Next Article in Journal
The Cytotoxic Effectiveness of Thiourea-Reduced Graphene Oxide on Human Lung Cancer Cells and Fungi
Next Article in Special Issue
Role of Crystalline Si and SiC Species in the Performance of Reduced Hybrid C/Si Gels as Anodes for Lithium-Ion Batteries
Previous Article in Journal
Room Temperature Ammonia Gas Sensor Based on p-Type-like V2O5 Nanosheets towards Food Spoilage Monitoring
Previous Article in Special Issue
Modulating Direct Growth of Copper Cobaltite Nanostructure on Copper Mesh as a Hierarchical Catalyst of Oxone Activation for Efficient Elimination of Azo Toxicant
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sustainable Preparation of Graphene Quantum Dots for Metal Ion Sensing Application

UNESCO Chair on Desalination and Water Treatment, Center for Advanced Materials, Qatar University, Doha P.O. Box 2713, Qatar
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(1), 148; https://doi.org/10.3390/nano13010148
Submission received: 6 December 2022 / Revised: 18 December 2022 / Accepted: 21 December 2022 / Published: 28 December 2022
(This article belongs to the Special Issue Nanomaterials in Catalysis for Environmental and Energy Applications)

Abstract

:

Highlights

What are the main finding?
  • Preparation of GQDs from ethanolic extract of eucalyptus tree leaves.
  • Hydrothermal treatment of extract at different temperatures and times was carried out.
What is the implication of the main finding?
  • The GQDs were formed in the size range of 2–5 nm, as validated by TEM images.
  • Developed GQDs were successfully used for metal ion sensing.

Abstract

Over the past several years, graphene quantum dots (GQDs) have been extensively studied in water treatment and sensing applications because of their exceptional structure-related properties, intrinsic inert carbon property, eco-friendly nature, etc. This work reported on the preparation of GQDs from the ethanolic extracts of eucalyptