Moreover, the testing of Cr6+ in actual water (tap/bottled drinking/lake) samples showed satisfactory recoveries (92.6–103.3%) with RSDs less than 4.5%.

Summary: IFE-driven fluorescence quenching in the detection of Cr6+ (using GQDs-based systems) is the most common approach. The considerable selectivity and sensitivity achieved by N-GQDs/N,S-GQDs with Cr6+ HMI are some notable results. Furthermore, the extension of doped-GQDs to the fabrication of hydrogel-based MC devices and paper-based convenient devices is demonstrated, which served as sensing platforms (on-site level) for Cr6+ through smart-phone image analyses. The applicability of GQDs/PANI as electrode materials for the rapid and continuous (90 samples/h) detection of Cr6+ opens a new avenue to construct GQDs involved systems for the EC detection of Cr6+.

6.7. Cd2+

An earlier report showed ECL-based Cd2+ sensing using an N-GQDs/K2S2O8 co-reactant system with a good LOD (13 nM) in the lower concentration domain. The coordination of Cd2+ with the functional groups of N-GQDs caused ECL quenching, which was further recovered by the addition of EDTA.195 The subsequent development of GQDs-based systems and their performance in the detection of Cd2+ are compiled in Table 7.
Table 7 GQDs, modified-GQDs, and GQDs involved with other counterparts for Cd2+ sensing application
GQDs-based sensor Synthesis conditions Size range/average size (nm)a QY (%) Sensing process LR (µM) LOD (µM) Ref.
a Measured from TEM.b Size range/average size of GQDs/N-GQDs used in the composite/aerogel.c Absolute QY.d LR/LOD in µg L−1.e LR/LOD for N-GQDs containing paper-based sensor.f LOD in ppb without pre-reduction.g LOD in ppb with pre-reduction.h Analytical ability in real water/herbal medicine samples.
N-GQDs Acid oxidation of GO with HNO3:H2SO4 (4[thin space (1/6-em)]:[thin space (1/6-em)]1) under MW-reflux (240 W, 100 °C, 3 h); pH adjusted to 8.0; filtration; dialysis 2–7/4.5 11.7 ECL, turn-off 0.02–0.15 0.013 195
N-GQDs/TMPyP HT (nitrogen-doped GO in water, pH adjusted to 8.0, 200 °C, 12 h); filtration 3–6.4/— FL, turn-off 0.5–8 0.088 273h
COL 0.1–10 0.09
N-GQDs@GCE HT (PANI/2 M NaOH in water, 220 °C, 12 h); centrifugation; drop-casted on GCE 1.5–3.5/∼2.3 EC, DPV 1 × 10−5–100, 200–1000 1 × 10−5,f 1 × 10−7g 197h
N-GQDs HT (CA/ethylenediamine in water, 180 °C, 4 h); centrifugation; dialysis 0.5–8/3.2 80c FL, turn-on 1–25d 1.09d 274h
" 1–15d,e 0.59d,e
GQDs/TPPS (1[thin space (1/6-em)]:[thin space (1/6-em)]6) @SPCE MW (GA/triethylenetetramine in water, 300 W, 225 °C, 5 min); non-covalently modified with TPPS 0.5–6.5/—b EC, SWV 0–8, 8–13 0.436 262h
T–N-GQDs–CAA HT (CA/urea in water, 160 °C, 4 h); dialysis; covalently modified with sodium alginate; added TMPyP and freeze dried 3.5–7.5/—b COL, color change from red to green 10–2500d 5.10d 275h


For instance, owing to the enhanced conductivity and electrocatalytic activity of PANI-derived N-GQDs (∼10% nitrogen content, low bandgap), they showed high selectivity towards the EC detection of Cd2+ with a low LOD of 1 × 10−5 ppb (∼8.9 × 10−5 nM; without pre-reduction, Fig. 21a). The non-bonding electrons of the nitrogen atom and aromatic π moiety could spontaneously reduce Cd2+ into metallic Cd on the surface of N-GQDs for an EC response (Fig. 21c). Furthermore, by applying a pre-reduction step, N-GQDs@GCE exhibited an LOD as low as 1 × 10−7 ppb (∼8.9 × 10−7 nM; Fig. 21b), which is one of the best EC performances simply using N-GQDs without other components. The presence of nitrogen within the aromatic rings and as functional groups is crucial for the selective interaction of N-GQDs with Cd2+ rather than other HMIs. Moreover, the reusability/reproducibility/stability and applicability of the constructed sensor for Cd2+-spiked environmental samples (ground/sea/waste water) are quite satisfactory.197 Subsequently, the complexation of Cd2+ with the functional groups of N-GQDs (preferably –NH2) resulted in an unusual increment in fluorescence due to the inhibition of the PET process, and therefore followed the chelation-enhanced fluorescence (CHEF) mechanism. Consequently, the N-GQDs (absolute QY: 80%) showed considerable LODs of 1.09/0.59 µg l−1 in the solution-phase/paper-based detection process. Meanwhile, the solution/paper-based sensor systems also showed applicability in Cd2+-spiked real water and herbal medicine samples; however, the paper-based sensor required a large N-GQDs loading (3.4 mg mL−1) to achieve reasonable selectivity and the sensing platform was restricted to determining higher concentrations of Cd2+.274


image file: d5ra04935k-f21.tif
Fig. 21 (a and b) DPV-based current response of N-GQDs@GCE in the presence of different concentrations of Cd2+ without and with the pre-reduction step, respectively. (c) Schematic of the deposition of N-GQDs on GCE, resulting in a 25-fold larger current response than bare GCE. Reproduced/adapted from ref. 197 with permission from The Royal Society of Chemistry, 2021.

Recently, the EC detection of Cd2+ using a GQDs/TPPS (1[thin space (1/6-em)]:[thin space (1/6-em)]6) (1[thin space (1/6-em)]:[thin space (1/6-em)]6 implies mass ratio of GQDs to TPPS) electrode material showed inferior sensitivity compared to previous EC results (Table 7). Additionally, the current response of Cd2+ with this electrode system was significantly reduced in the presence of Cu2+, showing a considerable interference issue.262 In another recent report, Tang et al.275 employed the EDC/NHS coupling reaction between N-GQDs and sodium alginate, followed by the addition of TMPyP and freeze drying to obtain a T–N-GQDs–CAA aerogel (Fig. 22a and Table 7) for the rapid (∼4 min) detection of Cd2+ via the COL method. The electrostatic and π–π interactions originating from the N-GQDs facilitated the effective self-assembly of TMPyP in the aerogel and formation of ion channels for the favourable and rapid transportation of Cd2+ during the sensing operation. Due to the strong binding affinity of Cd2+ with TMPyP rather than the functional groups of N-GQDs, the aerogel pellet could trace an increasing concentration of Cd2+ via the change in its colour from red to green (Fig. 22b) and showed a broad LR of 10–2500 µg l−1 (Fig. 22c), along with an acceptable LOD (5.10 µg l−1). The developed system was also validated for its practical utility to detect Cd2+ in tap (recoveries: 99.33–102%)/river (recoveries: 100–103%)/lake (recoveries: 102.4–104%) water with RSDs ranging from 1.08% to 4.32%.


image file: d5ra04935k-f22.tif
Fig. 22 (a) Synthetic steps involved in the preparation of T–N-GQDs–CAA aerogel. (b) Changes in the colour of aerogel pellet with 0–2.5 mg per l Cd2+ and corresponding red (R)/green (G)/blue (B) values captured using a smart-phone. (c) Linear fitted plot between R/B ratio and concentration of Cd2+. Reprinted from ref. 275, copyright 2025, with permission from Elsevier.

Summary: According to the research developments in the GQDs-based detection of Cd2+, N-GQDs are the most suitable choice among doped-GQDs. The low-level detection of Cd2+ using the N-GQDs electrode material via the EC method and the applicability of N-GQDs/TMPyP-containing aerogel in the COL detection of Cd2+ via smart-phone-based simple image analyses are notable results.

6.8. Co2+

The doped-GQDs encountered in the detection of Co2+ are summarized in Tables 8 and S6. For instance, Wang et al.276 demonstrated an AIE quenching (AIEQ) pathway in the selective and sensitive FL detection of Co2+ using N-GQDs with a low LOD (2 nM). The high affinity of Co2+ with amino functional groups present at the edges of N-GQDs induced their aggregation, and consequently fluorescence quenching. This nanoprobe was also extended to the intracellular monitoring and in vitro tracing of Co2+-induced cellular apoptosis in A549 living cells.
Table 8 GQDs, modified-GQDs, and GQDs involved with other counterparts for Co2+, Ni2+, Al3+, and As3+ sensing application
GQDs-based sensor Synthesis conditions Size range/average sizea (nm) QY (%) Sensing process LR (µM) LOD (µM) Ref.
a Measured from TEM.b QY of GQDs.c LR/LOD in ppb.d Analytical ability in living cells.e Analytical ability in real water samples.f Paper-based sensing capability.
Co2+
N-GQDs MW (PEI/L-lysine in water, 400 W, 120 °C, 5 min); dissolved in water and pH adjusted to 7.0; dialysis 4–6/5.2 FL, turn-off 0.01–5 0.002 276d
N,S-GQDs HT (CA/cysteamine·HCl in water, 160 °C, 4 h); centrifugation 1.1–5.4/3 FL, turn-off 0–40 1.25 277e,f
[thin space (1/6-em)]
Ni2+
EDA–GQDs HT (GO/6 wt% H2O2/W18O49 nanowires in water, 200 °C, 96 h); covalently functionalized with EDA under HT (150 °C, 24 h) 2–6/4.2 83 FL, turn-off 0.1–50 0.03 278d
[thin space (1/6-em)]
Al3+
GQDs MW (glucose/ethylene glycol in water, 800 W, 9 min); filtration; dialysis 1.5–7/∼3.3 2.5 FL, turn-on 0.4–500 0.0598 279d
[thin space (1/6-em)]
As3+
GQDs/GSH–rCQDs GQDs purchased from Sigma-Aldrich; mixed with GSH–rCQDs 18.9b FL, ratiometric 0.5–100c 0.5c 280f


Subsequently, the metal–ligand interaction between Co2+ and functional groups of N,S-GQDs (carboxyl, amino, and thiol) caused the aggregation of N,S-GQDs, and consequently the weakening of their fluorescence intensity (Fig. 23a). Moreover, N,S-GQDs showed potential to detect Co2+ in real water specimens (recoveries/RSDs: 91.2–108.2/0.1–7.3%) and construction of paper-based strip towards the visual monitoring of Co2+ (Fig. 23b). However, although this probe responded to Co2+ selectively in a broad LR (solution-phase), the calculated LOD (1.25 µM) is significantly higher than the normal level of Co2+ in human blood/urine (0.003/0.017 µM) and toxic level of 0.085 µM.277


image file: d5ra04935k-f23.tif
Fig. 23 (a) Schematic showing the interaction of Co2+ with the functional groups of N,S-GQDs for the weakening of fluorescence. (b) N,S-GQDs-coated paper strips for the visual monitoring of Co2+ (0–1000 µM) under a 365 nm light exposure. Reproduced/adapted from ref. 277 with permission from The Royal Society of Chemistry, 2020.

Summary: Based on reports, the aggregation effect of N-GQDs or N,S-GQDs in the presence of Co2+ can result in fluorescence quenching, and furthermore quantify Co2+ via the FL method. The presence of functional groups (nitrogen- and oxygen-containing) on the surface/edge of GQDs is essential for effective interactions/complexation with Co2+.

6.9. Ni2+

GQDs-based sensors have also been used in the detection of Ni2+ (Tables 8 and S6). For example, Xu et al.278 used bright yellow-fluorescent ethylenediamine (EDA)-functionalized GQDs (EDA–GQDs, QY: 83%) to sense Ni2+ with acceptable sensitivity (Table 8). Due to the strong coordination ability of EDA with Ni2+, the fluorescence signal of EDA–GQDs is quenched after the addition of Ni2+ via the SQE mechanism (no change in the average lifetime (τav) before and after the introduction of Ni2+). Moreover, the in vitro detection of Ni2+ in rADSC cells was confirmed by the suppressed fluorescence after internalizing biocompatible EDA–GQDs and subsequent incubation with 5 µM Ni2+.

Summary: There are few reports on GQDs-based sensors for Ni2+. The results inferred that Ni2+-specific functional group (such as EDA)-containing GQDs can be a suitable platform for the FL quenching-based detection of Ni2+.

6.10. Al3+

The reasonable sensing performance of undoped GQDs in comparison to doped-GQDs in the fluorescence turn-on-based detection of Al3+ can be confirmed in Tables 8 and S6. For instance, Yao et al.279 employed a time-saving MW-assisted green protocol (800 W, 9 min) to synthesize GQDs (precursor: glucose) for the aggregation-induced enhanced emission (AIEE)-based detection of Al3+ (Fig. 24a). Although the QY of GQDs is quite low (2.5%), they could potentially detect Al3+ with a wide LR (Table 8) and LOD of 59.8 nM. Here, AIEE caused an increase in the fluorescence signal in the presence of Al3+ due to the restriction of GQDs to intra-molecularly rotate in the aggregated state (via coordination and electrostatic interaction), and therefore slowed down non-radiative decay. Moreover, satisfactory recoveries/RSDs of 96.8–109.7/4.72–8.87% in real complex samples (tab/drinking/pond/river water) validated the practical capability of the nanoprobe system.
image file: d5ra04935k-f24.tif
Fig. 24 (a) Synthesis of GQDs via an MW-assisted method, and their analytical applicability for Al3+ detection via AIEE mechanism. Reproduced/adapted from ref. 279 with permission from The Royal Society of Chemistry, 2020. (b) Fluorescence spectra of GQDs/GSH–rCQDs with 0.5 to 100 ppb concentration of As3+. (c) Linear relationship between the I410/I550 intensity ratio and concentration of As3+. (d) Colour variations of GQDs/GSH–rCQDs-containing paper strips with As3+ (5 to 100 ppb) under UV light. Reproduced/adapted from ref. 280 with permission from The Royal Society of Chemistry, 2019.

Summary: The FL detection of Al3+ involved the fluorescence enhancement phenomenon, which is commonly caused by the creation of aggregated-state GQDs/doped-GQDs in the presence of Al3+. Undoped GQDs have shown high selectivity as well as sensitivity compared to doped-GQDs (B-GQDs and N-GQDs) in the turn-on type FL identification of Al3+.

6.11. As3+

The successful detection of toxic As3+ could also be possible with systems containing undoped GQDs (Tables 8 and S6). The AIEE effect in magnetic GQDs (Fe3O4NPs–GQDs; synergised by the presence of Fe3O4NPs and –OH groups) after the addition of As3+ was used to sense As3+ (satisfactory performance) in artificial and real water solutions via a turn-on manner (Table S6). Subsequently, a completely carbon-based ratiometric sensor was constructed (combining violet-emitting GSH-functionalized reduced CQDs (GSH–rCQDs) and green-emitting GQDs) for detecting As3+, which relied on the –SH group-mediated preferential interaction and subsequent quenching of violet-emission (410 nm) over green ones (550 nm) (Fig. 24b). Consequently, the dual-emissive nanohybrid probe enabled the dose-dependent sensitive quantification of As3+ in a wide LR of 0.5–100 ppb (Fig. 24c) and LOD of 0.5 ppb. Additionally, the GQDs/GSH–rCQDs-coated nitrocellulose test paper could visually detect As3+ by colour variations (blue to cyan to green, Fig. 24d) under UV illumination.280

Summary: The advantage of ratiometric probes (nanohybrids containing two carbon-based QDs) is obvious, which not only enabled the highly sensitive FL detection of As3+ but also a visual detection possibility by the variation in colour under UV light (semi-quantitative analysis).

6.12. Ag+

By taking the advantage of the formation of AgNPs (within 5 min) via the reduction of Ag+ on the surface of GQDs, the first FL detection of Ag+ using GQDs was realized in 2013.281 Subsequently, functionalized GQDs, doped-GQDs, bare GQDs, and systems containing doped-GQDs were successfully employed in the detection of Ag+ (Tables 9 and S7). For example, biomass (Passiflora edulis sims)-derived N-GQDs (4.6% nitrogen) showed the selective quenching of the fluorescence intensity in the presence of Ag+ with high sensitivity (Table 9). Based on the UV-visible (280 nm absorbance increased with the appearance of new/strong absorption at 220 nm after adding Ag+) and PL decay lifetime (almost unchanged τav before/after the addition of Ag+) analyses, it is proposed that the combined effect of SQE and electron transfer is accountable for the turn-off based detection process; however, deep insights into the selectivity and high sensitivity are lacking.282
Table 9 GQDs, modified-GQDs, and GQDs involved with other counterparts for Ag+ and Au3+ sensing application
GQDs-based sensor Synthesis conditions Size range/average sizea (nm) QY (%) Sensing process LR (µM) LOD (µM) Ref.
a Measured from TEM.b Size range/average size of N,S-GQDs.c Size range/average size of GQDs measured from dynamic light scattering.d Analytical ability in real water/carbonated drink samples.e Visual detection capability.
Ag+
GQDs Acid oxidation of GO with HNO3:H2SO4 (4[thin space (1/6-em)]:[thin space (1/6-em)]1) under MW-reflux (650 W, 100 °C, 8–10 h); pH adjusted to 8.0; filtration; reduced with NaBH4 (room temperature, 10 h); dialysis 5–6.2/5.5 FL, turn-off 0–0.1 0.0035 281
N-GQDs HT (Passiflora edulis sims extract, 180 °C, 4 h); filtration 2–6/3.8 29 FL, turn-off 0.01–160 0.0012 282d
N,S-GQDs@PtNCLs HT (CA/TU in water, 160 °C, 4 h); dialysis; decorated with PtNCLs 18–28/21.53; 1.5–2.7/2.17b COL, turn-off 0.0005–0.3 0.0002 283d
N,S-GQDs/CdTeQDs Pyrolysis (CA/GSH, 200 °C, 5 min); mixed in 0.3 M NaOH solution; drying; mixed with CdTeQDs <5/—b FL, ratiometric 0.00117–0.00588, 0.0472–0.118, 1.7–4.2 0.000226, 0.004679, 0.143 119d,e
[thin space (1/6-em)]
Au3+
N,S-GQDs HT (CA/Cys in water, 200 °C, 8 h); centrifugation; dialysis 1–3.5/2.1 35.4 FL, turn-off 0.1–50 0.05 284d
GQDs@CFP@PM Pyrolysis (glucose, 200 °C, 20 min); dissolved in water; adsorbed on CFP; coated with PM solution 10–40/26.8c COL, turn-off 200–1000 70 114d,e


Xue et al.283 demonstrated the sensitive detection of Ag+ via the COL method (LOD: 0.2 nM) using N,S-GQDs-decorated Pt NCLs (N,S-GQDs@PtNCLs). This probe catalyzed the oxidation of 3,3′,5,5′-tetramethylbenzidine (TMB) with H2O2 (peroxidase-activity, active species: ˙OH radical) to develop blue colour (652 nm absorption from oxidized TMB (TMBox)), which was subsequently suppressed by the coverage of the probe surface with Ag metal (generated by the N,S-GQDs-mediated reduction of Ag+) (Fig. 25a). Meanwhile, the use of expensive Pt-salt for the generation of PtNCLs in the active probe cannot be neglected (N,S-GQDs did not show peroxidise-activity) and this probe is only applicable within the low concentrations of Ag+. Subsequently, a mixture of N,S-GQDs (nitrogen/sulfur content: 7.25/2.64%) and CdTeQDs was used to assemble a ratiometric platform (effective quenching of the 570 nm emission from CdTeQDs (93.31%) rather than the 424 nm emission from N,S-GQDs (10.76%)) for the FL as well as COL sensing of Ag+ with LOD up to 0.226 nM. When the concentration of Ag+ was low (below 0.14 µM), the quenching followed SQE due to the less possibility of collision, while both SQE and DQE were involved at higher concentrations of Ag+ (>0.14 µM). Moreover, the applicability of this nanoprobe in real water was justified by satisfactory recoveries/RSDs (97.85–100.6%/<3.5%). The low-level quantification and visual recognition (yellow-green to blue with increasing Ag+ concentration; under UV irradiation) of Ag+ with the ratiometric probe are notable, but the involvement of toxic semiconductor QDs and Fe3+/Fe2+ interference are some limitations.119


image file: d5ra04935k-f25.tif
Fig. 25 (a) Schematic of the COL sensing activity of N,S-GQDs@PtNCLs for Ag+ detection. Reprinted from ref. 283, copyright 2022, with permission from Elsevier. (b) GQDs-induced reduction of Au3+ to AuNPs. (c) Colour change in the assembled paper sensor with the addition of different concentrations of Au3+ and corresponding calibration plot. Reproduced/adapted from ref. 114 with permission from The Royal Society of Chemistry, 2021.

Summary: It can be inferred that N-GQDs are better FL probes for Ag+ compared to other single-heteroatom doped GQDs such as S-GQDs and undoped GQDs. The Ag+ detection capability of N,S-GQDs in heterostructured/mixed form with other counterparts such as PtNCLs and CdTeQDs through the COL and ratiometric manner is also considerable but with some limitations.

6.13. Au3+

The applicability of GQDs-based sensors for the recognition of Au3+ can be seen in Tables 9 and S7. For example, the blue-emission peak of N,S-GQDs (425 nm) is quenched in the presence of Au3+ due to the coverage of their surface with AuNPs, exhibiting a low LOD (50 nM) for Au3+. The formation of AuNPs is induced by the reactivity of Au3+ with the amine groups of N,S-GQDs and electron transfer from N,S-GQDs to Au3+. However, although this probe is very selective to Au3+, its optimum sensing activity occurs at elevated temperature (45 °C, 15 min incubation time) and the presence of Fe3+ potentially interferes with the quenching process (mercaptosuccinic acid is required to mask the interference from Fe3+). Moreover, this probe could successfully analyze Au3+ in spiked-real water samples and auranofin drug.284

Later, Thanomsak et al.114 fabricated a portable COL probe for the detection of Au3+ by adsorbing GQDs on cellulosic filter paper (CFP) and coating with a polymeric membrane (PM). The hydrophobicity created by the optimal PM coating on GQDs@CFP is important to avoid the leaching of GQDs from the paper sensor and effective ion-exchange in a short incubation time (10 min). The accumulation of Au3+ into the fabricated paper sensor via the cationic ion-exchange process and reduction to AuNPs through the electron-donating capability of GQDs (Fig. 25b) led to a visualize colour change (pale yellow to pink, without the assistance of UV light, Fig. 25c) in the LR of 200–1000 µM (Fig. 25c) with an LOD of 70 µM. Moreover, the paper-based sensor could be practically applied for the quantification of Au3+ in real water samples. However, the applicability of this paper sensor is limited to determining a high concentration of Au3+ and not suitable at trace level.

Summary: Dual-doped N,S-GQDs may be a good probe for the quenching-induced FL detection of Au3+. Although the sensitivity of bare GQDs-based paper sensors is inferior to the solution-phase FL detection results, the construction of a simple and low-cost platform for the rapid, real-time, and visual COL detection of Au3+ is notable.

6.14. Alkali/alkaline-earth MIs

Crown ether-like C-GQDs (edge possessing 78% oxygen atoms in the form of C–O–C) exhibited Ca2+ selectivity in the fluorescence quenching process (LOD: 2 pM) and in vitro detection proficiency of Ca2+ in hASCs (human adipose-derived stem cells). Besides, different sizes of the crown ether-like structures showed selectivity for Mg2+ (LOD: 20 pM), Sr2+ (LOD: 8 pM), and Ba2+ (LOD: 2 pM). Noticeably, the LODs for all the detected MIs are much lower than that of the previously reported PEG-modified N-GQDs (Table S8), which can be ascribed to the strong coordination ability of the crown ether-like structure with the respective MIs.285 Subsequently, GQDs were covalently modified with two crown ethers to create GQDs–15-crown-5 and GQDs–18-crown-6 composite systems for the sensing of both Na+ and K+ via the EC method (Table 10). However, both systems suffered from selectivity issues unlike the previous system, where selectivity is achieved by confining PEG–GQDs in Na+/K+ specific ionophores (the synthesis of PEG–GQDs-confined ionophores followed a complex/expensive procedure) (Table S8). GQDs–18-crown-6 (λem: 450/550 nm due to crown ether/GQDs components) was found to be selective up to a certain extent for K+ in a ratiometric manner (enhancement/suppression of 450/550 nm peaks with increasing concentration of K+), which is hypothesized to be due to the different interactions of K+ with the oxygen moieties of crown ether and GQDs. Meanwhile, detailed structural characterization of the synthesized probe, mechanism for its selectivity, interference study, and practical applicability are lacking in this report.286
Table 10 GQDs and modified-GQDs for alkali/alkaline-earth MI, rare-earth MI, and radioactive MI sensing application
GQDs-based sensor Synthesis conditions Size range/average sizea (nm) QY (%) Sensing process LR (µM) LOD (µM) MIs Ref.
a Measured from TEM.b Size range/average size of GQDs measured from dynamic light scattering.c Size range/average size of crown-GQDs.d Size range/average size of crown-GQDs–PEG5.e Dynamic concentration range.f LR/LOD in µg L−1 or ppb.g Analytical ability in living cells.h Analytical ability in real water samples.
Alkali/alkaline-earth MIs
C-GQDs HT (Alizarin in water, 150 °C, 24 h); dialysis 1–5.5/2.7 74 FL, turn-off 0–2 × 10−4 2 × 10−6 Ca2+ 285g
GQDs–15-crown-5@SPCE Acid oxidation of MWCNTs with HNO3:H2SO4 (1[thin space (1/6-em)]:[thin space (1/6-em)]3) under ultrasonication (60 °C, 4 days); dilution and filtration; pH adjusted to 7.0; dialysis; covalently modified with 15-crown-5 or 18-crown-6 —/4.93b EC, potentiometric 1–1 × 106e Na+ 286
    " " " K+
GQDs–18-crown-6@SPCE —/4.88b " " " Na+
    " " " K+
GQDs–18-crown-6 —/4.88b FL, ratiometric " " K+
DA–GQDs Cutting of GO paste at 120 °C, 12 h; after dilution, pH adjusted to 3.0; filtration; centrifugation; dialysis; covalently modified with DA 2–5/— FL, turn-on 4.93–10.61 0.05 Ca2+ 131g
Crown-GQDs–PEG5–Gd3+ ST (o-PDA/4-bromobenzo-18-crown 6 ether in ethanol, 180 °C, 50 h); ethanol replaced with water; filtration; dialysis; modified with PEG5; loaded with Gd3+; dialysis 1.84–10.12/3.73c   FL, turn-off 2500–25[thin space (1/6-em)]000 3800 K+ 287g
1–11/4.82d Relaxometry, turn-off 5000–150[thin space (1/6-em)]000 14[thin space (1/6-em)]120 K+
[thin space (1/6-em)]
Rare-earth MIs
N-GQDs Chemical oxidation of GO/lysine with 30% H2O2 under reflux (130 °C, 4 h); removal of excess H2O2; filtration; dialysis 1.5–3.5/— 13.2 FL, turn-off 0.3–15 0.11 Eu3+ 288
GQDs Acid oxidation of GSs with HNO3:H2SO4 (3[thin space (1/6-em)]:[thin space (1/6-em)]1) under ultrasonication (18 h); diluted with water, Filtered & pH adjusted to 8.0; HT (200 °C, 10 h); filtration; dialysis 15–20/— 7.2 FL, turn-off 50–230 0.38 Ce3+ 289
GQDs/o-PDA GQDs purchased from XFNANO —/4.2 FL, ratiometric 5–100 1.0 Ce4+ 290h
GQDs Acid oxidation of GO with HNO3:H2SO4 (1[thin space (1/6-em)]:[thin space (1/6-em)]3) under reflux (110 °C, 24 h); diluted with water & pH adjusted to 8.0; filtration; dialysis 1.5–3.8/2.5 FL, turn-on 0–30 0.3 Tb3+ 291
[thin space (1/6-em)]
Radioactive MIs
GQDs Gamma radiolysis of GO/25% H2O2 in water (270 kGy, 11.75 kGy h−1); drying & dispersion in water 2.3–8.8/4.6 10.2 FL, turn-off 2.24–21.43f 0.56f U6+ 292
ER-GQDs Drop-casted GQDs on Au or GCE and electrochemically reduced EC, SWV 23.4–345.8f 2f " 292h
PA@N-GQDs HT (CA/urea in water, 160 °C, 4 h); filtration; covalently modified with PA 4.2–8.7/6.5 FL, turn-off 10–80 2.01 × 10−3 U6+ 132
" 10–60 1.35 × 10−3 Th4+


A recent report demonstrated the detection of Ca2+ and its intracellular tracking using DA–GQDs through an atypical FL turn-on fashion. It was observed that Ca2+ is prone to coordinate with the oxygen groups of DA–GQDs to improve the selectivity and sensitivity. The blocking of PET and strengthening of internal charge transfer after the coordination of Ca2+ with DA–GQDs caused a fluorescence enhancement in the detection process. Although the LOD of Ca2+ with DA–GQDs (50 nM) is much higher in comparison to previous results (Tables 10 and S8), their wide LR (4.93–10.61 µM), selectivity from specific functional group, and turn-on type FL detection are advantageous for the analysis of a broad range of concentrations with high accuracy. This probe is also applicable to quantify Ca2+ in an organic compound (calcium gluconate) with slightly lower sensitivity (LR/LOD: 14.56–45.45 µM/100 nM). Moreover, the progressive intensification of the blue fluorescence with an increase in amount of Ca2+ (0, 2.5, 5.0, and 7.5 µM) in ARPE-19 (human retinal pigment epithelium cells) living cells (Fig. 26) confirmed the Ca2+ monitoring capability of the biocompatible probe (>70% cell viability at 150 µg mL−1 concentration, incubation time: 72 h) in biological matrices.131


image file: d5ra04935k-f26.tif
Fig. 26 Confocal images of living cells (ARPE-19) after incubation with DA–GQDs (i) and further treatment with 2.5 µM (ii), 5.0 µM (iii), and 7.5 µM (iv) Ca2+ under blue field (first column; a, d, g, j), red field (middle column; b, e, h, k), and overlay (right column; c, f, i, l). Reprinted from ref. 131, copyright 2024, with permission from Elsevier.

Recently, Chen et al.287 linked Gd3+ (magnetic site) and crown ether-possessing GQDs via bridging with PEG5 molecules to construct a crown-GQDs–PEG5–Gd3+ probe for the dual-mode (FL and nuclear magnetic resonance (NMR)-based relaxometry) identification of K+. Due to the binding selectivity of crown ether with K+, the fluorescence intensity gradually decreased in the FL sensing process. Additionally, the magnetic probe showed a significant change in the relaxometry response (T1; measured from NMR analyses) rather than the probe where Gd3+ is directly attached to crown-GQDs (crown-GQDs–Gd3+) (Fig. 27a). The gradual decrease in T1 with an increase in the concentration of K+ (5–150 mM, Fig. 27b) and linearly fitted curve (Fig. 27c) by the probe showed the relaxometry-based successful detection of K+. The changes in T1 with K+ are ascribed to the variations in proton concentration at the paramagnetic centre (Gd3+), which is facilitated by the PEG5 chains (proton transporter with low energy barrier). Although the sensitivity of K+ with 1H NMR-based detection is lower than the FL method (Table 10), it provides a new sensing opportunity to detect alkali MIs and other MIs in the future. After verifying the insignificant toxic effect, the probe effectively differentiated senescent cells (related to K+ concentration, high fluorescence and different morphology) from healthy ones (amount of probe: 200 µg mL−1, incubation time: 24 h; Fig. 27d) and exhibited easy penetration capability within the blood–brain barrier, which is a crucial achievement to monitor K+-induced aging effects and other related diseases.


image file: d5ra04935k-f27.tif
Fig. 27 (a) Changes in T1 for crown-GQDs–PEG5–Gd3+ and crown-GQDs–Gd3+ probes after adding 150 mM K+. Decreasing trend in T1 with an increase in the concentration of K+ (b) and linear fitting plot between ΔT and concentration of K+ (c) for crown-GQDs–PEG5–Gd3+. (d) Healthy and senescent cells incubated with crown-GQDs–PEG5–Gd3+ and imaged under bright-field (left) and fluorescence mode (right). Reprinted from ref. 287, copyright 2025, with permission from Elsevier.

Summary: The crown ether/crown ether-like structure in GQDs can significantly improve their coordination ability with alkali/alkaline-earth MIs to achieve promising selectivity as well as sensitivity in the FL detection process. Specifically, the presence of DA functionality in GQDs is advantageous for the selective and turn-on-type sensitive detection of Ca2+ even in biological media. Additionally, the selective/sensitive detection of K+ using the crown-GQDs–PEG5–Gd3+ probe via dual sensing techniques (FL and NMR) and identifying K+-induced senescence in healthy cells are considerable achievements.

6.15. Rare-earth MIs

Various rare-earth MIs such as Eu3+, Ce3+, Ce4+, and Tb3+ could be recognized by GQDs/doped-GQDs (Tables 10 and S8). For example, Ce4+ catalyzed the oxidation of o-PDA (oxidase-like activity), resulting in a strong emission from the oxidized o-PDA (o-PDAox, 562 nm) and quenching of the fluorescence of GQDs (444 nm) through IFE. As a result, the ratiometric detection of Ce4+ by analyzing the I562/I444 intensity ratio showed a satisfactory performance (Table 10) and practical ability in lake water samples. However, the Ce4+-induced oxidase-like reaction required a long incubation time (1 h) under dark conditions, and therefore the whole sensing process is time consuming. The sensing activity also exhibited significant interference in the presence of Cu2+ and Ag+, which required EDTA and I to inhibit their interference, respectively.290

Wang et al.291 demonstrated the selective and sensitive FL sensing of Tb3+ using GQDs, which involved a significant antenna effect to increase all four fluorescence peak intensities (490 (5D47F6)/546 (5D47F5)/585 (5D47F4)/620 (5D47F3) nm, originating from Tb3+–GQDs) with Tb3+ concentration (Fig. 28a) and LR of 0–30 µM (Fig. 28b). Fig. 28c and d depict the energy transfer operation (from GQDs to Tb3+) and resulting fluorescence in the Tb-GQDs via the antenna effect. The excitation of an electron from the ground state (S0) to the excited state (S1) of GQDs by the absorption of light, followed by the transfer of its energy to Tb3+ further excites and emits long-living fluorescence via a line-type f–f transition. Although the low fluorescence from Tb3+ in aqueous medium is reasonably enhanced by GQDs, the effective emission characteristic using a high energy λex (230 nm) and requirement of long sensing time (30 min for reaction between GQDs and Tb3+) are limiting factors.


image file: d5ra04935k-f28.tif
Fig. 28 Fluorescence spectra of GQDs in the presence of different amounts of Tb3+ (0 to 250 µM) (a) and linear plot corresponding to 546 nm peak intensity vs. Tb3+ concentration (b). Schematic of the energy transfer steps involved in the sensing operation (c) and corresponding fluorescence spectrum (d). Reprinted from ref. 291, copyright 2019, with permission from Elsevier. (e) Schematic of the Th4+/U6+ interaction with PA@N-GQDs during the sensing process. Reprinted from ref. 132, copyright 2024, with permission from Elsevier.

Summary: Both undoped and doped-GQDs (N-GQDs) are applied for the FL detection of rate-earth MIs. Good sensitivity in the FL detection of Ce4+ is achieved with bare GQDs through their oxidase-like activity and ratiometric manner rather than N-GQDs via a turn-off manner. Moreover, the specific antenna effect between GQDs and Tb3+ qualified Tb3+ according to the FL turn-on principle with a satisfactory performance.

6.16. Radioactive MIs

The applicability of GQDs-based platforms for the sensing of radioactive MIs is presented in Tables 10 and S8. For instance, GQDs synthesized from gamma radiolysis were used for the low-level detection of U6+ (LOD: 0.56 ppb) through electrostatic interaction and a FRET-based fluorescence quenching process. However, the FL detection strongly suffered from interference from Fe3+ (98% quenching). Furthermore, the electrochemically reduced GQDs (ER-GQDs) were utilized for the EC detection of U6+ (via cathodic SWV response, which is superior to the cathodic DPV signal). The estimated LOD (2 ppb) was found to be higher than FL method but EC detection is advantageous for higher concentration measurement (Table 10), elimination of Fe3+ interference (Fe3+ is precipitated to red-coloured Fe(OH)3 in saturated Na2CO3 electrolytic solution), and sensing of U6+ in ground water.292 The extraction of large-sized GQDs clusters (∼290 nm) from a supramolecular hydrogel and their utilization for the wide LR detection of U6+ can be revealed in Table S8. Subsequently, PA-functionalized N-GQDs (PA@N-GQDs, 7.4 at% nitrogen) were applied for the turn-off based FL detection of U6+ as well as Th4+ due to the strong binding interactions between them (Fig. 28e). Based on the KSV, binding constants, and association/dissociation constants analyses, it is deduced that Th4+ possesses higher binding affinity with PA@N-GQDs in comparison to U6+, resulting in 99/60% instant fluorescence quenching (within one minute) with the addition of Th4+/U6+. As a result, the probe could detect Th4+ in a more sensitive manner than U6+ (Table 10). It is also noticeable that the achieved LOD for U6+ is much lower than that in previous reports (Tables 10 and S8). However, the reversibility of the probe with EDTA is not satisfactory and cannot be used respectively.132

Summary: Nitrogen-doping and the presence of additional functional groups (e.g., PA) in GQDs are advantageous for their selective interaction with radioactive cations (U6+ and Th4+), and consequently the attainment of high sensitivity in FL detection. Moreover, the U6+ detectability of ER-GQDs via the EC method opens the possibility for the development of GQDs-based EC platforms to quantify hazardous radioactive MIs.

6.17. Multiple HMIs

6.17.1. EC detection. The GQDs-based electrode materials involved in the EC detection of multiple HMIs are consolidated in Tables 11 and S9. For instance, the highly selective and sensitive EC detection of Hg2+/Cu2+/Cd2+ (separately, LODs of Hg2+/Cu2+/Cd2+: 9.8/8.3/4300 pM) using vertically aligned mesoporous silica-nanochannel film (VMSF)-confined OH-GQDs and NH2-GQDs through the reduction of HMIs, followed by DPV-based anodic stripping at specific potentials is inspiring (Table 11); however, the complex/precise experimental conditions in the fabrication of modified-electrodes and the potential-specific electrodeposition of different HMIs before reduction/stripping operations cannot be avoided. Functional group-containing GQDs significantly amplified the signal response after their selective interaction with the analytes in the nanoconfined region via effective charge transfer, while the VMSF layer on the ITO electrode served as an anti-fouling and interference inhibitor. Besides Cd2+ recognition in soil-leached solution with NH2-GQDs@VMSF/ITO, OH-GQDs@VMSF/ITO was used for the quantification of Hg2+ in the presence of Cu2+ and Hg2+/Cu2+ sensing in seafood/human serum, dictating the practical approach of the fabricated probes.293
Table 11 GQDs, modified-GQDs, and GQDs involved with other counterparts for multiple HMI sensing application
GQDs-based sensor Synthesis conditions Size range/average sizea (nm) QY (%) Sensing process LR (µM) LOD (µM) HMIs Ref.
a Measured from TEM.b Size range/average size of GQDs in confined system/before functionalization/involved with other counterparts.c Absolute QY.d LR/LOD in ppm.e LOD in µg L−1 predicted from machine learning-based algorithm.f Analytical ability in real water/other real samples.g Simultaneous detection capability of multiple HMIs.h Paper-based sensing capability.
EC sensor
OH–GQDs@ VMSF/ITO HT (TNP in 0.125 M NaOH aqueous solution, 200 °C, 2 h); dialysis; filtration; electrophoresis confinement in VMSF/ITO electrode 0.9–2.9/1.83b 21c EC, DPV 1 ×10−5–0.001, 0.001–0.5 9.8 × 10−6 Hg2+ 293f,g
        " 1 × 10−5–0.001, 0.001–1.5 8.3 × 10−6 Cu2+ 293f
NH2–GQDs@ VMSF/ITO HT (TNP in 0.4 M NH3/1.5 M hydrazine hydrate aqueous solution, 200 °C, 2 h); dialysis; filtration; electrophoresis confinement in VMSF/ITO electrode 1.3–2.9/1.9b 29.8c " 0.02–1, 1–20 0.0043 Cd2+ 293f
N,S-GQDs@ GCE HT (PANI in 0.05 M H2SO4 aqueous solution, 220 °C, 12 h); drop-casted on GCE 3–5/5.4 EC, DPV 0.0001–100 1 × 10−6 Cd2+ 198f,g
" 0.0001–100 1 × 10−5 Pb2+
" 0.0001–100 1 × 10−6 Hg2+
[thin space (1/6-em)]
FL sensor
Undoped/doped-GQDs
N,S-GQDs HT (TNP/TU in 10 mM NaOH aqueous solution/10% DMF, 200 °C, 10 h); dialysis 1.6–2.8/2.1 23.2c FL, turn-off 0.01–25 0.008 Fe3+ 295f,g
" 0.4–180 0.25 Cu2+
" 0.1–140 0.05 Ag+
N-GQDs Plasma-contacting liquid synthesis (glucosamine in water, plasma irradiation, atmospheric pressure, below 80 °C, 10 min); filtration; dialysis 2–8/4.8 FL, turn-off 0–95 Fe3+ 296
" " " Pd2+
" " " Hg2+
" " " Cu2+
" " " Pb2+
" " " Co2+
N-GQDs HT (Bean dregs power in water, 180 °C, 12 h); filtration; dialysis 0.38–3.74/1.63 21.3 FL, turn-off 0–2000 2.5 Ce4+ 297
" 0–1600 1.9 Fe3+
N-GQDs HT (GO obtained from spent graphite/NH3·H2O in water, 200 °C, 1.5 h); filtration; dialysis 0.5–4.5/2.44 11.04 FL, turn-off 60–200 0.23 Fe3+ 298f
FL, turn-on 20–200 1.101 Al3+
[thin space (1/6-em)]
Functionalized GQDs
Am–GQDs Electrolysis of graphite rod in NaOH/ethanol, 24 h; dialysis; pH adjusted to 7.0 and re-dispersed in water; gamma irradiation (25 kGy, Ar) with 4 vol% ethylenediamine/3 vol% isopropyl alcohol; dialysis —/16 5.82 FL, turn-off 0–7.5 1.79 Co2+ 300f
" 0–4.0 0.657 Pd2+
" 0–45 2.55 Fe3+
PEG–Pb-GQDs HT (Cane molasses/lead acetate in water, 190 °C, 24 h); filtration; mixed with PEG-200 1–1.8/1.4 30.31 FL, turn-off 28–44 0.29 Fe3+ 301f,g
" 20–140 1.08 Cu2+ 301g
" 20–160 3.24 Ag+ 301g
NN–GQDs Electrolysis of graphene foam in 0.1 M NaOH/urea ethanolic solution, 30 V; centrifugation; filtration; covalent modification with NN 2–7/∼3b FL, turn-off 1e Hg2+ 302f,g
" 3e Fe3+
[thin space (1/6-em)]
GQDs involved with other counterparts
DPA–GQDs/Amino acid HT (CA/DPA in water, 200 °C, 2.5 h); diluted with water; combined with different amino acids —/0.8b FL/COL, turn-off 0.01–1d 0.1d Cu2+ 303f,h
" " " Hg2+
" " " Fe3+
N-GQDs/GSH–AuNCLs HT (GO/NH3·H2O in water, 170 °C, 6 h); filtration; mixed with GSH–AuNCLs aqueous solution 1.3–3.32/2.26b 24.42 FL, ratiometric 0.08–6 0.00412 Cu2+ 304f,h
" 1–40 0.943 Cd2+


Later, Saisree et al.198 employed N,S-GQDs@GCE for the detection of three highly toxic HMIs (Cd2+/Pb2+/Hg2+) with good sensitivities (12/13/5 µA µM−1) and experimental LODs of 1/10/1 pM during single HMI sensing (Table 11). The modified-electrode was fabricated by the simple-drop casting of an N,S-GQDs dispersion on a freshly polished GCE. Benefitting from the improved conductivity and electrocatalytic activity of N,S-GQDs, HMIs are directly reduced to the corresponding metal species with a well-separated EC response (Fig. 29). The high current response in the DPV curves with different concentrations of Cd2+ (Fig. 30a), Pb2+ (Fig. 30b), and Hg2+ (Fig. 30c) in the presence of two other HMIs and the corresponding broad LRs in the two concentration ranges (Fig. 30d–f) exhibited the simultaneous detection capability of the modified-electrode. The minimum concentration analyses (Fig. 30g–l) indicated that the LODs in the simultaneous detection are equivalent to that of the individual HMI measurement. Moreover, good reusability (∼98% current response after 30 cycles), reproducibility (≤2% RSDs from five independent electrodes), stability (≥80% after 60 days), and ground/sea/waste water sample analytical capability (∼100%/≤ 0.5% recoveries/RSDs) are promising aspects of this sensor system.


image file: d5ra04935k-f29.tif
Fig. 29 Schematic showing the deposition of N,S-GQDs on GCE surface to fabricate a working electrode for the EC cell, and its DPV response in the simultaneous presence of Cd2+, Pb2+, and Hg2+, resulting in peaks at different stripping voltages. Reprinted (adapted) with permission from ref. 198, copyright 2023, the American Chemical Society.

image file: d5ra04935k-f30.tif
Fig. 30 DPV signals of Cd2+ (a), Pb2+ (b) and Hg2+ (c) within the concentrations of 10−12 to 10−3 M in the presence of two other HMIs (1 mM). Plots of current vs. concentration of Cd2+ (d), Pb2+ (e), and Hg2+ (f). DPV signals of 0.1 M PBS, along with the minimum concentrations (LOD) of Cd2+ (g), Pb2+ (h), and Hg2+ (i). (j, k, l) Enlarged view of (g, h, i) showing LODs of 1/10/1 pM for Cd2+/Pb2+/Hg2+ during their simultaneous detection. Reprinted (adapted) with permission from ref. 198, copyright 2023, the American Chemical Society.
6.17.2. FL detection.
6.17.2.1. Undoped/doped-GQDs. Multiple HMIs are also detected by GQDs/doped-GQDs through the FL method (Tables 11 and S9). For instance, the fluorescence of starch-derived GQDs was quenched in the presence of Fe3+/Cu2+/Cr3+, and further recovered after the addition of group IIIA MIs (Al3+, gallium ion (Ga3+), and indium ion (In3+)), highlighting the cationic-controlled fluorescence switching behaviour of the GQDs due to the cation-driven changes in the π-conjugated region. The Ga3+ cation showed the highest fluorescence recovery of 67/81/70% after quenching operation with Fe3+/Cu2+/Cr3+. However, the authors could not quantify the sensitivity metrics.294 HT-synthesized N,S-GQDs (87.8% production yield) showed potential for the parallel detection of Fe3+, Cu2+, and Fe3+/Ag+ under different masking agents, namely, Cys, AA, and EDTA, respectively, with considerable sensitivities (Table 11). It was found that N,S-GQDs are more selective towards the three HMIs rather than undoped GQDs (containing abundant –OH groups). The –OH groups of N,S-GQDs are specific to interact with Fe3+, while the nitrogen-dopants of N,S-GQDs preferably interacted with Cu2+ and the electron-donating nature of N,S-GQDs reduced Ag+ to AgNPs. However, Ag+ detection required a longer incubation time (30 min) compared to Fe3+ (8 min) and Cu2+ (4 min) due to the contradictory HSAB interactions (N,S-GQDs: hard base, Ag+/Fe3+/Cu2+: soft acid/hard acid/moderate acid).295 The rapid synthesis of N-GQDs (∼10 min; 7.4% nitrogen content; 54% pyrrolic and graphitic nitrogen, 40% amine/pyridine/C3N4 configuration) using the plasma-contacting liquid (PCL) approach (Table 11) has become advantageous to incorporate plentiful oxygen/nitrogen-containing functional groups (both at the edges and on the surfaces) and numerous defect sites (Stone–Wales and trivacancy) for interaction with multiple HMIs. As a result, the N-GQDs showed capability for the successful detection of Fe3+, palladium ion (Pd2+), Hg2+, Cu2+, Pb2+, and Co2+ with good LRs (Table 11). PCL-synthesized N-GQDs showed a quenching effect with multiple HMIs (individual basis, minimum detectable concentration: 10 µM), showing the possibility to identify HMI-containing polluted water specimens; however, the authors did not demonstrate the simultaneous detection capability and practical utility of the probe.296

Recently, biomass (waste bean dregs)-derived N-GQDs were used in the FL method to determine two MIs (Ce4+ and Fe3+, Table 11). The fluorescence quenching-based identification capability of Ce4+ showed a broad LR (0–2000 µM) with the negotiation of larger LOD compared to previous reports on the detection of Ce4+ using chemical precursor-derived GQDs or N-GQDs (Tables 10 and S8). Meanwhile, this report did not provide the interference study, mechanistic investigation, and real-time applicability of the probe.297

Another very recent report utilized spent graphite (from waste lithium ion batteries) for its economical upcycling into crystalline N-GQDs (EIPL characteristic, QY: 11.04%). Their defective surface, small size (average size: 2.5 nm), and large amount of nitrogen-containing functionalities (nitrogen content: 2.67%; pyridinic and pyrrolic) facilitated active sites for the coordination of HMIs. Consequently, the N-GQDs could detect Fe3+ and Al3+ through the FL turn-off (98.4% quenching) and FL turn-on (38% enhancement) routes, respectively, and within the permissible limits according to the EPA (Table 11). The recovery tests of both HMIs in real water samples are satisfactory but the report lacks the detailed sensing mechanism.298 The distinct behaviours of the triple-colour emissive N,S-GQDs (440 nm (blue)/550 nm (green)/650 nm (red) emission at λex = 352/449/559 nm, Fig. 31a; QY: 65.4/61.4/24.6% at 440/540/630 nm emission) with 10 HMIs (Mn2+, Fe3+, Cu2+, Zn2+, Pb2+, Ni2+, Cd2+, Ag+, Co2+, and Ba2+, Fig. 31b) were explored to differentiate these HMIs using linear discriminant analysis (LDA) and hierarchical cluster analysis (HCA) (Fig. 31b). The analytical method could successfully quantify Fe3+/Cu2+/Pb2+/Cd2+/Ni2+/Co2+/Mn2+/Zn2+/Ba2+/Ag+ at a minimum concentration of 0.50/0.11/0.55/2.10/1.14/1.14/2.03/3.92/0.96/0.29 µM, discriminated HMIs (Fe3+ and Cd2+) from mixtures, and showed applicability in environmental water bodies (tap and lake water) with satisfactory recoveries.299


image file: d5ra04935k-f31.tif
Fig. 31 (a) Schematic of three colour (blue/green/red) emitting N,S-GQDs (a) and their application in the quantification and discrimination of 10 HMIs through LDA and HCA approach (b). Reprinted from ref. 299, copyright 2025, with permission from Elsevier.

6.17.2.2. Functionalized GQDs. The utility of functionalized GQDs for the identification of multiple HMIs can be ascertained from Tables 11 and S9. For example, apart from Co2+ and Fe3+ recognition, Am–GQDs showed the FL detection of Pd2+ (first report) with an LOD of 657 nM. Complexation of the HMIs with the functional groups of Am–GQDs and cation–π interactions between them induced their aggregation, and consequently fluorescence quenching. However, the sensitivity/selectivity of this probe is very specific to its preparation conditions via gamma irradiation and a proper understanding is lacking.300 Pb-doping along with PEG-modification in biomass (cane molasses)-derived GQDs (PEG–Pb-GQDs) showed a significant enhancement in fluorescence intensity (QY: 30.31%) compared to bare GQDs or PEG–GQDs (QY: 10.44 or 21.32%). Consequently, the PEG–Pb-GQDs fluorescent probe selectively detected Fe3+, Cu2+, and Ag+ from a mixed HMI solution using EDTA + TU, F + SCN +Cl, and EDTA + F masking agents, respectively, but with inferior sensitivities in comparison to the previous masking-based strategies using chemically synthesized N,S-GQDs probes (Table 11). Furthermore, the interference effect of Co2+, Ni2+, Mn2+, and Pb2+ on this probe (showed significant quenching) was not addressed.301 The introduction of specific functional groups such as Alizarine Red S and Eriochrome Black T on the surface of GQDs can be advantageous to develop simple COL sensors for the analysis of different HMIs according to colour variations (Table S9).

Recently, Llaver et al.302 developed an ML-enabled algorithm using a functionalized GQDs nanoprobe to selectively detect two HMIs in a standard solution as well as in a complex real water matrix. A schematic illustration of the synthesis of the urea-modified GQDs via an electrochemical method, followed by their chemical-functionalization with 1-nitroso-2-naphthol (NN) to obtain the NN–GQDs fluorescent probe for the discrimination as well as quantification of Hg2+ and Fe3+ with the accreditation of MI algorithm is shown in Fig. 32. The distinct and intense emission from the NN–GQDs (456 nm at 326 nm λex) exhibited a slight blue-shift, along with a quenching effect in the presence of Fe3+ rather than only quenched fluorescence with Hg2+, enabling the basis of assimilation in the algorithm to discriminate and quantify these HMIs. Based on the data analyses, the LODs for Hg2+ and Fe3+ were predicted to be 1.0 (9.0) and 3.0 (8.0) µg l−1 under single HMI (in the presence of other HMIs), respectively, indicating a good merit of quantification and simultaneous detection possibility. Moreover, the ML model was successfully applied to predict the HMI contents simultaneously in natural (tap, river, and dam) water systems with an accuracy close to that measured from standard instrumental methods.


image file: d5ra04935k-f32.tif
Fig. 32 Step 1: synthesis of urea-modified GQDs via the electrolysis of graphene foam in urea-containing electrolyte. Step 2: covalent modification with NN chelating molecules to obtain NN–GQDs for the selective and sensitive detection of Hg2+ and Fe3+ in a mixed solution. Step 3: employing ML algorithm to improve the quantification of HMIs. Reproduced/adapted from ref. 302 with permission from The Royal Society of Chemistry, 2024.

6.17.2.3. GQDs involved with other counterparts. Systems are also constructed by combining functionalized/doped-GQDs with other counterparts for multi HMI recognition (Table 11). For example, DPA–GQDs-supported amino acids were applied to recognize Cu2+/Hg2+/Fe3+ via the FL as well as COL method (LOD: 0.1 ppm). Cys amino acid is suggested to amplify the colour response in the COL detection process. As a result, a microfluidic paper-based platform (fabricated using DPA–GQDs and Cys) could recognize Cu2+/Fe3+ under natural light and quantify Cu2+ in environmental fluids/human urine samples.303 The quantification of Cu2+ and Cd2+ in a ratiometric manner with the assistance of N-GQDs and GSH-functionalized AuNCLs (GSH–AuNCLs) showed a satisfactory performance (Table 11). The fluorescence quenching/enhancement of GSH–AuNCLs with Cu2+/Cd2+ and insignificant effect on the fluorescence behaviour of N-GQDs resulted in the design of a sensitive ratiometric system. Apart from its recognition capacity in scallop samples, the fabricated paper-based platform using N-GQDs/GSH–AuNCLs showed possibility for the visual detection of both HMIs (distinct colour with Cu2+ and Cd2+ under UV light).304

Summary: The achievement of picomolar-level detection capability for multiple toxic HMIs (Pb2+, Hg2+, and Cd2+) in broad LRs without observable interference by the EC method using dual-doped N,S-GQDs is inspiring. FL methods also showed the simultaneous detection of multiple HMIs using GQDs-based platforms, but their selectivity is limited to the use of masking agents and their performance metrics are inferior in comparison to the EC method. N-GQDs and N,S-GQDs are effective probes compared to undoped GQDs for the efficient sensing of multiple HMIs through the FL method. Moreover, the creation of abundant defects and nitrogen/oxygen-containing covalent groups on GQDs are advantageous to achieve strong affinity with HMIs. The recent demonstration of LDA/HCA methods to discriminate multiple HMIs according to different colour responses with N,S-GQDs is notable. The functional group-enabled dissimilar interactions of functionalized GQDs with different HMIs can provide FL or COL detection platforms. Specifically, NN functional groups on GQDs can interact differently with Fe3+/Hg2+ to obtain a non-identical fluorescence outcome for the development of an ML algorithm and prediction of HMIs at very low concentrations with high accuracy. The ratiometric detection of Cu2+/Cd2+ with a binary system containing N-GQDs and GSH–AuNCLs is also considerable, which is extended to a paper-based device for the detection/discrimination according to different colour responses under UV light.

7. GQDs-based/involved sensors in the detection of anions

7.1. PO43−

The GQDs-based recognition of the PO43− anion can be accessed from Tables 12 and S10, where turn-off-on based strategies are the most common. The initial fluorescence quenching of GQDs/doped-GQDs is generally induced by rare-earth MIs or HMIs (Tables 12 and S10), but anion (Mo7O246−)-mediated quenching has also been reported in the sensitive sensing of PO43− (Table S10). For example, biomass (corn straw)-derived GQDs showed fluorescence quenching with Ce4+/Fe3+ via the AIQ mechanism and its recovery with PO43− due to the obstruction of their aggregation (Fig. 33a). The prominent recovery of the fluorescence with GQDs/Ce4+ rather than the GQDs/Fe3+ system in the presence of PO43− indicated the higher stability of the GQDs–Fe3+ complex. Consequently, the GQDs/Ce4+ system showed much better sensitivity in the detection of PO43−compared with GQDs/Fe3+ (Table 12) and applicability towards the analysis of real water samples as well as the construction of paper-based sensors.305
Table 12 GQDs, modified-GQDs, and GQDs involved with other counterparts for anion sensing applicationa
GQDs-based sensor Synthesis conditions Size range/average sizeb (nm) QY (%) Sensing process LR (µM) LOD (µM) Ref.
a THPC: tetrakis(hydroxymethyl)phosphonium chloride, PEI-EC: ethylenediamine-end-capped-polyethylenimine.b Measured from TEM.c Size range/average size of GQDs involved with other counterparts.d QY of GQDs involved with other counterparts.e Absolute QY.f LR in µL of 1 ppm analyte.g LR/LOD in µg mL−1 from paper-based sensor.h LR/LOD in the detection of S2O32−.i LR/LOD in the detection of ONOO.j LR/LOD in the detection of HSO3.k LOD in ppm.l Analytical ability in real samples.m Paper-based sensing capability.n Analytical ability in living cells.
PO43
GQDs/Ce4+ HT (Corn straw powder in water, 170 °C, 12 h); centrifugation; filtration 1.5–4/2.67 15.65 FL, turn-off-on 0.1–2, 2–20 0.06 305l,m
GQDs/Fe3+ " 0.1–1.4 0.09 305l
N-GQDs/Ce4+ HT (Bean dregs power in water, 180 °C, 12 h); filtration; dialysis 0.38–3.74/1.63 21.3 FL, turn-off-on 0–1400 297
[thin space (1/6-em)]
P2O74
N,S-GQDs/Fe3+ Pyrolysis (CA/GSH, 200 °C, 15 min); dissolved in water and pH adjusted to 5.0; dialysis <8/3 36.3 FL, turn-off-on 1–1000 0.81 306l
[thin space (1/6-em)]
ClO
DAP–GQDs Acid oxidation of graphite flake with HNO3:H2SO4 (1[thin space (1/6-em)]:[thin space (1/6-em)]3) under ultrasonication (2 h) and reflux (120 °C, 24 h); pH adjusted to 7.0; filtration; dialysis; covalent modification with DAP 1–5/2.9 13.4 FL, turn-off 0–8 0.0126 130l,n
[thin space (1/6-em)]
S2
Eu-GQDs/ZIF-8 ST (GO/EuCl3·6H2O in DMF, 200 °C, 7 h); gel permeation chromatography; non-covalent adsorption on ZIF-8; centrifugation <10/—c FL, turn-on 0–600f 0.12k 307
[thin space (1/6-em)]
CN
GQDs@ZIF-11 Pyrolysis (CA, 200 °C, 15 min); mixed in 10 mg per mL NaOH solution and pH adjusted to 7.0; in situ encapsulation in ZIF-11 2.5–8/∼5.2c 27d FL, turn-off 0.15–30 0.0145 308l
N-GQDs/Ag+ HT (CA/tris(hydroxymethyl)-aminomethane in water, 205 °C, 2.5 h); dialysis 1–12/5.4 57.9e FL, turn-off-on 0.5–25g 0.08g 309l,m
[thin space (1/6-em)]
NO2(FL sensor)
N,P-GQDs HT (THPC/PEI-EC in water, 230 °C, 8 h); pH adjusted to 7.0; dialysis 1.5–7.5/4.2 9.4e FL, turn-off 0.005–0.03 0.0025 310n
N-GQDs Pyrolysis (Onion slice, 220 °C, 4 h, N2); ST (obtained solid in DMF/H2O, 190 °C, 4 h); dialysis <15/10 15.7 FL, turn-off 0.3–1400 0.1 311
[thin space (1/6-em)]
NO2(EC sensor)
GQDs/PCN-222 @FTO Ar/DC microplasma treatment of starch in 0.1 M NaOH aqueous solution, 1 h; filtration; impregnated in mesoporous PCN-222; drop-casted on FTO 1.5–5/3.1c EC, Amp 40–18000 6.4 312
CoPc/GQDs@ GCE HT (CA/NaOH in water, 160 °C, 4 h); centrifugation, dialysis; non-covalently conjugated with CoPc; drop-casted on GCE 2–6/∼3.5c EC, ChAmp 0–1000 0.17 313
CoPc/N-GQDs @ GCE HT (CA/urea in water, 160 °C, 4 h); centrifugation, dialysis; non-covalently conjugated with CoPc; drop-casted on GCE 2–5/∼3.2c " 0–1000 0.25
[thin space (1/6-em)]
I
N,S-GQDs/Ce4+ Pyrolysis (CA/Cys, 200 °C); diluted with water 2–4/— 85.6 FI-CL, turn-off 0.04–3 0.00423 192l
[thin space (1/6-em)]
F
Gd3+-loaded PEG–GQDs Purchased from CASYUEDA materials Technology 2–7.2/4.2 ULF-NMR relaxometry 0.01–100 0.01 314l
[thin space (1/6-em)]
SCN
GQDs/AuNPs hybrid Chemical oxidation of graphite powder with KMnO4/H2SO4; HT (obtained solid/TSC in water, 150 °C, 2 h); filtration; dialysis; GQD-assisted synthesis of AuNPs 3–5/—c 9.6d COL, turn-on 0.01–0.1 0.003 315l
[thin space (1/6-em)]
Other anions
N-GQDs/I2 ST (julolidine/acetic acid in ethanol, 200 °C, 12 h); filtration; dialysis —/4.8 53 FL, turn-off-on 0.002–0.01h 9.3 × 10−5h 316l,n
Cy5.5–N-GQDs Acid oxidation of graphene with HNO3:H2SO4 (1[thin space (1/6-em)]:[thin space (1/6-em)]4) under reflux (90 °C, 10 h); diluted with water and pH adjusted to 7.0; filtration; covalently modified with Cy5.5 dye 1–5/3.5c 11.6 FL, ratiometric 0–6i 0.03i 317n
OH-GQDs/PPy-Br HT (TNP in 0.2 M NaOH aqueous solution, 200 °C, 10 h); filtration; dialysis; non-covalently conjugated with PPy-Br dye <5/—c 21 FL, ratiometric 0.1–2j 0.036j 141l,n



image file: d5ra04935k-f33.tif
Fig. 33 (a) Schematic showing the aggregation-based quenching of the fluorescence of GQDs in the presence of Ce4+/Fe3+ and further switch-on of their fluorescence due to the inhibition of their aggregation in the presence of PO43−. Reprinted from ref. 305, copyright 2022, with permission from Elsevier. Fluorescence spectra (b) and corresponding fluorescence intensity plot (c) of N-GQDs/Ce4+ system according to different concentrations of PO43− (0–1.4 mM). Reprinted from ref. 297, copyright 2025, with permission from Elsevier.

Recently, PO43− was again sensed with the same strategy using a biomass (Bean dregs)-derived N-GQDs/Ce4+ system via the gradual enhancement of its fluorescence intensity as the PO43− concentration increased step-by-step (Fig. 33b) and exhibited a very broad LR (0–1400 µM, Fig. 33c) in comparison to previous results (Tables 12 and S10).297

Summary: The turn-off-on-based strategy is the most common for the detection of PO43−, where the quenching of the fluorescence of GQDs or doped-GQDs with cheap Ce4+ rather than costly rare-earth cations (Eu3+ and Dy3+) is notable. Among the doped-GQDs, N-GQDs can be employed as an efficient probe for the sensing of PO43−. Moreover, the effective detection capability for PO43− using biomass-derived GQDs or N-GQDs is considerable.

7.2. P2O74−

The turn-off-on-based sensing attributes for P2O74− using doped-GQDs can be revealed from Tables 12 and S10. For instance, the Fe3+-mediated fluorescence quenching of N,S-GQDs and its recovery in the presence of P2O74− (Fe3+ possesses high affinity for P2O74− rather than the nitrogen/sulphur-containing functional groups of N,S-GQDs) avoided the use of expensive Eu3+ (used earlier, Table S10) and demonstrated a wide LR of 1–1000 µM (LOD: 810 nM) in the detection of P2O74−. Moreover, a clinical diagnosis of arthritis is possible with the sensor due to the P2O74− assay in human synovial fluid.306

Summary: N,S-GQDs have shown better potential compared to single-heteroatom doped N-GQDs in the turn-off-on-type FL detection of P2O74−. Moreover, fluorescence quenching with cheap Fe3+ rather than the costly Eu3+, and subsequent P2O74−-driven recovery is an applicable sensing approach to attain a reasonable selectivity/sensitivity for P2O74−.

7.3. Hypochlorite (ClO)

GQDs-based systems have been successfully used for the sensing of ClO via FL, CL, and COL methods (Tables 12 and S10). For instance, the CL signal produced by the oxidation of GQDs with ClO greatly increased (∼18-fold) in the presence of cetyltrimethylammonium bromide, which enabled the detection of ClO in a wide LR (Table S10). Later, edge-functionalized GQDs (DAP–GQDs) exhibited acceptable sensitivity (Table 12) in the identification of ClO. The nanoprobe triggered AIEE-based turn-on behaviour with Pb2+, while a turn-off response to ClO with the participation of hydrogen-bonding-induced energy transfer (Fig. 34a). The presence of amine moieties in DAP favoured the formation of a hydrogen-bonded framework between DAP–GQDs and O–Cl. Consequently, the probe and analyte reached in close proximity, resulting in little aggregation and energy migration. The recoveries of 92–123% for ClO in tap/lake water samples and the monitoring of exogenic ClO in HeLa cells justified the practicability of this sensor.130
image file: d5ra04935k-f34.tif
Fig. 34 (a) Schematic of the AIEE-based sensing of Pb2+ and hydrogen-bonding/energy transfer-induced turn-off detection of ClO. Reprinted (adapted) with permission from ref. 130, copyright 2021, the American Chemical Society. (b) Schematic of the sensing process of S2− using Eu-GQDs confined in a ZIF-8 framework. Reprinted from ref. 307, copyright 2017, with permission from Elsevier.

Summary: Although ClO can trigger CL enhancement or fluorescence switch-off of bare GQDs/modified-GQDs, the utility of functionalized GQDs and FL method for the purpose of ClO sensing is advantageous to achieve nanomolar-level sensitivity. Nitrogen-rich functionalities (specifically, DAP) on GQDs have shown capability for the selective and sensitive detection of ClO via an energy transfer process.

7.4. S2−

GQDs-based sensors have also been fabricated to monitor toxic S2− (Tables 12 and S10). Besides the previous turn-off-on-based approach (Table S10), Sammi et al.307 explored a new approach to sense S2− with a satisfactory performance (Table 12). The non-covalent host–guest interaction between Eu-GQDs and zeolitic imidazole framework (ZIF-8) NPs significantly increased the fluorescence intensity of the Eu-GQDs/ZIF-8 nanocomposite due to the better separation/dispersion of Eu-GQDs in the host matrix (Fig. 34b). After the addition of S2−, the separation of Eu-GQDs became more prominent in ZIF-8, resulting in the further intensification of fluorescence and analysis of S2− via the turn-on mode.

Summary: The turn-on detection of S2− via the host–guest interaction between metal-doped GQDs (Eu-GQDs) and ZIF-8 is an effective sensing approach. Additionally, the S2− sensing performance of dual-functionalized GQDs (SA, GSH–GQDs; SA: sulfanilic acid) via Cu2+-mediated fluorescence quenching and S2−-driven recovery is also considerable.

7.5. Cyanide (CN)

Various sensing platforms have been constructed with the participation of GQDs/doped-GQDs to monitor CN (Tables 12 and S10). For instance, a GQDs-encapsulated ZIF-11 (GQDs@ZIF-11) composite was utilized for the selective/sensitive detection of CN with a low LOD (14.5 nM). Here, CN effectively suppressed the interaction between the functional groups of GQDs and benzimidazole moiety present on ZIF-11 to turn-off the fluorescence signal. Moreover, the detection of CN in real samples (apple seeds/bitter almonds) was also accomplished by this probe with reasonable recoveries (96.7–102.7%).308

Malahom et al.309 developed a fluorescent paper-based test kit using N-GQDs (absolute QY: 57.9%) for the selective quantification of CN with considerable sensitivity (Table 12). A schematic representation of the turn-off-on-based detection process is shown in Fig. 35, where the fluorescence signal of N-GQDs is quenched by Ag+ via the PET mechanism, followed by switched-on behaviour in the presence of CN (production of HCN and [Ag(CN)2] complex formation according to eqn (1) and (2), respectively, Fig. 35) via the leaving-off of Ag+ from the N-GQDs surface. Additionally, the fabricated kit was found to be a promising analytical tool for the quantification of CN in real juice/food samples (recoveries: 102.6–109.3/97.1–109.4%), along with satisfactory storage capability (30 days) and inter-/intra-day precision below 2% RSD.


image file: d5ra04935k-f35.tif
Fig. 35 Schematic of the possible mechanism involved in the detection of CN using an N-GQDs-containing test paper kit. Reprinted (adapted) with permission from ref. 309, copyright 2023, the American Chemical Society.

Summary: Both N-GQDs and N,S-GQDs are applicable to detect CN with the involvement of Ag+ or AgNPs; however, the construction of portable paper kits using N-GQDs in the CN detection process up to real sample level is a good attempt. Additionally, the compositing of GQDs with a ZIF moiety can also build a selective/sensitive platform for CN.

7.6. NO2

7.6.1. FL detection. Although a 2.5 nM LOD was achieved for NO2 (first report) using N,P-GQDs, this probe is limited to a very narrow concentration range (Table 12). Electron transfer between the –NH2 groups of N,P-GQDs (nitrogen/phosphorus content: 5.85/8.71%) and NO2 quenched the inherent fluorescence of N,P-GQDs (incubation time: 20 min). Moreover, the fluorescence imaging-based detection of NO2 in living cells (T24 cells) using N,P-GQDs demonstrated their intracellular detection possibility.310 Later, a similar interaction/electron transfer-based quenching mechanism using N-GQDs (onion biomass-derived) could detect NO2 in a broad LR with satisfactory LOD (Table 12). However, effective quenching activity with this probe required a relatively longer interaction time (30 min) compared to previous reports.311
7.6.2. EC detection. The EC method is also used in the detection of NO2 using GQDs involved systems (Tables 12 and S10). For instance, GQDs were directly impregnated into a porphyrinic zirconium-based metal–organic framework (PCN-222) to construct a GQDs/PCN-222 mesoporous structure, which exhibited 100-times higher electrical conductivity (9 × 10−11 S cm−1) in comparison to bare PCN-222 (6 × 10−13 S cm−1). As a result, the amperometric (Amp) response of GQDs/PCN-222@FTO (FTO: F-doped tin oxide) showed a superior EC sensing performance for NO2 (Table 12) in comparison to PCN-222@FTO (LR/LOD: 200–20[thin space (1/6-em)]000/50 µM). The response current in the EC detection is the result of NO2 oxidation, which is effectively electrocatalyzed by the GQDs involved composite material.312 Subsequently, Ndebele et al.313 conjugated GQDs or N-GQDs (via π stacking) with tris(4-tert-butylphenoxy)-(5-phenoxylpicolinic acid)phthalocyanato cobalt(II) (CoPc) to obtain two electrocatalyst materials, namely, CoPc/GQDs and CoPc/N-GQDs, respectively, for the chronoamperometric (ChAmp)-based EC detection of NO2 with improved LODs in comparison to previous reports (Tables 12 and S10). The GQDs/N-GQDs significantly magnified the electrocatalytic activity of CoPc for a better EC response. The electrode materials were also tested for 20 consecutive cycles (≤10% reduction in peak currents) to access their satisfactory stability.

Summary: The existence of nitrogen-containing functional groups in doped-GQDs is favourable to interact with NO2 and execute electron transfer for the quenching of their fluorescence signal. Although the N,P-GQDs achieved the nanomolar-level detection of NO2, the achievement of a wide sensitivity range with N-GQDs is noticeable. The sensitivity in the EC detection of NO2 using conjugate systems (GQDs or N-GQDs with CoPc counterparts) is also considerable. GQDs-based/involved systems can effectively sense NO2 through the FL as well as EC method.

7.7. Halide ions

7.7.1. I. The initial detection of I via the FL turn-off-on approach using the N-GQDs/Ag+ system (Table S10) showed the possibility to utilize GQDs-based systems for this purpose. As a result, the sensing of I was performed again with an entirely new flow-injection CL (FI-CL) method (Table 12). The redox reaction between N,S-GQDs (nitrogen/sulfur content: 5.31/5.12%) and Ce4+ under acidic pH produced a strong CL signal, which was quenched by I and exhibited a good LOD (4.23 nM) in the detection process. The production of the redox reaction energy required for the excitation of N,S-GQDs facilitated the generation of a strong CL response. Moreover, this system showed applicability to determine I in kelp/tea samples. However, although the detection of I via the FI-CL approach resulted in a good performance, its complicated setup and precise sensing operation cannot be ignored.192
7.7.2. Fluoride (F). Besides the turn-off-on-based FL method to detect F using the B,N-GQDs/Hg2+ platform with low sensitivity (Table S10), Li et al.314 developed a highly sensitive ultra-low-field NMR (ULF-NMR, 118 µT) relaxometry method to sense F with a low LOD of 10 nM in a wide LR (Table 12). An aqueous solution of magnetic GQDs (Gd3+-loaded PEG–GQDs) showed a decrease in relaxation time with an increase in the concentration of F. Here, the probe-F-selective coordination and localized polarization/ionization effect of the surrounding water facilitated the proton exchange rate and relaxivity in the sensing operation, which is evidenced by the large negative electrostatic potential of the probe in the presence of F (−304.1 kJ mol−1 with F and −52.7 kJ mol−1 without F). Moreover, this sensor probe was found to be reliable for the analysis of F in domestic water samples.

Summary: Doped-GQDs, particularly, N-GQDs and N,S-GQDs have shown potential to detect I with considerable selectivity/sensitivity through the FL and FI-CL approach, respectively. Between these two sensing methods, the FL turn-off-on route is straightforward and user friendly. The selective and reasonably sensitive identification of F through the NMR relaxometry technique by utilizing the magnetic nature of Gd3+-loaded PEG–GQDs is recognizable.

7.8. SCN

After the qualitative demonstration of the recovery of fluorescence in the presence of SCN using a quenched system (GQDs/Hg2+) (Table S10), the quantitative detection of SCN was realized with a GQDs/AuNPs hybrid. The oxidation of TMB by scavenging more ˙OH radicals and growth of small-sized AuNPs in the presence of SCN collectively boosted the peroxidise-like nanozymatic activity to intensify the blue colour (from TMBox), and therefore COL sensing of SCN with a low LOD (3 nM) in a narrow LR (Table 12). Moreover, the detection of SCN in spiked-milk samples showed a convenient LOD of 8 nM and recoveries/RSDs of 93.8–104.8%/<4.0%.315

Summary: The COL detection of SCN with GQDs/AuNPs hybrids is a considerable sensing activity, which dictates the utility of GQDs involved systems in the selective-sensitive detection of SCN.

7.9. Other anions

GQDs-based systems have also been used for the sensing of persulfate (S2O82−), sulfite (SO32−), ferricyanide (Fe(CN)63−), ethyl xanthate (EtX), thiosulfate (S2O32−), peroxynitrite (ONOO), and bisulfite (HSO3) anions (Tables 12 and S10). IL@GQDs showed selective fluorescence quenching with Fe(CN)63−, and consequently Fe(CN)63− detection capability in a standard aqueous solution as well as in real river water samples (Table S10). The anion exchange nature of the IL (contains BF4) favoured the ferricyanide anion to reach close to the probe for effective interaction and quenching-based analytical activity. However, the requirement of a long equilibrium time (30 min; for significant quenching) and interference effect from Fe3+ (suggested to manage by AA) are some drawbacks in this sensing approach. The association of IFE suppressed the fluorescence intensity of N-GQDs with I2; however, the iodimetry reaction between S2O32− and I2 caused weakening of IFE to switch-on the fluorescence signal. As a result, this probe could achieve a low LOD for S2O32− (0.093 nM) in a narrow LR (Table 12) and practical utility in real water/biological fluid (human blood, serum, and urine) samples as well as in RAW cells. However, although this probe is reusable for at least up to 7 repeated cycles without much compromise in its selectivity-sensitivity, the interferences from S2− and SO32− cannot be neglected.316 FRET between N-GQDs and cyanine 5.5 (Cy5.5) in the covalently functionalized Cy5.5–N-GQDs, and furthermore the disappearance of FRET after the addition of ONOO built a ratiometric probe for the sensitive detection of ONOO (Table 12). However, the selectivity result of the sensor is missing in this report. This probe could be internalized in the mitochondria of living cells (Cy5.5 possesses mitochondria-targeting groups) to ratio metrically image exogenous as well as endogenous ONOO.317

Subsequently, a similar type of ratiometric detection was realized for HSO3 (LOD: 36 nM, Table 12) using a nanoconjugate system containing a red-emissive dye (PPy-Br, λem: 750 nm) and green-emissive OH-GQDs (λem: 535 nm). Fig. 36a illustrates the sensing process, where the FRET between the dye molecules and OH-GQDs enhanced/diminished the corresponding red/green fluorescence, and furthermore switch-off (due to the formation of non-fluorescent Michael adducts)/switch-on (termination of FRET) behaviour in the presence of HSO3. The FRET-involved sensing mechanism was validated by time-resolved fluorescence spectra (reduction of OH-GQDs τav from 6.92 to 3.25 ns after conjugation with PPy-Br, and further increase to 6.48 ns after the addition of HSO3, Fig. 36b). Moreover, the good recovery range of 95–100% in biological samples and ratiometric monitoring of HSO3 in MDA-MB-231 living cells confirmed the practical utility of this sensor.141


image file: d5ra04935k-f36.tif
Fig. 36 (a) Schematic of the ratiometric detection of HSO3 using the OH-GQDs/PPy-Br FRET conjugate system. (b) Time-resolved fluorescence decay curves of OH-GQDs, OH-GQDs/PPy-Br, and OH-GQDs/PPy-Br in the presence of HSO3. Reprinted (adapted) with permission from ref. 141, copyright 2023, the American Chemical Society.

Summary: Undoped GQDs/N-GQDs have shown potential to detect many other inorganic anions. Specifically, the FRET–weakened FRET-type FL turn-off-on process in a ratiometric manner (with GQDs or N-GQDs) may result the sensitive detection of ONOO/HSO3 anions, even up to the living cell tracking level. The high selectivity/sensitivity in the turn-off detection of Fe(CN)63− using IL@GQDs is also a considerable achievement.

8. Conclusion and future prospects

A variety of readily available precursors (from bulk carbon to common laboratory chemicals/biomass) has made it viable to attain the facile and low-cost synthesis of GQDs and their modified-derivatives. The experimental parameters, starting precursors, and post-modifications played crucial roles in attaining various physiochemical characteristics, compositional/structural modulations, and tuneable photo-physical properties in the GQDs/modified-GQDs. Consequently, GQDs/modified-GQDs have gained the necessary enhancement in optical features, electronic conductivity, and electrocatalytic activity for analytical purposes. Microplasma (at room temperature) and atmospheric pressure reflux synthesis are representative examples that showed feasibility to convert green precursors (such as chitosan, starch, and lignin) and economical aromatic compounds (phloroglucinol; nearly 100% conversion) into high-quality GQDs/doped-GQDs. The presence of nitrogen-heteroatoms in the lattice as well as in the form of amino/amido functional groups at the edge/surface of GQDs is found to be supportive for the nanoprobing of cationic MIs. Moreover, some amount of sulfur or boron along with nitrogen is also favourable to prepare N,S-GQDs or B,N-GQDs as effective probes for MIs.

Benefitting from their favourable features, along with limitations, GQDs/modified-GQDs and many of the platforms that involve GQDs/modified-GQDs have been shown as examples to selectively and sensitively detect many types of MIs and anions (brief summary presented after the discussion of each ions). Broadly speaking, electron density/functional groups enriched N-GQDs and N,S-GQDs have shown significant opportunities to selectively interact with many positively charged MIs, resulting in a sensitive detection performance. The sulfur-containing functional groups in GQDs can have specific affinity with Hg2+ and Pb2+. Also, additional functional moieties on functionalized GQDs such as DMC, Am, DA and crown ether exhibited specificity in the sensitive quantification of Hg2+, Cu2+, Ca2+, and alkali MIs, respectively. Conversely, negatively charged anions are frequently detected using GQDs-based/involved systems through a mediation process. GQDs-based/involved systems not only specifically interacted with inorganic ions to alter their inherent fluorescence signals (FL-based detection) but also emerged as a superior platform for EC, ECL, and COL detection. The sensing of inorganic ions through ratiometric design exhibited better sensitivity, reliability, and accuracy compared to analyzing via single-signal (emission or ECL response) upturn/downfall. Recent developments have enabled the quantification of hazardous HMIs such as Hg2+, Cd2+ and Pb2+ using GQDs-based platforms to sub-nanomolar/picomolar magnitudes, and even in the presence of other interfering analytes. Moreover, the detection of HMIs by fabricated test-paper strips, hydrogel kits, and aerogel pellets through simple smart-phone captured image analyses, assimilation of LDA/HCA to distinguish multiple HMIs, selectivity/tracking of bio-relevant alkali/alkaline-earth MIs in living cells, realization of worthwhile selectivity-sensitivity for rare-earth/radioactive MIs, integration of ML for the accomplishment of nanomolar-level accuracy, and sensing capability of various anions are some of the notable results.

However, although the sensing of inorganic ions by GQDs-based/involved systems has gained significant attention in the past, it is still in the developing stage, opening an opportunity for new and deeper-level research. Some of the foreseeable challenges/prospects where attention may be paid for a bright future are as follows:

(1) Obtaining nearly monodispersed GQDs with precise control over the number of layers and chemical composition is vital for understanding their structure–property relationship. Here, the choice of precursors and fine-tuning of the experimental conditions need further improvement, especially through the bottom-up approach.

(2) There are a few examples where red-/NIR-emissive or UCPL-featured GQDs/modified-GQDs have been explored for the detection of inorganic ions. Considering the importance of these GQD structures (particularly, owing to their high brightness, easy penetration capability in biological components, and low background effect), their easy/repeatable production with sufficiently high QY may be a suitable choice especially for application in living systems.

(3) It is surprising why GQDs/modified-GQDs with similar compositions and structures (synthesized using different processes/conditions/precursors) show selectivity with different HMIs and a range of sensitivity metrics with a particular HMI. Therefore, much work is required to divulge the genesis of the selectivity and sensitivity of GQDs-based systems with molecular-level observations, and also with the assistance of advanced algorithms-artificial intelligence.

(4) Surface functionalities, heteroatom-doping in their lattice, and active surface of GQDs are crucial for their selective interactions with inorganic ions. Moreover, fine-tuning of the functional groups at the edge rather than on their surface or vice versa, controllable sp2/sp3 carbon content, tunable defects/vacancies in their structure, and incorporation of chirality features may give some fruitful insights about their specificity/sensitivity mechanism for inorganic ions.

(5) GQDs-based platforms have shown promising relevance for the selective/sensitive detection of bio-related alkali/alkaline-earth MIs. Recently, Ca2+ detection/monitoring in different cell lines via functional group (DA)-mediated uncommon turn-on fluorescence has prompted researchers to develop higher wavelength-emitting GQDs/modified-GQDs for the purpose of safe bio-implantation.

(6) The NMR-based relaxometry detection of K+ (alkali MI) and F (anion) inspired the exploration of a new sensing approach for inorganic ions. However, the employment of GQDs-containing probes with toxic/expensive Gd3+ cannot be ignored. Therefore, the fabrication of NMR-active sensor systems by creating a paramagnetic center in GQDs-based probes with a benign element may be an interesting task.

(7) It is worthwhile to combine GQDs/modified-GQDs with other functional counterparts (e.g., LDH, PB analogues, metal–organic frameworks, and covalent-organic frameworks) at the molecular level to improve their chemical/mechanical stability and optical–electronic properties for the fabrication of advanced and reliable sensor devices.

(8) The well-dispersed and in situ implantation of GQDs/modified-GQDs in an interlocked polymeric matrix (smart wearable hydrogels/aerogels) can furnish self-transportation channels for the selective/rapid migration of inorganic ions. Therefore, the development of highly porous and low-density platforms (with high stability, easy processability, degradability, and tolerance against harsh environments) for the sensing of inorganic ions is worthwhile and should be explored.

(9) According to the available literature, the detection of toxic Cr6+/Cr3+, As3+, alkali/alkaline-earth MIs, rare-earth/radioactive MIs, and many anions is at the very early state of investigation, and thus there is much hope in the search of GQDs-based/involved advanced platforms, especially suitable using the EC, ECL, and COL sensing methods.

(10) The simultaneous detection/discrimination of multiple HMIs using GQDs-based/involved systems is another area of expansion, which should pay special attention to meet the requirement of robust sensors in real complex systems. Here, ML- and LDA/HCA-enabled semiempirical quantification with the requirement of minimum experimental data is showing a new future direction.

(11) Most inorganic anions are detected by GQDs-based systems through the involvement of mediating steps (because of their similar surface charge). Therefore, post-functionalization of GQDs/doped-GQDs with different cationic moieties may enable them to directly interact with selective anions and sense them with a high performance output.

Conflicts of interest

There are no conflicts to declare for financial interests or personal relationships.

Data availability

Data availability statement is not applicable as no new data were generated or analyzed in this study.

Supplementary information is available. See DOI: https://doi.org/10.1039/d5ra04935k.

Acknowledgements

PD thanks to the University of Allahabad, Prayagraj, India, for the infrastructural facility and Science & Engineering Board (SERB), New Delhi, India, for the financial support through a fast track grants (SB/FT/CS-190/2011). He also thanks the Government of India for “One Nation One Subscription” facility by which majority of research articles are accessible and Mr Shishir Singh, IIT Kanpur for providing some unsubscribed contents used in this article.

References

  1. M. Balali-Mood, K. Naseri, Z. Tahergorabi, M. R. Khazdair and M. Sadeghi, Toxic mechanisms of five heavy metals: mercury, lead, chromium, cadmium, and arsenic, Front. Pharmacol, 2021, 12, 643972 CrossRef CAS PubMed.
  2. O. Dagdag, T. W. Quadri, R. Haldhar, S. C. Kim, W. Daoudi, E. Berdimurodov, E. D. Akpan and E. E. Ebenso, An overview of heavy metal pollution and control, Heavy Metals in the Environment: Management Strategies for Global Pollution, 2023, ch. 1, pp. 3–24 Search PubMed.
  3. V. Singh, G. Ahmed, S. Vedika, P. Kumar, S. K. Chaturvedi, S. N. Rai, E. Vamanu and A. Kumar, Toxic heavy metal ions contamination in water and their sustainable reduction by eco-friendly methods: isotherms, thermodynamics and kinetics study, Sci. Rep., 2024, 14, 7595 CrossRef CAS.
  4. K. Jomova, S. Y. Alomar, E. Nepovimova, K. Kuca and M. Valko, Heavy metals: toxicity and human health effects, Arch. Toxicol., 2025, 99, 153–209 CrossRef CAS PubMed.
  5. T. E. Oladimeji, M. Oyedemi, M. E. Emetere, O. Agboola, J. B. Adeoye and O. A. Odunlami, Review on the impact of heavy metals from industrial wastewater effluent and removal technologies, Heliyon, 2024, 10, e40370 CrossRef CAS.
  6. S. Mitra, A. J. Chakraborty, A. M. Tareq, T. B. Emran, F. Nainu, A. Khusro, A. M. Idris, M. U. Khandaker, H. Osman, F. A. Alhumaydhi and J. Simal-Gandara, Impact of heavy metals on the environment and human health: novel therapeutic insights to counter the toxicity, J. King Saud Univ., Sci., 2022, 34, 101865 CrossRef.
  7. Y. S. Wu, A. I. Osman, M. Hosny, A. M. Elgarahy, A. S. Eltaweil, D. W. Rooney, Z. Chen, N. S. Rahim, M. Sekar, S. C. B. Gopinath, N. N. I. M. Rani, K. Batumalaie and P. S. Yap, The toxicity of mercury and its chemical compounds: molecular mechanisms and environmental and human health implications: a comprehensive review, ACS Omega, 2024, 9, 5100–5126 CrossRef CAS.
  8. M. Samuel Collin, S. K. Venkatraman, N. Vijayakumar, V. Kanimozhi, S. M. Arbaaz, R. G. S. Stacey, J. Anusha, R. Choudhary, V. Lvov, G. I. Tovar, F. Senatov, S. Koppala and S. Swamiappan, Bioaccumulation of lead (Pb) and its effects on human: a review, J. Hazard. Mater. Adv., 2022, 7, 100094 Search PubMed.
  9. F. Qu and W. Zheng, Cadmium exposure: mechanisms and pathways of toxicity and implications for human health, Toxics, 2024, 12, 388 CrossRef CAS PubMed.
  10. S. Y. Ganie, D. Javaid, Y. A. Hajam and M. S. Reshi, Arsenic toxicity: sources, pathophysiology and mechanism, Toxicol. Res., 2024, 13, 1–20 Search PubMed.
  11. J. P. Wise Jr, J. L. Young, J. Cai and L. Cai, Current understanding of hexavalent chromium [Cr(VI)] neurotoxicity and new perspectives, Environ. Int., 2022, 158, 106877 CrossRef PubMed.
  12. J. Chem. Educ. Staff, Biochemical roles of some essential metal ions, J. Chem. Educ., 1977, 54, 761–762 CrossRef.
  13. M. A. Zoroddu, J. Aaseth, G. Crisponi, S. Medici, M. Peana and V. M. Nurchi, The essential metals for humans: a brief overview, J. Inorg. Biochem., 2019, 195, 120–129 CrossRef CAS.
  14. E. A. B. Hughes, T. E. Robinson, D. B. Bassett, S. C. Cox and L. M. Grover, Critical and diverse roles of phosphates in human bone formation, J. Mater. Chem. B, 2019, 7, 7460–7470 RSC.
  15. E. S. Rudge, A. H. Y. Chan and F. J. Leeper, Prodrugs of pyrophosphates and bisphosphonates: disguising phosphorus oxyanions, RSC Med. Chem., 2022, 13, 375–391 RSC.
  16. G. Qiu, Y. Han, X. Zhu, J. Gong, T. Luo, C. Zhao, J. Liu, J. Liu and X. Li, Sensitive detection of sulfide ion based on fluorescent ionic liquid–graphene quantum dots nanocomposite, Front. Chem., 2021, 9, 658045 CrossRef CAS.
  17. L. Singh and N. Ranjan, Highly selective and sensitive detection of nitrite ion by an unusual nitration of a fluorescent benzimidazole, J. Am. Chem. Soc., 2023, 145, 2745–2749 CrossRef CAS.
  18. M. Maruthupandi, M. Chandhru, S. K. Rani and N. Vasimalai, Highly selective detection of iodide in biological, food, and environmental samples using polymer-capped silver nanoparticles: preparation of a paper-based testing kit for on-site monitoring, ACS Omega, 2019, 4, 11372–11379 CrossRef CAS PubMed.
  19. R. Saha, I. Bhattacharya, S. Pyne and R. K. Mitra, Thiocyanate ion (SCN) offers a major impact in rapid protein amyloidosis: a salient role played by protein solvation, J. Phys. Chem. B, 2025, 129, 1946–1955 CrossRef CAS.
  20. WHO Team, Guidelines for drinking-water quality: fourth edition incorporating the first and second addenda, World Health Organization, 2022, https://iris.who.int/bitstream/handle/10665/352532/9789240045064-eng.pdf Search PubMed.
  21. National Primary Drinking Water Regulations, EPA, https://www.epa.gov/ground-water-and-drinking-water/national-primary-drinking-water-regulations, 12 December 2024 Search PubMed.
  22. F. Tamirat, W. D. Adane, M. Tessema, E. Tesfaye and G. Tesfaye, Determination of major and trace metals in date palm fruit (Phoenix dactylifera) samples using flame atomic absorption spectrometry and assessment of the associated public health risks, Int. J. Anal. Chem., 2024, 2024, 9914300 Search PubMed.
  23. D. Rajendiran, N. Harikrishnan and K. Veeramuthu, Heavy metal concentrations and pollution indicators in the Ennore ecosystem, east coast of Tamilnadu, India using atomic absorption spectrometry study with statistical approach, Sci. Rep., 2025, 15, 9161 CrossRef CAS.
  24. E. Bozorgzadeh, A. Pasdaran and H. Ebrahimi-Najafabadi, Determination of toxic heavy metals in fish samples using dispersive micro solid phase extraction combined with inductively coupled plasma optical emission spectroscopy, Food Chem., 2021, 346, 128916 CrossRef CAS PubMed.
  25. J. Zhong, Z. Wang, Y. Chen, W. Huan, M. Shi, L. Lei, X. Yu and L. Chen, Determination of trace heavy metal elements in litterfall by inductively coupled plasma optical emission spectrometry after extraction using choline chloride-based deep eutectic solvents, RSC Adv., 2024, 14, 22497–22503 RSC.
  26. T. V. Acker, S. Theiner, E. Bolea-Fernandez, F. Vanhaecke and G. Koellensperger, Inductively coupled plasma mass spectrometry, Nat. Rev. Methods Primers, 2023, 3, 52 CrossRef.
  27. V. Balaram, L. Copia, U. S. Kumar, J. Miller and S. Chidambaram, Pollution of water resources and application of ICP-MS techniques for monitoring and management-a comprehensive review, Geosystems Geoenviron., 2023, 2, 100210 CrossRef.
  28. K. Pytlakowska, K. Kocot, B. Hachuła, M. Pilch, R. Wrzalik and M. Zubko, Determination of heavy metal ions by energy dispersive X-ray fluorescence spectrometry using reduced graphene oxide decorated with molybdenum disulfide as solid adsorbent, Spectrochim. Acta, Part B, 2020, 167, 105846 CrossRef CAS.
  29. V. G. Povarov, T. N. Kopylova, M. A. Sinyakova and V. A. Rudko, Quantitative determination of trace heavy metals and selected rock-forming elements in porous carbon materials by the X-ray fluorescence method, ACS Omega, 2021, 6, 24595–24601 CrossRef CAS PubMed.
  30. J. Li, W. Peng, A. Wang, M. Wan, Y. Zhou, X. G. Zhang, S. Jin and F. L. Zhang, Highly sensitive and selective SERS substrates with 3D hot spot buildings for rapid mercury ion detection, Analyst, 2023, 148, 4044–4052 RSC.
  31. A. H. A. Hassan, M. M. A. Zeinhom, M. Shaban, A. M. Korany, A. Gamal, N. S. Abdel-Atty and S. I. Al-Saeedi, Rapid and sensitive in situ detection of heavy metals in fish using enhanced Raman spectroscopy, Spectrochim. Acta, Part A, 2024, 313, 124082 CrossRef CAS.
  32. K. Kancharla and K. K. Tadi, MoS2 quantum dots-based low-cost and disposable electrochemical sensor for the detection of lead (II) ions, J. Electrochem. Soc., 2023, 170, 127509 CrossRef CAS.
  33. F. Farahmandzadeh, K. Kermanshahian, E. Molahosseini, M. Molaei and M. Karimipour, Highly fluorescent CdTe/ZnSe quantum dot-based “turn-off” sensor for the on-site rapid detection of lead ions in aqueous medium, Spectrochim. Acta, Part A, 2025, 324, 124914 CrossRef CAS PubMed.
  34. R. Guan, L. Tao, Y. Hu, C. Zhang, Y. Wang, M. Hong and Q. Yue, Selective determination of Ag+ in the presence of Cd2+, Hg2+ and Cu2+ based on their different interactions with gold nanoclusters, RSC Adv., 2020, 10, 33299–33306 RSC.
  35. K. Zhao, L. Ge, T. I. Wong, X. Zhou and G. Lisak, Gold-silver nanoparticles modified electrochemical sensor array for simultaneous determination of chromium(III) and chromium(VI) in wastewater samples, Chemosphere, 2021, 281, 130880 CrossRef PubMed.
  36. X. Shao, D. Yang, M. Wang and Q. Yue, A colorimetric detection of Hg2+ based on gold nanoparticles synthesized oxidized N-methylpyrrolidone as a reducing agent, Sci. Rep., 2023, 13, 22208 CrossRef.
  37. J. Liu, Y. Bai, J. Shi, Q. Yu, J. Liu, J. Yang, C. Fu and Q. Zhang, Selective detection of mercury ions based on tin oxide quantum dots: performance and fluorescence enhancement model, J. Mater. Chem. C, 2021, 9, 8274–8284 RSC.
  38. G. A. El-Fatah, H. S. Magar, R. Y. A. Hassan, R. Mahmoud, A. A. Farghali and M. E. M. Hassouna, A novel gallium oxide nanoparticles-based sensor for the simultaneous electrochemical detection of Pb2+, Cd2+ and Hg2+ ions in real water samples, Sci. Rep., 2022, 12, 20181 CrossRef PubMed.
  39. Z. Ali, R. Ullah, M. Tuzen, S. Ullah, A. Rahim and T. A. Saleh, Colorimetric sensing of heavy metals on metal doped metal oxide nanocomposites: a review, Trends Environ. Anal. Chem., 2023, 37, e00187 CrossRef.
  40. R. Ayranci and M. Ak, An electrochemical sensor platform for sensitive detection of iron (III) ions based on pyrene-substituted poly(2,5-dithienylpyrrole), J. Electrochem. Soc., 2019, 166, B291–B296 CrossRef.
  41. R. Iftikhar, I. Parveen, Ayesha, A. Mazhar, M. S. Iqbal, G. M. Kamal, F. Hafeez, A. L. Pang and M. Ahmadipour, Small organic molecules as fluorescent sensors for the detection of highly toxic heavy metal cations in portable water, J. Environ. Chem. Eng., 2023, 11, 109030 CrossRef.
  42. M. Núñez-Martínez, M. Fernández-Míguez, E. Quiñoá and F. Freire, Colorimetric detection of oxidizing metal ions using anilide-poly(phenylacetylene)s, Nanoscale, 2025, 17, 4439–4443 RSC.
  43. Z. Yang, T. Xu, H. Li, M. She, J. Chen, Z. Wang, S. Zhang and J. Li, Zero-dimensional carbon nanomaterials for fluorescent sensing and imaging, Chem. Rev., 2023, 123, 11047–11136 CrossRef PubMed.
  44. L. A. Ponomarenko, F. Schedin, M. I. Katsnelson, R. Yang, E. W. Hill, K. S. Novoselov and A. K. Geim, Chaotic dirac billiard in graphene quantum dots, Science, 2008, 320, 356–358 CrossRef PubMed.
  45. S. Zhu, Y. Song, J. Wang, H. Wan, Y. Zhang, Y. Ning and B. Yang, Photoluminescence mechanism in graphene quantum dots: quantum confinement effect and surface/edge state, Nano Today, 2017, 13, 10–14 CrossRef.
  46. A. B. Siddique, S. M. Hossain, A. K. Pramanick and M. Ray, Excitation dependence and independence of photoluminescence in carbon dots and graphene quantum dots: insights into the mechanism of emission, Nanoscale, 2021, 13, 16662–16671 RSC.
  47. M. J. Sweetman, S. M. Hickey, D. A. Brooks, J. D. Hayball and S. E. Plush, A practical guide to prepare and synthetically modify graphene quantum dots, Adv. Funct. Mater., 2019, 29, 1808740 CrossRef.
  48. B. Li, Y. Wang, L. Huang, H. Qu, Z. Han, Y. Wang, M. J. Kipper, L. A. Belfiore and J. Tang, Review of performance improvement strategies for doped graphene quantum dots for fluorescence-based sensing, Synth. Met., 2021, 276, 116758 CrossRef.
  49. R. Rabeya, S. Mahalingam, A. Manap, M. Satgunam, M. Akhtaruzzaman and C. H. Chia, Structural defects in graphene quantum dots: a review, Int. J. Quantum Chem., 2022, 122, e26900 CrossRef.
  50. T. Gao, X. Wang, L. Y. Yang, H. He, X. X. Ba, J. Zhao, F. L. Jiang and Y. Liu, Red, yellow, and blue luminescence by graphene quantum dots: syntheses, mechanism, and cellular imaging, ACS Appl. Mater. Interfaces, 2017, 9, 24846–24856 CrossRef CAS PubMed.
  51. L. Chen, S. Yang, Y. Li, Z. Liu, H. Wang, Y. Zhang, K. Qi, G. Wang, P. He and G. Ding, Precursor symmetry triggered modulation of fluorescence quantum yield in graphene quantum dots, Adv. Funct. Mater., 2024, 34, 2401246 CrossRef CAS.
  52. A. Poszwa, Electron transport properties of graphene quantum dots with non-centro-symmetric Gaussian deformation, Sci. Rep., 2022, 12, 9908 CrossRef CAS.
  53. Y. Y. Ju, X. X. Shi, S. Y. Xu, X. H. Ma, R. J. Wei, H. Hou, C. C. Chu, D. Sun, G. Liu and Y. Z. Tan, Atomically precise water-soluble graphene quantum dot for cancer sonodynamic therapy, Adv. Sci., 2022, 9, 2105034 CrossRef CAS.
  54. A. Raghavan, M. Radhakrishnan, K. Soren, P. Wadnerkar, A. Kumar, S. Chakravarty and S. Ghosh, Biological evaluation of graphene quantum dots and nitrogen-doped graphene quantum dots as neurotrophic agents, ACS Appl. Bio Mater., 2023, 6, 2237–2247 CrossRef CAS.
  55. Z. Song, J. Gong, R. Soltani, J. D. Fauny, C. Ménard-Moyon, P. Chen and A. Bianco, Cellular impact and biodegradability of S- and N-doped graphene quantum dots on human monocytes and macrophages, Adv. Funct. Mater., 2024, 34, 2405856 CrossRef CAS.
  56. H. Wang, R. Firouzi-Haji, M. Aghajamali, M. A. Vieira, J. Y. Cho, Q. Lu, X. Zhang, A. J. Bergren, J. G. C. Veinot, H. Hassanzadeh and A. Meldrum, Graphene quantum dot bearing liquid droplets for ultrasensitive fluorescence-based detection of nitroaromatics, ACS Appl. Nano Mater., 2022, 5, 14639–14645 CrossRef CAS.
  57. P. A. Rasheed, M. Ankitha, V. K. Pillai and S. Alwarappan, Graphene quantum dots for biosensing and bioimaging, RSC Adv., 2024, 14, 16001–16023 RSC.
  58. P. Sharma, P. Yadav, A. Kumar and H. Mudila, Exploration of graphene quantum dots: design, properties, energy storage and conversion, J. Power Sources, 2025, 630, 236177 CrossRef CAS.
  59. A. Zarepour, A. Khosravi, N. Y. Ayten, P. C. Hatır, S. Iravani and A. Zarrabi, Innovative approaches for cancer treatment: graphene quantum dots for photodynamic and photothermal therapies, J. Mater. Chem. B, 2024, 12, 4307–4334 RSC.
  60. C. Zhao, X. Song, Y. Liu, Y. Fu, L. Ye, N. Wang, F. Wang, L. Li, M. Mohammadniaei, M. Zhang, Q. Zhang and J. Liu, Synthesis of graphene quantum dots and their applications in drug delivery, J. Nanobiotechnol., 2020, 18, 142 CrossRef CAS.
  61. M. Goswami, S. Mandal and V. K. Pillai, Efect of hetero-atom doping on the electrocatalytic properties of graphene quantum dots for oxygen reduction reaction, Sci. Rep., 2023, 13, 5182 CrossRef CAS.
  62. X. Liu, N. Li, X. Lv, J. Wang, Q. Ma and Z. Xie, One-step preparation of N, S-doped graphene quantum dots for white light-emitting diodes, J. Alloys Compd., 2025, 1022, 179827 CrossRef CAS.
  63. S. Zhou, H. Xu, W. Gan and Q. Yuan, Graphene quantum dots: recent progress in preparation and fluorescence sensing applications, RSC Adv., 2016, 6, 110775–110788 RSC.
  64. L. Zhang, D. Peng, R. P. Liang and J. D. Qiu, Graphene-based optical nanosensors for detection of heavy metal ions, Trends Anal. Chem., 2018, 102, 280–289 CrossRef CAS.
  65. M. Li, T. Chen, J. J. Gooding and J. Liu, Review of carbon and graphene quantum dots for sensing, ACS Sens., 2019, 4, 1732–1748 CrossRef CAS.
  66. N. A. A. Anas, Y. W. Fen, N. A. S. Omar, W. M. E. M. M. Daniyal, N. S. M. Ramdzan and S. Saleviter, Development of graphene quantum dots-based optical sensor for toxic metal ion detection, Sensors, 2019, 19, 3850 CrossRef CAS.
  67. I. A. Revesz, S. M. Hickey and M. J. Sweetman, Metal ion sensing with graphene quantum dots: detection of harmful contaminants and biorelevant species, J. Mater. Chem. B, 2022, 10, 4346–4362 RSC.
  68. R. Wu, Y. Cao, Z. Chen and J. J. Zhu, Fluorescent graphene quantum dots: properties regulation, sensing applications, and future prospects, Adv. Sens. Energy Mater., 2025, 4, 100140 CrossRef.
  69. S. Saisree, A. S. Nair, E. Dais and K. Y. Sandhya, Electrochemical sensors for monitoring water quality: recent advances in graphene quantum dot-based materials for the detection of toxic heavy metal ions Cd(II), Pb(II) and Hg(II) with their mechanistic aspects, J. Environ. Chem. Eng., 2025, 13, 116545 CrossRef.
  70. L. Lu, Y. Zhu, C. Shi and Y. T. Pei, Large-scale synthesis of defects-selective graphene quantum dots by ultrasonic-assisted liquid-phase exfoliation, Carbon, 2016, 109, 373–383 CrossRef CAS.
  71. Y. Zhou, H. Sun, F. Wang, J. Ren and X. Qu, How functional groups influence the ROS generation and cytotoxicity of graphene quantum dots, Chem. Commun., 2017, 53, 10588–10591 RSC.
  72. C. Zhang, J. Li, X. Zeng, Z. Yuan and N. Zhao, Graphene quantum dots derived from hollow carbon nano-onions, Nano Res., 2018, 11, 174–184 CrossRef CAS.
  73. S. Kapoor, A. Jha, H. Ahmad and S. S. Islam, Avenue to large-scale production of graphene quantum dots from high-purity graphene sheets using laboratory-grade graphite electrodes, ACS Omega, 2020, 5, 18831–18841 CrossRef CAS.
  74. Y. X. Chen, D. Lu, G. G. Wang, J. Huangfu, Q. B. Wu, X. F. Wang, L. F. Liu, D. M. Ye, B. Yan and J. Han, Highly efficient orange emissive graphene quantum dots prepared by acid-free method for white LEDs, ACS Sustainable Chem. Eng., 2020, 8, 6657–6666 CrossRef CAS.
  75. M. O. Danilov, S. S. Fomanyuk, G. I. Dovbeshko, O. P. Gnatyuk, I. A. Rusetskyi and G. Ya. Kolbasov, Graphene quantum dots from partially unzipped multi-walled carbon nanotubes: promising materials for oxygen electrodes, J. Electrochem. Soc., 2021, 168, 044514 CrossRef CAS.
  76. I. Hazarika, T. Kalita, P. Deka, S. K. Gogoi, K. Althubeiti, R. Thakuria and B. Gogoi, Soot-based reduced graphene quantum dot/hemin conjugate for favipiravir sensing, ACS Appl. Nano Mater., 2021, 4, 13927–13937 CrossRef CAS.
  77. W. Lu, N. Shen, C. Celia, Y. Xie, Q. Chang and X. Deng, Polyethylenimine-functionalized graphene quantum dots for Cd2+ ion adsorption, New J. Chem., 2023, 47, 20966–20975 RSC.
  78. N. S. Singh and P. K. Giri, Ultrasmall graphene quantum dots as exceptionally stable and efficient substrates for graphene-enhanced Raman spectroscopy detection of dyes, ACS Appl. Nano Mater., 2024, 7, 13579–13589 CrossRef.
  79. G. S. Lekshmi, A. Krzeminska, S. Sundararaju, S. J. Hinder, A. Zatylna, P. Paneth, J. Pietrasik, C. Sudip, W. Hendrickx, A. J. Nathanael, B. Januszewicz, L. Kolodziejczyk, L. Kaczmarek and V. Kumaravel, Engineering of brewery waste-derived graphene quantum dots with ZnO nanoparticles for treating multi-drug resistant bacterial infections, J. Environ. Chem. Eng., 2024, 12, 112263 CrossRef.
  80. Q. Hu, K. Zhao, M. Liu, S. Riaz, Y. Qi, P. Wei, J. Cheng and Y. Xie, A dual passivation strategy based on F/N co-doped coal-based graphene quantum dots for high-efficiency carbon-based perovskite solar cells, J. Mater. Chem. A, 2024, 12, 5980–5989 RSC.
  81. C. Zhou, Y. Chen, X. You, Y. Dong and Y. Chi, An electrochemiluminescent biosensor based on interactions between a graphene quantum dot-sulfite co-reactant system and hydrogen peroxide, ChemElectroChem, 2017, 4, 1783–1789 CrossRef.
  82. M. Zhao, Direct synthesis of graphene quantum dots with different fluorescence properties by oxidation of graphene oxide using nitric acid, Appl. Sci., 2018, 8, 1303 CrossRef.
  83. S. Maiti, S. Kundu, C. N. Roy, T. K. Das and A. Saha, Synthesis of excitation independent highly luminescent graphene quantum dots through perchloric acid oxidation, Langmuir, 2017, 33, 14634–14642 CrossRef PubMed.
  84. B. Lyu, H. J. Li, F. Xue, L. Sai, B. Gui, D. Qian, X. Wang and J. Yang, Facile, gram-scale and eco-friendly synthesis of multi-color graphene quantum dots by thermal-driven advanced oxidation process, Chem. Eng. J., 2020, 388, 124285 CrossRef.
  85. X. Zhu, J. Yu, Y. Yan, W. Song and X. Hai, One-pot alkali cutting-assisted synthesis of fluorescence tunable amino-functionalized graphene quantum dots as a multifunctional nanosensor for sensing of pH and tannic acid, Talanta, 2022, 236, 122874 CrossRef PubMed.
  86. R. V. Nair, R. T. Thomas, V. Sankar, H. Muhammad, M. Dong and S. Pillai, Rapid, acid-free synthesis of high-quality graphene quantum dots for aggregation induced sensing of metal ions and bioimaging, ACS Omega, 2017, 2, 8051–8061 CrossRef.
  87. Y. Zhao, X. Wu, S. Sun, L. Ma, L. Zhang and H. Lin, A facile and high-efficient approach to yellow emissive graphene quantum dots from graphene oxide, Carbon, 2017, 124, 342–347 CrossRef.
  88. X. Zhou, S. Guo, P. Zhong, Y. Xie, Z. Li and X. Ma, Large scale production of graphene quantum dots through the reaction of graphene oxide with sodium hypochlorite, RSC Adv., 2016, 6, 54644–54648 RSC.
  89. Q. Xue, H. Huang, L. Wang, Z. Chen, M. Wu, Z. Li and D. Pan, Nearly monodisperse graphene quantum dots fabricated by amine-assisted cutting and ultrafiltration, Nanoscale, 2013, 5, 12098–12103 RSC.
  90. W. Kwon, Y. H. Kim, J. H. Kim, T. Lee, S. Do, Y. Park, M. S. Jeong, T. W. Lee and S. W. Rhee, High color-purity green, orange, and red light-emitting diodes based on chemically functionalized graphene quantum dots, Sci. Rep., 2016, 6, 24205 CrossRef PubMed.
  91. B. Ahmed, S. Kumar, A. K. Ojha, F. Hirsch, S. Riese and I. Fischer, Facile synthesis and photophysics of graphene quantum dots, J. Photochem. Photobiol., A, 2018, 364, 671–678 CrossRef.
  92. R. Das, K. K. Paul and P. K. Giria, Highly sensitive and selective label-free detection of dopamine in human serum based on nitrogen-doped graphene quantum dots decorated on Au nanoparticles: mechanistic insights through microscopic and spectroscopic studies, Appl. Surf. Sci., 2019, 490, 318–330 CrossRef.
  93. G. S. Kang, S. Lee, J. S. Yeo, E. S. Choi, D. C. Lee, S. I. Na and H. I. Joh, Graphene quantum dots with nitrogen and oxygen derived from simultaneous reaction of solvent as exfoliant and dopant, Chem. Eng. J., 2019, 372, 624–630 CrossRef CAS.
  94. Y. Yan, H. Li, Q. Wang, H. Mao and W. Kun, Controllable ionic liquid-assisted electrochemical exfoliation of carbon fibers for the green and large-scale preparation of functionalized graphene quantum dots endowed with multicolor emission and size tenability, J. Mater. Chem. C, 2017, 5, 6092–6100 RSC.
  95. K. Chu, J. R. Adsetts, S. He, Z. Zhan, L. Yang, J. M. Wong, D. A. Love and Z. Ding, Electrogenerated chemiluminescence and electroluminescence of N-doped graphene quantum dots fabricated from an electrochemical exfoliation process in nitrogen-containing electrolytes, Chem.–Eur. J., 2020, 26, 15892–15900 CrossRef CAS PubMed.
  96. H. Huang, S. Yang, Q. Li, Y. Yang, G. Wang, X. You, B. Mao, H. Wang, Y. Ma, P. He, Z. Liu, G. Ding and X. Xie, Electrochemical cutting in weak aqueous electrolyte: the strategy for efficient and controllable preparation of graphene quantum dots, Langmuir, 2018, 34, 250–258 CrossRef CAS.
  97. R. Qiang, W. Sun, K. Hou, Z. Li, J. Zhang, Y. Ding, J. Wang and S. Yang, Electrochemical trimming of graphene oxide affords graphene quantum dots for Fe3+ detection, ACS Appl. Nano Mater., 2021, 4, 5220–5229 CrossRef CAS.
  98. R. L. Calabro, D. S. Yang and D. Y. Kim, Controlled nitrogen doping of graphene quantum dots through laser ablation in aqueous solutions for photoluminescence and electrocatalytic applications, ACS Appl. Nano Mater., 2019, 2, 6948–6959 CrossRef CAS.
  99. S. Kang, Y. K. Jeong, J. H. Ryu, Y. Son, W. R. Kim, B. Lee, K. H. Jung and K. M. Kim, Pulsed laser ablation based synthetic route for nitrogen-doped graphene quantum dots using graphite flakes, Appl. Surf. Sci., 2020, 506, 144998 CrossRef CAS.
  100. M. Buzaglo, M. Shtein and O. Regev, Graphene quantum dots produced by microfluidization, Chem. Mater., 2016, 28, 21–24 CrossRef CAS.
  101. Y. Xu, J. Chang, C. Liang, X. Sui, Y. Ma, L. Song, W. Jiang, J. Zhou, H. Guo, X. Liu and Y. Zhang, Tailoring multi-walled carbon nanotubes into graphene quantum sheets, ACS Appl. Mater. Interfaces, 2020, 12, 47784–47791 CrossRef CAS.
  102. Z. Azimi, M. Alimohammadian and B. Sohrabi, Graphene quantum dots based on mechanical exfoliation methods: a simple and eco-friendly technique, ACS Omega, 2024, 9, 31427–31437 CrossRef CAS.
  103. M. Esmaeili, F. Ahour and S. Keshipour, Sensitive and selective determination of trace amounts of mercury ions using a dimercaprol functionalized graphene quantum dot modified glassy carbon electrode, Nanoscale, 2021, 13, 11403–11413 RSC.
  104. D. Huang, H. Zhou, Y. Wu, T. Wang, L. Sun, P. Gao, Y. Sun, H. Huang, G. Zhou and J. Hu, Bottom-up synthesis and structural design strategy for graphene quantum dots with tunable emission to near infrared region, Carbon, 2019, 142, 673–684 CrossRef CAS.
  105. L. Yang, C. R. De-Jager, J. R. Adsetts, K. Chu, K. Liu, C. Zhang and Z. Ding, Analyzing near-infrared electrochemiluminescence of graphene quantum dots in aqueous media, Anal. Chem., 2021, 93, 12409–12416 CrossRef CAS PubMed.
  106. M. K. Kumawat, M. Thakur, R. B. Gurung and R. Srivastava, Graphene quantum dots for cell proliferation, nucleus imaging, and photoluminescent sensing applications, Sci. Rep., 2017, 7, 15858 CrossRef.
  107. W. Li, M. Li, Y. Liu, D. Pan, Z. Li, L. Wang and M. Wu, Three-minute ultra-rapid microwave-assisted synthesis of bright fluorescent graphene quantum dots for live cell staining and white LEDs, ACS Appl. Nano Mater., 2018, 1, 1623–1630 CrossRef CAS.
  108. L. Wang, W. Li, B. Wu, Z. Li, D. Pan and M. Wu, Room-temperature synthesis of graphene quantum dots via electron-beam irradiation and their application in cell imaging, Chem. Eng. J., 2017, 309, 374–380 CrossRef CAS.
  109. J. Zhu, Y. Tang, G. Wang, J. Mao, Z. Liu, T. Sun, M. Wang, D. Chen, Y. Yang, J. Li, Y. Deng and S. Yang, Green, rapid, and universal preparation approach of graphene quantum dots under ultraviolet irradiation, ACS Appl. Mater. Interfaces, 2017, 9, 14470–14477 CrossRef CAS.
  110. S. H. Lee, D. Y. Kim, J. Lee, S. B. Lee, H. Han, Y. Y. Kim, S. C. Mun, S. H. Im, T. H. Kim and O. O. Park, Synthesis of single-crystalline hexagonal graphene quantum dots from solution chemistry, Nano Lett., 2019, 19, 5437–5442 CrossRef CAS PubMed.
  111. D. Kurniawan, N. Sharma, M. R. Rahardja, Y. Y. Cheng, Y. T. Chen, G. X. Wu, Y. Y. Yeh, P. C. Yeh, K. K. Ostrikov and W. H. Chiang, Plasma nanoengineering of bioresource-derived graphene quantum dots as ultrasensitive environmental nanoprobes, ACS Appl. Mater. Interfaces, 2022, 14, 52289–52300 CrossRef CAS PubMed.
  112. J. Yao and L. Wang, Graphene quantum dots as nanosensor for rapid and label-free dual detection of Cu2+ and tiopronin by means of fluorescence ‘‘on–off–on’’ switching: mechanism and molecular logic gate, New J. Chem., 2021, 45, 20649–20659 RSC.
  113. G. L. Hong, H. L. Zhao, H. H. Deng, H. J. Yang, H. P. Peng, Y. H. Liu and W. Chen, Fabrication of ultra-small monolayer graphene quantum dots by pyrolysis of trisodium citrate forfluorescent cell imaging, Int. J. Nanomed., 2018, 13, 4807–4815 CrossRef CAS PubMed.
  114. S. Thanomsak, C. Insombat, P. Chaiyo, T. Tuntulani and W. Janrungroatsakul, Fabrication of a paper-based sensor from graphene quantum dots coated with a polymeric membrane for the determination of gold(III) ions, Anal. Methods, 2021, 13, 4785–4792 RSC.
  115. M. Nafiujjaman, J. Kim, H. K. Park and Y. K. Lee, Preparation of blue-color-emitting graphene quantum dots and their in vitro and in vivo toxicity evaluation, J. Ind. Eng. Chem., 2018, 57, 171–180 CrossRef CAS.
  116. E. Sari, Synthesis and characterization of high quantum yield graphene quantum dots via pyrolysis of glutamic acid and aspartic acid, J. Nanopart. Res., 2022, 24, 37 CrossRef CAS.
  117. C. C. Fu, C. T. Hsieh, R. S. Juang, S. Gu, Y. A. Gandomi, R. E. Kelly and K. D. Kihm, Electrochemical sensing of mercury ions in electrolyte solutions by nitrogen-doped graphene quantum dot electrodes at ultralow concentrations, J. Mol. Liq., 2020, 302, 112593 CrossRef CAS.
  118. S. Gu, C. T. Hsieh, Y. Y. Tsai, Y. A. Gandomi, S. Yeom, K. D. Kihm, C. C. Fu and R. S. Juang, Sulfur and nitrogen co-doped graphene quantum dots as a fluorescent quenching probe for highly sensitive detection toward mercury ions, ACS Appl. Nano Mater., 2019, 2, 790–798 CrossRef CAS.
  119. L. Jin, L. Li, X. Zeng, S. Yu and J. Zhang, The ratiometric fluorescent sensor based on the mixture of CdTe quantum dots and graphene quantum dots for quantitative analysis of silver in drinks, Food Chem., 2023, 429, 136926 CrossRef CAS.
  120. T. T. Bezuneh, T. H. Fereja, H. Li and Y. Jin, Solid-phase pyrolysis synthesis of highly fluorescent nitrogen/sulfur codoped graphene quantum dots for selective and sensitive diversity detection of Cr(VI), Langmuir, 2023, 39, 1538–1547 CrossRef CAS PubMed.
  121. J. Cai, G. Han, J. Ren, C. Liu, J. Wang and X. Wang, Single-layered graphene quantum dots with self-passivated layer from xylan for visual detection of trace chromium(Vl), Chem. Eng. J., 2022, 435, 131833 CrossRef CAS.
  122. S. Ye, F. Su, J. Li, B. Yu, L. Xu, T. Xiong, K. Shao and X. Yuan, Enhanced in vivo antiviral activity against pseudorabies virus through transforming gallic acid into graphene quantum dots with stimulation of interferon-related immune responses, J. Mater. Chem. B, 2024, 12, 122–130 RSC.
  123. C. T. Hsieh, P. Y. Sung, Y. A. Gandomi, K. S. Khoo and J. K. Chang, Microwave synthesis of boron- and nitrogen-codoped graphene quantum dots and their detection to pesticides and metal ions, Chemosphere, 2023, 318, 137926 CrossRef CAS PubMed.
  124. R. K. Ratnesh, M. K. Singh, V. Kumar, S. Singh, R. Chandra, M. Singh and J. Singh, Mango leaves (Mangifera indica)-derived highly florescent green graphene quantum dot nanoprobes for enhanced on-off dual detection of cholesterol and Fe2+ ions based on molecular logic operation, ACS Appl. Bio Mater., 2024, 7, 4417–4426 CrossRef CAS.
  125. K. Ajravat, S. Rajput and L. K. Brar, Microwave assisted hydrothermal synthesis of N doped graphene for supercapacitor applications, Diamond Relat. Mater., 2022, 129, 109373 CrossRef CAS.
  126. D. Kurniawan, F. Caielli, K. Thyagajaran, K. (Ken) Ostrikov, W. H. Chiang and D. Z. Pai, Operando time and space-resolved liquid-phase diagnostics reveal the plasma selective synthesis of nanographenes, Nanoscale, 2024, 16, 15104–15112 RSC.
  127. Y. Ochi, A. Otani, R. Katakami, A. Ogura, K. I. Takao, Y. Iso and T. Isobe, Open system massive synthesis of narrow-band blue and green fluorescent graphene quantum dots and their application in water sensing, J. Mater. Chem. C, 2024, 12, 6548–6558 RSC.
  128. A. Anter, E. Orhan, M. Ulusoy, B. Polat, M. Yıldız, A. Kumar, A. D. Bartolomeo, E. Faella, M. Passacantando and J. Bi, Lanthanum(III)hydroxide nanoparticles and polyethyleneimine-functionalized graphene quantum dot nanocomposites in photosensitive silicon heterojunctions, ACS Appl. Mater. Interfaces, 2024, 16, 22421–22432 CrossRef CAS.
  129. J. Yuan, Z. Yang, J. Zou, Z. Wu, Z. Wang, L. Wang, W. Shen, Q. Zhang and H. Xu, Functionalized sulfonic acid groups enhance solubility and stability of graphene quantum dots for efficient photosynthesis of lettuce, J. Photochem. Photobiol., A, 2025, 463, 116280 CrossRef CAS.
  130. N. Nandi, S. Gaurav, P. Sarkar, S. Kumar and K. Sahu, Hit multiple targets with one arrow: Pb2+ and ClO detection by edge functionalized graphene quantum dots and their applications in living cells, ACS Appl. Bio Mater., 2021, 4, 7605–7614 CrossRef CAS PubMed.
  131. R. Cheng, Z. Li, P. Chang, S. Shan, X. Jiang, Z. Hu, B. Zhang, Y. Zhao and S. Ou, Enhanced intracellular calcium detection using dopamine-modified graphene quantum dots with dual emission mechanisms, Spectrochim. Acta, Part A, 2025, 328, 125475 CrossRef CAS PubMed.
  132. A. S. Shilpa, T. D. Thangadurai, G. M. Bhalerao and S. Maji, Tailor-designed carbon-based novel fluorescent architecture for nanomolar detection of radioactive elements U(VI) and Th(IV) in pH ± 5.0, Talanta, 2024, 272, 125783 CrossRef CAS.
  133. R. Guo, S. Zhou, Y. Li, X. Li, L. Fan and N. H. Voelcker, Rhodamine-functionalized graphene quantum dots for detection of Fe3+ in cancer stem cells, ACS Appl. Mater. Interfaces, 2015, 7, 23958–23966 CrossRef CAS PubMed.
  134. M. B. Miltenburg, T. B. Schon, E. L. Kynaston, J. G. Manion and D. S. Seferos, Electrochemical polymerization of functionalized graphene quantum dots, Chem. Mater., 2017, 29, 6611–6615 CrossRef CAS.
  135. C. H. Park, H. Yang, J. Lee, H. H. Cho, D. Kim, D. C. Lee and B. J. Kim, Multicolor emitting block copolymer-integrated graphene quantum dots for colorimetric, simultaneous sensing of temperature, pH, and metal ions, Chem. Mater., 2015, 27, 5288–5294 CrossRef CAS.
  136. S. Lee, J. Lee and S. Jeon, Aggregation-induced emission of matrix-free graphene quantum dots via selective edge functionalization of rotor molecules, Sci. Adv., 2023, 9, eade2585 CrossRef CAS.
  137. M. Ou, Y. Zhu, J. Liu, Z. Liu, J. Wang, J. Sun, C. Qin and L. Dai, Polyvinyl alcohol fiber with enhanced strength and modulus and intense cyan fluorescence based on covalently functionalized graphene quantum dots, Chin. Chem. Lett., 2025, 36, 110510 CrossRef CAS.
  138. R. Wang, X. Du, Y. Wu, J. Zhai and X. Xie, Graphene quantum dots integrated in ionophore-based fluorescent nanosensors for Na+ and K+, ACS Sens., 2018, 3, 2408–2414 CrossRef CAS.
  139. P. Mohammadian, M. Masteri-Farahani and N. Mosleh, A fluorescent turn-on nanosensor for selective detection of L-morphine using D-cysteine-functionalized graphene quantum dots, J. Photochem. Photobiol., A, 2025, 458, 115970 CrossRef CAS.
  140. L. Wang, J. Zheng, S. Yang, C. Wu, C. Liu, Y. Xiao, Y. Li, Z. Qing and R. Yang, Two-photon sensing and imaging of endogenous biological cyanide in plant tissues using graphene quantum dot/gold nanoparticle conjugate, ACS Appl. Mater. Interfaces, 2015, 7, 19509–19515 CrossRef CAS.
  141. S. D. Hiremath, A. Thakuri, M. M. Joseph, A. A. Bhosle, K. K. Maiti, M. Banerjee and A. Chatterjee, Cationic donor-two-acceptor dye-graphene quantum dot nanoconjugate for the ratiometric detection of bisulfite ions and monitoring of SO2 levels in heat-stressed cells, ACS Appl. Nano Mater., 2023, 6, 9958–9967 CrossRef CAS.
  142. J. Cai, G. Ma, X. Li, S. On and X. Wang, Saccharide-passivated graphene quantum dots from graphite for iron(III) sensing and bioimaging, ACS Appl. Nano Mater., 2023, 6, 11001–11012 CrossRef CAS.
  143. A. Colburn, N. Wanninayake, D. Y. Kim and D. Bhattacharyya, Cellulose-graphene quantum dot composite membranes using ionic liquid, J. Membr. Sci., 2018, 556, 293–302 CrossRef CAS.
  144. Y. H. Chang, C. C. Chang, L. Y. Chang, P. C. Wang, P. Kanokpaka and M. H. Yeh, Self-powered triboelectric sensor with N-doped graphene quantum dots decorated polyaniline layer for non-invasive glucose monitoring in human sweat, Nano Energy, 2023, 112, 108505 CrossRef CAS.
  145. R. Roy, R. Thapa, S. Biswas, S. Saha, U. K. Ghorai, D. Sen, E. M. Kumar, G. S. Kumar, N. Mazumder, D. Roy and K. K. Chattopadhyay, Resonant energy transfer in van der Waal stacked MoS2-functionalized graphene quantum dots composite with ab initio validation, Nanoscale, 2018, 10, 16822–16829 RSC.
  146. S. Chung, R. A. Revia and M. Zhang, Graphene quantum dots and their applications in bioimaging, biosensing, and therapy, Adv. Mater., 2021, 33, 1904362 CrossRef CAS PubMed.
  147. H. Bian, Q. Wang, S. Yang, C. Yan, H. Wang, L. Liang, Z. Jin, G. Wang and S. (Frank) Liu, Nitrogen-doped graphene quantum dots for 80% photoluminescence quantum yield for inorganic γ-CsPbI3 perovskite solar cells with efficiency beyond 16%, J. Mater. Chem. A, 2019, 7, 5740–5747 RSC.
  148. R. S. Li, B. Yuan, J. H. Liu, M. L. Liu, P. F. Gao, Y. F. Li, M. Li and C. Z. Huang, Boron and nitrogen co-doped single-layered graphene quantum dots: a high-affinity platform for visualizing the dynamic invasion of HIV DNA into living cells through fluorescence resonance energy transfer, J. Mater. Chem. B, 2017, 5, 8719–8724 RSC.
  149. S. Abbas, A. Abbas, T. Zahra, J. Kazmi, W. Ahmad, N. Ahmed, T. M. Lim and H. Cong, Green and gram-scale synthesis of uniform graphene quantum dots from biomass waste: a highly selective probe for nanomolar Hg2+ sensing, Mater. Today Chem., 2025, 47, 102830 CrossRef CAS.
  150. L. Y. Chang, M. Rinawati, Y. T. Guo, Y. C. Lin, C. Y. Chang, W. N. Su, H. Mizuguchi, W. H. Huang, J. L. Chen and M. H. Yeh, Nitrogen-doped graphene quantum dots incorporated into MOF-derived NiCo layered double hydroxides for nonenzymatic lactate detection in noninvasive biosensors, ACS Appl. Nano Mater., 2024, 7, 14431–14442 CrossRef CAS.
  151. P. Kadyan, M. Kumar, A. Tufail, A. Ragusa, S. K. Kataria and A. Dubey, Microwave-assisted green synthesis of fluorescent graphene quantum dots (GQDs) using Azadirachta indica leaves: enhanced synergistic action of antioxidant and antimicrobial effects and unveiling computational insights, Mater. Adv., 2025, 6, 805–826 RSC.
  152. L. Zhu, D. Li, H. Lu, S. Zhang and H. Gao, Lignin-based fluorescence-switchable graphene quantum dots for Fe3+ and ascorbic acid detection, Int. J. Biol. Macromol., 2022, 194, 254–263 CrossRef CAS PubMed.
  153. V. Kansara and M. Patel, Modulating the properties of graphene quantum dots by heteroatom doping for biomedical applications, Colloids Surf., A, 2024, 691, 133906 CrossRef CAS.
  154. C. Zhu, S. Yang, J. Sun, P. He, N. Yuan, J. Ding, R. Mo, G. Wang, G. Ding and X. Xie, Deep ultraviolet emission photoluminescence and high luminescece efficiency of ferric passivated graphene quantum dots: strong negative inductive effect of Fe, Synth. Met., 2015, 209, 468–472 CrossRef CAS.
  155. J. W. Lee, J. H. Kwak, J. Kim, Y. K. Jang, J. T. Han, T. J. Kim, K. S. Hong, H. J. Jeong and I. H. S. Yang, Highly emissive blue graphene quantum dots with excitation-independent emission via ultrafast liquid-phase photoreduction, RSC Adv., 2024, 14, 11524–11532 RSC.
  156. C. Carrera, A. Galan-Gonzalez, W. K. Maser and A. M. Benito, Multifaceted role of H2O2 in the solvothermal synthesis of green-emitting nitrogen-doped graphene quantum dots, Chem. Sci., 2025, 16, 3662–3670 RSC.
  157. Z. Wang, D. Chen, B. Gu, B. Gao, Z. Liu, Y. Yang, Q. Guo, X. Zheng and G. Wang, Yellow emissive nitrogen-doped graphene quantum dots as a label-free fluorescent probe for Fe3+ sensing and bioimaging, Diamond Relat. Mater., 2020, 104, 107749 CrossRef.
  158. J. Li, H. Zhao, X. Zhao and X. Gong, Aggregation-induced enhanced red emission graphene quantum dots for integrated fabrication of luminescent solar concentrators, Nano Lett., 2024, 24, 11722–11729 CrossRef.
  159. A. Valimukhametova, O. Zub, N. Castro-Lopez, D. Vashani, H. Paul, U. C. Topkiran, A. Gasimli, K. Malkova, F. L. Wormley Jr and A. V. Naumov, Combination diagnostics in vivo: dual-mode ultrasound/NIR fluorescence imaging with neodymium- and thulium-doped graphene quantum dots, ACS Appl. Bio Mater., 2025, 8, 4303–4314 CrossRef.
  160. Y. Yan, J. Chen, N. Li, J. Tian, K. Li, J. Jiang, J. Liu, Q. Tian and P. Chen, Systematic bandgap engineering of graphene quantum dots and applications for photocatalytic water splitting and CO2 reduction, ACS Nano, 2018, 12, 3523–3532 CrossRef PubMed.
  161. S. Zhu, J. Zhang, X. Liu, B. Li, X. Wang, S. Tang, Q. Meng, Y. Li, C. Shi, R. Hu and B. Yang, Graphene quantum dots with controllable surface oxidation, tunable fluorescence and up-conversion emission, RSC Adv., 2012, 2, 2717–2720 RSC.
  162. S. Jeong, R. L. Pinals, B. Dharmadhikari, H. Song, A. Kalluri, D. Debnath, Q. Wu, M. H. Ham, P. Patra and M. P. Landry, Graphene quantum dot oxidation governs noncovalent biopolymer adsorption, Sci. Rep., 2020, 10, 7074 CrossRef.
  163. D. Kurniawan, R. J. Weng, O. Setiawan, K. (Ken) Ostrikov and W. H. Chiang, Microplasma nanoengineering of emission-tuneable colloidal nitrogen-doped graphene quantum dots as smart environmental-responsive nanosensors and nanothermometers, Carbon, 2021, 185, 501–513 CrossRef.
  164. R. Wang, H. Fan, W. Jiang, G. Ni and S. Qu, Amino-functionalized graphene quantum dots prepared using high-softening point asphalt and their application in Fe3+ detection, Appl. Surf. Sci., 2019, 467–468, 446–455 CrossRef.
  165. H. Zhang, J. Wang, S. Wei, C. Wang, X. Yin, X. Song, C. Jiang and G. Sun, Nitrogen-doped graphene quantum dot-based portable fluorescent sensors for the sensitive detection of Fe3+ and ATP with logic gate operation, J. Mater. Chem. B, 2023, 11, 6082–6094 RSC.
  166. L. Ruiyi, Q. Xie, Y. Wang, Z. Li and X. Liu, Synthesis of folic acid, histidine, and serine-functionalized and boron and phosphorus-doped graphene quantum dots with excellent yellow luminescence behavior in aqueous/solid states and their use for the fluorescence detection of Fe2+ in urine, New J. Chem., 2024, 48, 14984–14994 RSC.
  167. A. Abbas, S. Rubab, A. Rehman, S. Irfan, H. M. A. Sharif, Q. Liang and T. A. Tabish, One-step green synthesis of biomass-derived graphene quantum dots as a highly selective optical sensing probe, Mater. Today Chem., 2023, 30, 101555 CrossRef.
  168. S. Tang, D. Chen, Y. Yang, C. Wang, X. Li, Y. Wang, C. Gu and Z. Cao, Mechanisms behind multicolor tunable near-infrared triple emission in graphene quantum dots and ratio fluorescent probe for water detection, J. Colloid Interface Sci., 2022, 617, 182–192 CrossRef.
  169. S. Zhuo, M. Shao and S. T. Lee, Upconversion and downconversion fluorescent graphene quantum dots: ultrasonic preparation and photocatalysis, ACS Nano, 2012, 6, 1059–1064 CrossRef PubMed.
  170. X. Wen, P. Yu, Y. R. Toh, X. Ma and J. Tang, On the upconversion fluorescence in carbon nanodots and graphene quantum dots, Chem. Commun., 2014, 50, 4703–4706 RSC.
  171. T. V. Huynh, N. T. N. Anh, W. Darmanto and R. A. Doong, Erbium-doped graphene quantum dots with up- and down-conversion luminescence for effective detection of ferric ions in water and human serum, Sens. Actuators, B, 2021, 328, 129056 CrossRef.
  172. F. Khan and J. H. Kim, N-functionalized graphene quantum dots with ultrahigh quantum yield and large stokes shift: efficient downconverters for CIGS solar cells, ACS Photonics, 2018, 5, 4637–4643 CrossRef.
  173. H. Zhou, M. Ou, D. Sun and C. Yang, Facile preparation of highly fluorescent nitrogen-doped graphene quantum dots for sensitive Fe3+ detection, Opt. Laser Technol., 2022, 156, 108542 CrossRef.
  174. Y. X. Wang, M. Rinawati, J. D. Zhan, K. Y. Lin, C. J. Huang, K. J. Chen, H. Mizuguchi, J. C. Jiang, B. J. Hwang and M. H. Yeh, Boron-doped graphene quantum dots anchored to carbon nanotubes as noble metal-free electrocatalysts of uric acid for a wearable sweat sensor, ACS Appl. Nano Mater., 2022, 5, 11100–11110 CrossRef.
  175. S. Li, Y. Li, J. Cao, J. Zhu, L. Fan and X. Li, Sulfur-doped graphene quantum dots as a novel fluorescent probe for highly selective and sensitive setection of Fe3+, Anal. Chem., 2014, 86, 10201–10207 CrossRef PubMed.
  176. N. T. N. Anh, P. Y. Chang and R. A. Doong, Sulfur-doped graphene quantum dot-based paper sensor for highly sensitive and selective detection of 4-nitrophenol in contaminated water and wastewater, RSC Adv., 2019, 9, 26588–26597 RSC.
  177. T. V. Tam, K. C. Bhamu, M. J. Kim, S. G. Kang, J. S. Chung, S. H. Hur and W. M. Choi, Engineering phosphorous doped graphene quantum dots decorated on graphene hydrogel as effective photocatalyst and high-current density electrocatalyst for seawater splitting, Chem. Eng. J., 2024, 480, 148190 CrossRef.
  178. P. Yang, J. Su, R. Guo, F. Yao and C. Yuan, B,N-co-doped graphene quantum dots as fluorescence sensor for detection of Hg2+ and F ions, Anal. Methods, 2019, 11, 1879–1883 RSC.
  179. Z. Geng, A highly sensitive fluorescent detection method utilizing B and S co-doped graphene quantum dots for ibuprofen analysis, Alexandria Eng. J., 2025, 112, 17–25 CrossRef.
  180. G. Guo, X. Qian, T. Li, S. Gao, B. Zhang, L. Wang, K. Liu, C. Gu and D. Chen, A smartphone-integrated nanosensor based on N, P co-doped graphene quantum dots for fluorescence detection of acid red 18 in food, Curr. Appl. Phys., 2023, 56, 92–99 CrossRef.
  181. K. S. Tiras, E. Soheyli, Z. Sharifirad and E. Mutlugun, Optimization of high efficiency blue emissive N-, S-doped graphene quantum dots, Opt. Mater., 2025, 159, 116544 CrossRef.
  182. B. Li, X. Xiao, M. Hu, Y. Wang, Y. Wang, X. Yan, Z. Huang, P. Servati, L. Huang, J. Tang and B. Mn, N co-doped graphene quantum dots for fluorescence sensing and biological imaging, Arabian J. Chem., 2022, 15, 103856 CrossRef.
  183. A. R. Valimukhametova, B. H. Lee, U. C. Topkiran, K. Gries, R. Gonzalez-Rodriguez, J. L. Coffer, G. Akkaraju and A. Naumov, Cancer therapeutic siRNA delivery and imaging by nitrogen- and neodymium-doped graphene quantum dots, ACS Biomater. Sci. Eng., 2023, 9, 3425–3434 CrossRef PubMed.
  184. F. A. P. K, G. R. Tharani, A. Sundaramoorthy, V. K. Shanmugam, K. Subramani, S. Chinnathambi, G. N. Pandian, V. Raghavan, A. N. Grace, S. Ganesan and M. Rajendiran, An ultra-sensitive detection of melamine in milk using rare-earth doped graphene quantum dots- synthesis and optical spectroscopic approach, Microchem. J., 2024, 196, 109670 CrossRef.
  185. Y. Wu, C. Combs, B. O. Okosun, K. Tayutivutikul, D. C. Darland and J. X. Zhao, Fe3+-doped graphene quantum dots-based nanozyme for H2O2 detection in cellular metabolic distress, ACS Appl. Nano Mater., 2025, 8, 2774–2784 CrossRef.
  186. S. Nangare, K. Chaudhari and P. Patil, Poly-L-lysine functionalized graphene quantum dots embedded zirconium metal–organic framework-based fluorescence switch on-off-on nanoprobe for highly sensitive and selective detection of taurine, J. Photochem. Photobiol., A, 2024, 446, 115158 CrossRef.
  187. W. Xuan, L. Ruiyi, F. Saiying, L. Zaijun, W. Guangli, G. Zhiguo and L. Junkang, D-penicillamine-functionalized graphene quantum dots for fluorescent detection of Fe3+ in iron supplement oral liquids, Sens. Actuators, B, 2017, 243, 211–220 CrossRef.
  188. Y. Ping, L. Ruiyi, Y. Yongqiang, L. Zaijun, G. Zhiguo, W. Guangli and L. Junkang, Pentaethylenehexamine and D-penicillamine co-functionalized graphene quantum dots for fluorescent detection of mercury(II) and glutathione and bioimaging, Spectrochim. Acta, Part A, 2018, 203, 139–146 CrossRef PubMed.
  189. T. Du, J. She, X. Yang, Y. Zhao, S. Zhou and J. Zhao, Role of functionalization in the fluorescence quantum yield of graphene quantum dots, Appl. Phys. Lett., 2023, 122, 142107 CrossRef.
  190. L. Gao, L. Ju and H. Cui, Chemiluminescent and fluorescent dual-signal graphene quantum dots and their application in pesticide sensing arrays, J. Mater. Chem. C, 2017, 5, 7753–7758 RSC.
  191. D. Li, F. Nie, T. Tang and K. Tian, Determination of ferric ion via its effect on the enhancement of the chemiluminescece of the permanganate-sulfite system by nitrogen-doped graphene quantum dots, Microchim. Acta, 2018, 185, 431 CrossRef PubMed.
  192. J. Zhang, Y. Li and S. Han, Simultaneous detection of iodide and mercuric ions by nitrogen-sulfur co-doped graphene quantum dots based on flow injection “turn off-on” chemiluminescence analysis system, Microchem. J., 2019, 147, 1141–1146 CrossRef CAS.
  193. X. Qin, Z. Zhan, R. Zhang, K. Chu, Z. Whitworth and Z. Ding, Nitrogen- and sulfur-doped graphene quantum dots for chemiluminescence, Nanoscale, 2023, 15, 3864–3871 RSC.
  194. X. L. Cai, B. Zheng, Y. Zhou, M. R. Younis, F. B. Wang, W. M. Zhang, Y. G. Zhou and X. H. Xia, Synergistically mediated enhancement of cathodic and anodic electrochemiluminescence of graphene quantum dots through chemical and electrochemical reactions of coreactants, Chem. Sci., 2018, 9, 6080–6084 RSC.
  195. L. Li, J. Li, R. Fei, C. Wang, Q. Lu, J. Zhang, L. Jiang and J. Zhu, A facile microwave avenue to electrochemiluminescent two-color graphene quantum dots, Adv. Funct. Mater., 2012, 22, 2971–2979 CrossRef.
  196. J. Pizarro, R. Segura, D. Tapia, F. Navarro, F. Fuenzalida and M. J. Aguirre, Inexpensive and green electrochemical sensor for the determination of Cd (II) and Pb(II) by square wave anodic stripping voltammetry in bivalve molluscs, Food Chem., 2020, 321, 126682 CrossRef PubMed.
  197. S. Saisree, R. Aswathi, J. S. Arya Nair and K. Y. Sandhya, Radical sensitivity and selectivity in the electrochemical sensing of cadmium ions in water by polyaniline-derived nitrogen-doped graphene quantum dots, New J. Chem., 2021, 45, 110–122 RSC.
  198. S. Saisree, A. Nair J. S and S. K. Yesodha, Graphene quantum dots doped with sulfur and nitrogen as versatile electrochemical sensors for heavy metal ions Cd(II), Pb(II), and Hg(II), ACS Appl. Nano Mater., 2023, 6, 1224–1234 CrossRef.
  199. A. Rakovich and T. Rakovich, Semiconductor versus graphene quantum dots as fluorescent probes for cancer diagnosis and therapy applications, J. Mater. Chem. B, 2018, 6, 2690–2712 RSC.
  200. J. Chen, S. Yin, F. Yang, S. Guo, J. Zhang, Z. Lu and T. Gao, A single-molecule graphene quantum dot: a novel efficient photosensitizer for photodynamic cancer therapy, Chem. Sci., 2025, 16, 13923–13934 RSC.
  201. G. Papaparaskeva, P. Papagiorgis, G. Itskos, C. Michael, E. Tanasa and T. Krasia-Christoforou, Color-tunable, white light emission from electrospun nanocomposite fibers sensitized by graphene quantum dots, ACS Appl. Opt. Mater., 2025, 3, 839–851 CrossRef.
  202. B. Ouyang, Q. Zhong, P. Ouyang, Y. Yuan, X. Wu and S. T. Yang, Graphene quantum dots enhance the biological nitrogen fixation by up-regulation of cellular metabolism and electron transport, Chem. Eng. J., 2024, 487, 150694 CrossRef.
  203. H. Mirzaei, M. H. Ehsani and A. Shakeri, Development and characterization of a graphene quantum dot/g-C3N4 photocatalyst for efficient degradation of Rhodamine B, Sci. Rep., 2025, 15, 27276 CrossRef PubMed.
  204. A. Raheem, R. Kumar and S. Dutta, Designing of N-doped graphene quantum dot/NiAl layered double hydroxide/TiO2 heterostructure for enhanced photocatalytic hydrogen production, ACS Appl. Energy Mater., 2025, 8, 10433–10444 CrossRef.
  205. S. Fatima, M. W. Hakim, X. Zheng, Y. Sun, Z. Li, N. Han, M. Li, Z. Wang, L. Han, L. Wang, S. Khan, J. Liu and H. Li, Nitrogen doped graphene quantum dots (NGQD) pillared Ta4C3Tx MXene as high-performance electrochemical supercapacitors, J. Power Sources, 2025, 645, 237190 CrossRef.
  206. Q. Ding, C. Li, H. Wang, C. Xu and H. Kuang, Electrochemical detection of heavy metal ions in water, Chem. Commun., 2021, 57, 7215–7231 RSC.
  207. J. Lee, M. Kim, S. Park, J. Lee, Q. Chen, J. Kim, T. Defferriere, H. Park, S. Jeon and I. D. Kim, Bandgap-engineered graphene quantum dot photosensitizers for tunable light spectrum-activated NO2 sensors, ACS Nano, 2025, 19, 32732–32743 CrossRef PubMed.
  208. Y. H. Chiu, M. Rinawati, L. Y. Chang, S. Aulia, C. Li, P. C. Shi, K. J. Chen, W. H. Huang, H. Mizuguchi and M. H. Yeh, Promoting the signal reliability of non-invasive biosensors via a N-doped graphene quantum dot modified Prussian blue analogue protective layer for glucose monitoring, J. Mater. Chem. B, 2025, 13, 7381–7392 RSC.
  209. A. Y. Tong, N. Xu, T. Yan, F. B. Chen and C. Y. Yang, Graphene quantum dot-enhanced chemiluminescence sensor based on surface molecular imprinting recognition for the detection of tetrabromobisphenol A, Spectrochim. Acta, Part A, 2026, 347, 126981 CrossRef PubMed.
  210. W. Pengsook, C. Thanachayanont, W. T. Wahyuni and P. Hasin, Cobalt-modified exfoliated zirconium phosphate/histidine-functionalized graphene quantum dots-based electrochemical biosensor for promoting sensitive detection of methyl parathion in agricultural foods and water bodies samples, Talanta, 2026, 296, 128403 CrossRef.
  211. L. Zhou, J. Geng and B. Liu, Graphene quantum dots from polycyclic aromatic hydrocarbon for bioimaging and sensing of Fe3+ and hydrogen heroxide, Part. Part. Syst. Charact., 2013, 30, 1086–1092 CrossRef.
  212. J. Ju and W. Chen, Synthesis of highly fluorescent nitrogen-doped graphene quantum dots for sensitive, label-free detection of Fe (III) in aqueous media, Biosens. Bioelectron., 2014, 58, 219–225 CrossRef.
  213. T. T. Xu, J. X. Yang, J. M. Song, J. S. Chen, H. L. Niu, C. J. Mao, S. Y. Zhang and Y. H. Shen, Synthesis of high fluorescence graphene quantum dots and their selective detection for Fe3+ in aqueous solution, Sens. Actuators, B, 2017, 243, 863–872 CrossRef.
  214. Y. P. Zhang, J. M. Ma, Y. S. Yang, J. X. Ru, X. Y. Liu, Y. Ma and H. C. Guo, Synthesis of nitrogen-doped graphene quantum dots (N-GQDs) from marigold for detection of Fe3+ ion and bioimaging, Spectrochim. Acta, Part A, 2019, 217, 60–67 CrossRef PubMed.
  215. Z. G. Khan, T. N. Agrawal, S. B. Bari, S. N. Nangare and P. O. Patil, Application of surface nitrogen-doped graphene quantum dots in the sensing of ferric ions and glutathione: spectroscopic investigations and DFT calculations, Spectrochim. Acta, Part A, 2024, 306, 123608 CrossRef.
  216. C. Ren, M. Zhang, N. Zheng, Z. Nie, F. Zhang, J. Tang and G. Chen, Metal/non-metal doped graphene quantum dots as dual fluorescent probes for detection of pesticide residues and heavy metal ions, Spectrochim. Acta, Part A, 2026, 346, 126886 CrossRef.
  217. L. Chen, C. Wu, P. Du, X. Feng, P. Wu and C. Cai, Electrolyzing synthesis of boron-doped graphene quantum dots for fluorescence determination of Fe3+ ions in water samples, Talanta, 2017, 164, 100–109 CrossRef PubMed.
  218. R. Wang, L. Jiao, X. Zhou, Z. Guo, H. Bian and H. Dai, Highly fluorescent graphene quantum dots from biorefinery waste for tri-channel sensitive detection of Fe3+ ions, J. Hazard. Mater., 2021, 412, 125096 CrossRef.
  219. X. Zhu, Z. Zhang, Z. Xue, C. Huang, Y. Shan, C. Liu, X. Qin, W. Yang, X. Chen and T. Wang, Understanding the selective detection of Fe3+ based on graphene quantum dots as fluorescent probes: the Ksp of a metal hydroxide-assisted mechanism, Anal. Chem., 2017, 89, 12054–12058 CrossRef.
  220. W. Wang, Z. Wang, J. Liu, Y. Peng, X. Yu, W. Wang, Z. Zhang and L. Sun, One-pot facile synthesis of graphene quantum dots from rice husks for Fe3+ sensing, Ind. Eng. Chem. Res., 2018, 57, 9144–9150 CrossRef.
  221. M. H. M. Facure, R. Schneider, L. A. Mercante and D. S. Correa, Rational hydrothermal synthesis of graphene quantum dots with optimized luminescent properties for sensing applications, Mater. Today Chem., 2022, 23, 100755 CrossRef.
  222. A. D. Chowdhury and R. A. Doong, Highly sensitive and selective detection of nanomolar ferric ions using dopamine functionalized graphene quantum dots, ACS Appl. Mater. Interfaces, 2016, 8, 21002–21010 CrossRef.
  223. W. Ye, L. Ruiyi and L. Zaijun, Arginine- and serine-functionalized boron-doped graphene quantum dots with dual-emission fluorescence for ultrasensitive detection of Fe3+ in iron supplementation beverages, New J. Chem., 2025, 49, 8743–8751 RSC.
  224. L. Xu, W. Mao, J. Huang, S. Li, K. Huang, M. Li, J. Xia and Q. Chen, Economical, green route to highly fluorescence intensity carbon materials based on ligninsulfonate/graphene quantum dots composites: application as excellent fluorescent sensing platform for detection of Fe3+ ions, Sens. Actuators, B, 2016, 230, 54–60 CrossRef.
  225. C. Wang, Y. Sun, J. Jin, Z. Xiong, D. Li, J. Yao and Y. Liu, Highly selective, rapid-functioning and sensitive fluorescent test paper based on graphene quantum dots for on-line detection of metal ions, Anal. Methods, 2018, 10, 1163–1171 RSC.
  226. R. Das, H. Sugimoto, M. Fujii and P. K. Giri, Quantitative Understanding of charge-transfer-mediated Fe3+ sensing and fast photoresponse by N-doped graphene quantum dots decorated on plasmonic Au nanoparticles, ACS Appl. Mater. Interfaces, 2020, 12, 4755–4768 CrossRef.
  227. K. B. R. Teodoro, M. H. M. Facure, R. Schneider, A. D. Alvarenga, R. S. Andre and D. S. Correa, Self-standing thin films of cellulose nanocrystals and graphene quantum dots for detection of trace iron(III), ACS Appl. Nano Mater., 2023, 6, 11561–11571 CrossRef.
  228. R. Das, A. Paria and P. K. Giri, Machine learning framework for selective and sensitive metal ion sensing with nitrogen-doped graphene quantum dots heterostructure, Carbon, 2025, 232, 119800 CrossRef.
  229. G. Y. Karaca, Graphene quantum dot-based gold-nickel micromotors for sensitive detection of ferric ions, J. Fluoresc., 2025 DOI:10.1007/s10895-025-04238-6.
  230. K. Saenwong, P. Nuengmatcha, P. Sricharoen, N. Limchoowong and S. Chanthai, GSH-doped GQDs using citric acid rich-lime oil extract for highly selective and sensitive determination and discrimination of Fe3+ and Fe2+ in the presence of H2O2 by a fluorescence “turn-off” sensor, RSC Adv., 2018, 8, 10148–10157 RSC.
  231. N. Pimsin, C. Keawprom, Y. Areerob, N. Limchoowong, P. Sricharoen, P. Nuengmatcha, W. C. Oh and S. Chanthai, Selective Fe(II)-fluorescence sensor with validated two-consecutive working range using N,S,I-GQDs associated with garlic extract as an auxiliary green chelating agent, RSC Adv., 2022, 12, 14356–14367 RSC.
  232. Z. Li, Y. Wang, Y. Ni and S. Kokot, A rapid and label-free dual detection of Hg (II) and cysteine with the use of fluorescence switching of graphene quantum dots, Sens. Actuators, B, 2015, 207, 490–497 CrossRef CAS.
  233. J. Kappen, M. K. Aravind, P. Varalakshmi, B. Ashokkumar and S. A. John, Hydroxyl rich graphene quantum dots for the determination of Hg(II) in the presence of large concentration of major interferents and in living cells, Microchem. J., 2020, 157, 104915 CrossRef CAS.
  234. J. Kappen, S. Ponkarpagam and S. A. John, Study on the interactions between graphene quantum dots and Hg(II): unraveling the origin of photoluminescence quenching of graphene quantum dots by Hg(II), Colloids Surf., A, 2020, 591, 124551 CrossRef CAS.
  235. S. Reagen, Y. Wu, X. Liu, R. Shahni, J. Bogenschuetz, X. Wu, Q. R. Chu, N. Oncel, J. Zhang, X. Hou, C. Combs, A. Vasquez and J. X. Zhao, Synthesis of highly near-infrared fluorescent graphene quantum dots using biomass-derived materials for in vitro cell imaging and metal ion detection, ACS Appl. Mater. Interfaces, 2021, 13, 43952–43962 CrossRef CAS.
  236. B. Shi, L. Zhang, C. Lan, J. Zhao, Y. Su and S. Zhao, One-pot green synthesis of oxygen-rich nitrogen-doped graphene quantum dots and their potential application in pH-sensitive photoluminescence and detection of mercury (II) ions, Talanta, 2015, 142, 131–139 CrossRef CAS PubMed.
  237. D. Peng, L. Zhang, R. P. Liang and J. D. Qiu, Rapid detection of mercury ions based on nitrogen-doped graphene quantum dots accelerating formation of manganese porphyrin, ACS Sens., 2018, 3, 1040–1047 CrossRef CAS.
  238. P. Naksen, P. Khamlam, P. Khemthong, N. Yodsin, J. Phanthasri, S. Youngjan, A. Prakobkij, N. Kitchawengkul, S. Jungsuttiwong, A. Samphao and P. Jarujamrus, Nitrogen and sulfur doped graphene quantum dots as a fluorometric paper-based sensor for highly selective and sensitive detection of mercury ions in aqueous samples, Microchem. J., 2025, 216, 114623 CrossRef CAS.
  239. Z. Liu, Z. Mo, N. Liu, R. Guo, X. Niu, P. Zhao and X. Yang, One-pot synthesis of highly fluorescent boron and nitrogen co-doped graphene quantum dots for the highly sensitive and selective detection of mercury ions in aqueous media, J. Photochem. Photobiol., A, 2020, 389, 112255 CrossRef CAS.
  240. A. N. Nair, V. S. N. Chava, S. Bose, T. Zheng, S. Pilla and S. T. Sreenivasan, In situ doping-enabled metal and nonmetal codoping in graphene quantum dots: synthesis and application for contaminant sensing, ACS Sustainable Chem. Eng., 2020, 8, 16565–16576 CrossRef CAS.
  241. Z. Xiaoyan, L. Zhangyi and L. Zaijun, Fabrication of valine-functionalized graphene quantum dots and its use as a novel optical probe for sensitive and selective detection of Hg2+, Spectrochim. Acta, Part A, 2017, 171, 415–424 CrossRef CAS.
  242. R. Li, X. Wang, Z. Li, H. Zhu and J. Liu, Folic acid-functionalized graphene quantum dots with tunable fluorescence emission for cancer cell imaging and optical detection of Hg2+, New J. Chem., 2018, 42, 4352–4360 RSC.
  243. O. J. Achadu and T. Nyokong, Application of graphene quantum dots functionalized with thymine and thymine-appended zinc phthalocyanine as novel photoluminescent nanoprobes, New J. Chem., 2017, 41, 1447–1458 RSC.
  244. Z. Wu, C. Dai, Y. Wang, L. Ma, G. Zang, Q. Liu and S. Zhu, A novel sensor for visual and selective detection of Hg2+ based on functionalized doped quantum dots, Anal. Methods, 2022, 14, 2368–2375 RSC.
  245. B. Tian, Y. Kou, X. Jiang, J. Lu, Y. Xue, M. Wang and L. Tan, Ultrasensitive determination of mercury ions using a glassy carbon electrode modified with nanocomposites consisting of conductive polymer and amino-functionalized graphene quantum dots, Microchim. Acta, 2020, 187, 210 CrossRef CAS PubMed.
  246. X. Qi and Z. Wang, Graphene quantum dots functionalized Ce-ZnO nanofibers with enriched oxygen vacancy sites morphology to improve the efficiency of selective electrochemical detection of Hg (II), Diamond Relat. Mater., 2023, 139, 110241 CrossRef CAS.
  247. A. Dhanagar and A. Shaheen, Self-assembled luminescent droplets from graphene quantum dots induced by a Gemini surfactant for selective detection of mercury(II), Langmuir, 2025, 41, 4136–4145 CrossRef CAS.
  248. T. V. Tam and W. M. Choi, One-pot synthesis of highly fluorescent amino-functionalized graphene quantum dots for effective detection of copper ions, Curr. Appl. Phys., 2018, 18, 1255–1260 CrossRef.
  249. C. Ren, C. Tian, W. Cao, M. Zhang, T. Zhang, J. Tang, F. Zhang, G. Chen and J. Tang, Controllable functionalization of amino-functionalized graphene quantum dots as fluorescent probe for detection of Cu(II) ions detection, Mater. Lett., 2024, 364, 136393 CrossRef CAS.
  250. Y. Liu and D. Y. Kim, Ultraviolet and blue emitting graphene quantum dots synthesized from carbon nano-onions and their comparison for metal ion sensing, Chem. Commun., 2015, 51, 4176–4179 RSC.
  251. J. Wen, M. Li, J. Xiao, C. Liu, Z. Li, Y. Xie, P. Ning, H. Cao and Y. Zhang, Novel oxidative cutting graphene oxide to graphene quantum dots for electrochemical sensing application, Mater. Today Commun., 2016, 8, 127–133 CrossRef CAS.
  252. L. Ding, Z. Zhao, D. Li, X. Wang and J. Chen, An “off-on” fluorescent sensor for copper ion using graphene quantum dots based on oxidation of L-cysteine, Spectrochim. Acta, Part A, 2019, 214, 320–325 CrossRef CAS.
  253. P. C. Yeh, S. Yoon, D. Kurniawan, Y. G. Chung and W. H. Chiang, Unraveling the fluorescence quenching of colloidal graphene quantum dots for selective metal ion detection, ACS Appl. Nano Mater., 2021, 4, 5636–5642 CrossRef CAS.
  254. Y. Guo, Q. Huang, F. Xu, Z. Luo, Y. Wei, Z. Chen, Z. Zeng, H. Zhang and H. Shi, A graphene quantum dots based dual-modal fluorometric and visualized detection of copper ions, Spectrochim. Acta, Part A, 2025, 328, 125442 CrossRef CAS.
  255. Z. Zhu, R. Li, Y. Li, P. Pan, J. Liu, Y. Qi, B. Zhou and Z. Yang, Paper-based electrodes with nitrogen-doped graphene quantum dots for detection of copper ions via electrochemiluminescence, Mater. Chem. Phys., 2023, 296, 127300 CrossRef CAS.
  256. D. Kumar, S. Rani, B. T. Mathew, R. K. Chikara, B. Nandan and R. K. Srivastava, Hydrothermal pyrolysis of styrofoam waste for efficient copper ion sensing using graphene quantum dots, ACS Sustainable Resour. Manage., 2024, 1, 1824–1833 CrossRef CAS.
  257. W. Li, Q. Niu, X. Pang, S. Li, Y. Liu, B. Li, S. Li, L. Wang, H. Guo and L. Wang, Optimized sensitivity in copper(II) ion detection: sustainable fabrication of fluorescence red-shifted graphene quantum dots via electron-withdrawing modulation, Molecules, 2025, 30, 1244 CrossRef CAS PubMed.
  258. N. Limchoowong, P. Sricharoen, Y. Areerob, P. Nuengmatcha, T. Sripakdee, S. Techawongstien and S. Chanthai, Preconcentration and trace determination of copper (II) in Thai food recipes using Fe3O4@Chi–GQDs nanocomposites as a new magnetic adsorbent, Food Chem., 2017, 230, 388–397 CrossRef CAS.
  259. Y. Li, W. Zhang, X. Jiang, Y. Kou, J. Lu and L. Tan, Investigation of photo-induced electron transfer between amino-functionalized graphene quantum dots and selenium nanoparticle and it's application for sensitive fluorescent detection of copper ions, Talanta, 2019, 197, 341–347 CrossRef CAS PubMed.
  260. I. Ibrahim, H. N. Lim, N. M. Huang, Z. T. Jiang and M. Altarawneh, Selective and sensitive visible-light-prompt photoelectrochemical sensor of Cu2+ based on CdS nanorods modified with Au and graphene quantum dots, J. Hazard. Mater., 2020, 391, 122248 CrossRef CAS PubMed.
  261. Y. Liu, Y. Sun and M. Yang, A double-potential ratiometric electrochemiluminescence platform based on g-C3N4 nanosheets (g-C3N4 NSs) and graphene quantum dots for Cu2+ detection, Anal. Methods, 2021, 13, 903–909 RSC.
  262. M. Sara, K. Abid, P. G. Gucciardi, L. M. Scolaro and G. Neri, Electrochemical determination of heavy metals by a graphene quantum dots/porphyrin-based supramolecular system, Synth. Met., 2025, 311, 117824 CrossRef.
  263. Y. X. Qi, M. Zhang, Q. Q. Fu, R. Liu and G. Y. Shi, Highly sensitive and selective fluorescent detection of cerebral lead(II) based on graphene quantum dot conjugates, Chem. Commun., 2013, 49, 10599–10601 RSC.
  264. X. Sun, Y. Peng, Y. Lin, L. Cai, F. Li and B. Liu, G-quadruplex formation enhancing energy transfer in self-assembled multilayers and fluorescence recognize for Pb2+ ions, Sens. Actuators, B, 2018, 255, 2121–2125 CrossRef.
  265. S. Saisree, V. S. Archana and S. K. Yesodha, Picomolar selective electrochemical sensing of lead ions by a gold-copper nanocluster-nitrogen-doped graphene quantum dot combination, ACS EST Water, 2024, 4, 3145–3152 CrossRef.
  266. V. Dhrishya, S. Saisree, V. S. Archana and S. K. Yesodha, Hybrid boron nitride/N-doped graphene quantum dots for specific and picomolar electrochemical detection of Pb(II) ions, ACS Appl. Nano Mater., 2025, 8, 1586–1595 CrossRef.
  267. S. Saisree, V. Dhrishya, G. Meena, A. S. Nair and K. Y. Sandhya, Unraveling the stability and specific electrochemical sensing of lead ions using copper nanoclusters on sulfur and nitrogen-doped graphene quantum dots, Analyst, 2025, 150, 4031–4040 RSC.
  268. F. Cai, X. Liu, S. Liu, H. Liu and Y. Huang, A simple one-pot synthesis of highly fluorescent nitrogen-doped graphene quantum dots for the detection of Cr(VI) in aqueous media, RSC Adv., 2014, 4, 52016–52022 RSC.
  269. S. Huang, H. Qiu, F. Zhu, S. Lu and Q. Xiao, Graphene quantum dots as on-off-on fluorescent probes for chromium(VI) and ascorbic acid, Microchim. Acta, 2015, 182, 1723–1731 CrossRef.
  270. E. Punrat, C. Maksuk, S. Chuanuwatanakul, W. Wonsawat and O. Chailapakul, Polyaniline/graphene quantum dot-modified screen-printed carbon electrode for the rapid determination of Cr(VI) using stopped-flow analysis coupled with voltammetric technique, Talanta, 2016, 150, 198–205 CrossRef PubMed.
  271. D. Nugroho, R. Benchawattananon, J. Janshongsawang, N. Pimsin, P. Porrawatkul, R. Pimsen, P. Nuengmatcha, P. Nueangmatcha and S. Chanthai, Ultra-trace analysis of chromium ions (Cr3+/Cr6+) in water sample using selective fluorescence turn-off sensor with natural carbon dots mixed graphene quantum dots nanohybrid composite synthesis by pyrolysis, Arabian J. Chem., 2024, 17, 105443 CrossRef.
  272. C. S. Ni, W. J. Zhang, W. Z. Bi, M. X. Wu, S. X. Feng, X. L. Chen and L. B. Qu, Facile synthesis of N-doped graphene quantum dots as a fluorescent sensor for Cr(VI) and folic acid detection, RSC Adv., 2024, 14, 26667–26673 RSC.
  273. L. Zhang, D. Peng, R. P. Liang and J. D. Qiu, Nitrogen-doped graphene quantum dots as a new catalyst accelerating the coordination reaction between cadmium(II) and 5,10,15,20- tetrakis(1-methyl-4-pyridinio)porphyrin for cadmium(II) sensing, Anal. Chem., 2015, 87, 10894–10901 CrossRef.
  274. P. Naksen, S. Boonruang, N. Yuenyong, H. L. Lee, P. Ramachandran, W. Anutrasakda, M. Amatatongchai, S. Pencharee and P. Jarujamrus, Sensitive detection of trace level Cd (II) triggered by chelation enhanced fluorescence (CHEF) “turn on”: nitrogen-doped graphene quantum dots (N-GQDs) as fluorometric paper-based sensor, Talanta, 2022, 242, 123305 CrossRef.
  275. S. Tang, H. J. Hu, X. Chang, H. Zhang, Y. Chen, Y. Li, W. X. Li, X. Hu, X. Liao and G. B. Jiang, Group competition mechanism enabled by nitrogen-doped graphene quantum dots for efficient Cd(II) detection and removal, Chem. Eng. J., 2025, 510, 161590 CrossRef.
  276. N. Wang, Z. X. Liu, R. S. Li, H. Z. Zhang, C. Z. Huang and J. Wang, The aggregation induced emission quenching of graphene quantum dots for visualizing the dynamic invasions of cobalt(II) into living cells, J. Mater. Chem. B, 2017, 5, 6394–6399 RSC.
  277. W. Boonta, C. Talodthaisong, S. Sattayaporn, C. Chaicham, A. Chaicham, S. Sahasithiwat, L. Kangkaewe and S. Kulchat, The synthesis of nitrogen and sulfur co-doped graphene quantum dots for fluorescence detection of cobalt(II) ions in water, Mater. Chem. Front., 2020, 4, 507–516 RSC.
  278. A. Xu, P. He, T. Huang, J. Li, X. Hu, P. Xiang, D. Chen, S. Yang, G. Wang and G. Ding, Selective supramolecular interaction of ethylenediamine functionalized graphene quantum dots: ultra-sensitive photoluminescence detection for nickel ion in vitro, Synth. Met., 2018, 244, 106–112 CrossRef CAS.
  279. M. Yao, J. Huang, Z. Deng, W. Jin, Y. Yuan, J. Nie, H. Wang, F. Du and Y. Zhang, Transforming glucose into fluorescent graphene quantum dots via microwave radiation for sensitive detection of Al3+ ions based on aggregation-induced enhanced emission, Analyst, 2020, 145, 6981–6986 RSC.
  280. S. Gogoi, R. Devi, H. S. Dutta, M. Bordoloi and R. Khan, Ratiometric fluorescence response of a dual light emitting reduced carbon dot/graphene quantum dot nanohybrid towards As(III), J. Mater. Chem. C, 2019, 7, 10309–10317 RSC.
  281. X. Ran, H. Sun, F. Pu, J. Ren and X. Qu, Ag nanoparticle-decorated graphene quantum dots for label-free, rapid and sensitive detection of Ag+ and biothiols, Chem. Commun., 2013, 49, 1079–1081 RSC.
  282. Z. Wang, D. Chen, B. Gu, B. Gao, T. Wang, Q. Guo and G. Wang, Biomass-derived nitrogen doped graphene quantum dots with color-tunable emission for sensing, fluorescence ink and multicolor cell imaging, Spectrochim. Acta, Part A, 2020, 227, 117671 CrossRef CAS PubMed.
  283. Z. Xue, D. Bai, Y. Li, Q. Zhang, T. Shao, P. Guo, D. Zhang and X. Zhou, Pt nanoclusters based on N, S co-doped graphene quantum dots as a smart probe for ultrasensitive Ag+ sensing, Microchem. J., 2022, 182, 107827 CrossRef CAS.
  284. T. Yang, F. Cai, X. Zhang and Y. Huang, Nitrogen and sulfur codoped graphene quantum dots as a new fluorescent probe for Au3+ ions in aqueous media, RSC Adv., 2015, 5, 107340–107347 RSC.
  285. Y. Huang, W. Zhou, Y. Wang and Y. Zhang, Crown ether-like structure in graphene quantum dots: ultra-sensitive photoluminescence sensor for Ca2+ in vitro, Synth. Met., 2020, 270, 116581 CrossRef CAS.
  286. D. Iannazzo, C. Espro, A. Ferlazzo, C. Celesti, C. Branca and G. Neri, Electrochemical and fluorescent properties of crown ether functionalized graphene quantum dots for potassium and sodium ions detection, Nanomaterials, 2021, 11, 2897 CrossRef CAS.
  287. X. Chen, Q. Tao, S. Yang, H. Wang, X. Wang, G. Ding, X. Gao, H. Dong and L. Rong, K+ concentration-based NMR-fluorescence dual-functional senescence sensing using graphene quantum dots with crown ether structure, Sens. Actuators, B, 2025, 431, 137442 CrossRef CAS.
  288. L. Lin, X. Song, Y. Chen, M. Rong, T. Zhao, Y. Jiang, Y. Wang and X. Chen, One-pot synthesis of highly greenish-yellow fluorescent nitrogen-doped graphene quantum dots for pyrophosphate sensing via competitive coordination with Eu3+ ions, Nanoscale, 2015, 7, 15427–15433 RSC.
  289. F. Salehnia, F. Faridbod, A. S. Dezfuli, M. R. Ganjali and P. Norouzi, Cerium(III) ion sensing based on graphene quantum dots fluorescent turn-off, J. Fluoresc., 2017, 27, 331–338 CrossRef CAS PubMed.
  290. M. Xia, X. E. Zhao, J. Sun, Z. Zheng and S. Zhu, Graphene quantum dots combined with the oxidase-mimicking activity of Ce4+ for ratiometric fluorescent detection of Ce4+ and alendronate sodium, Sens. Actuators, B, 2020, 319, 128321 CrossRef CAS.
  291. S. Wang, X. Chu, X. Xiang and Y. Cao, Highly selective antenna effect of graphene quantum dots (GQDs): a new fluorescent sensitizer for rare earth element terbium in aqueous media, Talanta, 2020, 209, 120504 CrossRef CAS.
  292. S. K. Guin, A. S. Ambolikar, J. P. Guin and S. Neogy, Exploring the excellent photophysical and electrochemical properties of graphene quantum dots for complementary sensing of uranium, Sens. Actuators, B, 2018, 272, 559–573 CrossRef CAS.
  293. L. Lu, L. Zhou, J. Chen, F. Yan, J. Liu, X. Dong, F. Xi and P. Chen, Nanochannel-confined graphene quantum dots for ultrasensitive electrochemical analysis of complex samples, ACS Nano, 2018, 12, 12673–12681 CrossRef CAS PubMed.
  294. J. Li, Z. Wang, J. Yang, X. Xia, R. Yi, J. Jiang, W. Liu, J. Chen, L. Chen and J. Xu, “On-off-on” fluorescence switch of graphene quantum dots: a cationic control strategy, Appl. Surf. Sci., 2021, 546, 149110 CrossRef CAS.
  295. C. Shen, S. Ge, Y. Pang, F. Xi, J. Liu, X. Dong and P. Chen, Facile and scalable preparation of highly luminescent N,S co-doped graphene quantum dots and their application for parallel detection of multiple metal ions, J. Mater. Chem. B, 2017, 5, 6593–6600 RSC.
  296. V. P. Thai, D. N. Tran, K. Kosugi, K. Takahashi, T. Sasaki and T. Kikuchi, One-step synthesis of N-doped graphene quantum dots via plasma contacting liquid for multiple heavy metal ion detection, ACS Appl. Nano Mater., 2024, 7, 12664–12672 CrossRef CAS.
  297. Y. Zhao, M. Luo, R. Yang, S. Yao, Q. Zhang, W. Yu, Z. Feng, Y. Yan and Y. Xu, Hydrothermal synthesis of biomass waste derived graphene quantum dots with high ion detection ability, Diamond Relat. Mater., 2025, 155, 112363 CrossRef CAS.
  298. A. Liu, Z. Dai, D. Ouyang, B. Mahara, L. Yang and X. Chen, In situ conversion of graphite into graphene quantum dots (GQDs) towards upcycling of spent lithium-ion batteries, Green Chem., 2025, 27, 12460–12471 RSC.
  299. M. Tan, G. Yang, M. Wang, Y. Wang, S. Gao, Y. Wang, W. Zhu, Y. Peng, Y. Lu and C. Song, A fluorescence sensor array based on triple-color emission S,N-co-doped graphene quantum dots for metal ions discrimination, Spectrochim. Acta, Part A, 2026, 347, 126986 CrossRef CAS PubMed.
  300. S. Dorontic, A. Bonasera, M. Scopelliti, M. Mojsin, M. Stevanovic, O. Markovic and S. Jovanovic, Blue luminescent amino-functionalized graphene quantum dots as a responsive material for potential detection of metal ions and malathion, J. Lumin., 2022, 252, 119311 CrossRef CAS.
  301. Y. Lou, W. Sun, L. Jiang, X. Fan, K. Zhang, S. Chen, Z. Li, J. Ji, J. Ou, L. Liao and A. Qin, Double-enhanced fluorescence of graphene quantum dots from cane molasses via metal and PEG modification for detecting metal ions and pigments, Opt. Mater., 2022, 133, 113037 CrossRef CAS.
  302. M. Llaver, S. D. Barrionuevo, J. M. Núñez, A. L. Chapana, R. G. Wuilloud, M. H. Aguirre and F. J. Ibañez, Fluorescent graphene quantum dots-enhanced machine learning for the accurate detection and quantification of Hg2+ and Fe3+ in real water samples, Environ. Sci.: Nano, 2024, 11, 2703–2715 RSC.
  303. P. Abdollahiyan, M. Hasanzadeh, F. Seidi and P. Pashazadeh-Panahi, An innovative colorimetric platform for the low-cost and selective identification of Cu(II), Fe(III), and Hg(II) using GQDs-DPA supported amino acids by microfluidic paper-based (µPADs) device: multicolor plasmonic patterns, J. Environ. Chem. Eng., 2021, 9, 106197 CrossRef CAS.
  304. X. Gao, Z. Ma, M. Sun, X. Liu, K. Zhong, L. Tang, X. Li and J. Li, A highly sensitive ratiometric fluorescent sensor for copper ions and cadmium ions in scallops based on nitrogen doped graphene quantum dots cooperating with gold nanoclusters, Food Chem., 2022, 369, 130964 CrossRef CAS PubMed.
  305. Y. Wang, Q. He, X. Zhao, J. Yuan, H. Zhao, G. Wang and M. Li, Synthesis of corn straw-based graphene quantum dots (GQDs) and their application in PO43− detection, J. Environ. Chem. Eng., 2022, 10, 107150 CrossRef CAS.
  306. Z. Liu, J. Xiao, X. Wu, L. Lin, S. Weng, M. Chen, X. Cai and X. Lin, Switch-on fluorescent strategy based on N and S co-doped graphene quantum dots (N-S/GQDs) for monitoring pyrophosphate ions in synovial fluid of arthritis patients, Sens. Actuators, B, 2016, 229, 217–224 CrossRef CAS.
  307. H. Sammi, D. Kukkar, J. Singh, P. Kukkar, R. Kaur, H. Kaur, M. Rawat, G. Singh and K. H. Kim, Serendipity in solution–GQDs zeolitic imidazole frameworks nanocomposites for highly sensitive detection of sulfide ions, Sens. Actuators, B, 2018, 255, 3047–3056 CrossRef CAS.
  308. P. G. Oorimi, A. Tarlani, R. Zadmard and J. Muzart, Synthesis of photoluminescent composite based on graphene quantum dot@ZIF-11: a novel sensor for extremely efficient nano-molar detection of CN, Microchem. J., 2023, 189, 108494 CrossRef.
  309. N. Malahom, T. Ma-In, P. Naksen, W. Anutrasakda, M. Amatatongchai, D. Citterio and P. Jarujamrus, Nitrogen-doped graphene quantum dots as “off-on” fluorescent probes in paper-based test kits for selective monitoring of cyanide in food, ACS Appl. Nano Mater., 2023, 6, 11144–11153 CrossRef CAS.
  310. R. Liu, J. Zhao, Z. Huang, L. Zhang, M. Zou, B. Shi and S. Zhao, Nitrogen and phosphorus co-doped graphene quantum dots as a nano-sensor for highly sensitive and selective imaging detection of nitrite in live cell, Sens. Actuators, B, 2017, 240, 604–612 CrossRef CAS.
  311. P. Mohammadnejad, E. Jabbary, S. M. M. Hosseini, B. Sohrabi and M. N. Jamal, Facile synthesis and modifying of nitrogen-doped graphene quantum dots (N-doped GQDs) as fluorescent probes for detection of nitrite ions, Environ. Nanotechnol., Monit. Manage., 2023, 20, 100887 CAS.
  312. Y. C. Chen, W. H. Chiang, D. Kurniawan, P. C. Yeh, K. I. Otake and C. W. Kung, Impregnation of graphene quantum dots into a metal-organic framework to render increased electrical conductivity and activity for electrochemical sensing, ACS Appl. Mater. Interfaces, 2019, 11, 35319–35326 CrossRef CAS PubMed.
  313. N. Ndebele and T. Nyokong, The use of carbon-based nanomaterials conjugated to cobalt phthalocyanine complex in the electrochemical detection of nitrite, Diamond Relat. Mater., 2023, 132, 109672 CrossRef CAS.
  314. Y. Li, Y. Xiao, Q. Tao, M. Yu, L. Zheng, S. Yang, G. Ding, H. Dong and X. Xie, Selective coordination and localized polarization in graphene quantum dots: detection of fluoride anions using ultra-low-field NMR relaxometry, Chin. Chem. Lett., 2021, 32, 3921–3926 CrossRef CAS.
  315. S. R. Ahmed, M. Sherazee, S. Srinivasan and A. R. Rajabzadeh, Nanozymatic detection of thiocyanate through accelerating the growth of ultra-small gold nanoparticles/graphene quantum dots hybrids, Food Chem., 2022, 379, 132152 CrossRef.
  316. Y. Wang, Y. Hu, W. Weng, S. Chang, H. Xu, D. Li and D. Li, Nitrogen-doped graphene quantum dots based fluorescent probe for highly sensitive detection of thiosulfate anion and oxidative compounds, J. Photochem. Photobiol., A, 2021, 412, 113234 CrossRef CAS.
  317. K. Yang, L. Hou, Z. Li, T. Lin, J. Tian and S. Zhao, A mitochondria-targeted ratiometric fluorescent nanoprobe for imaging of peroxynitrite in living cells, Talanta, 2021, 231, 122421 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.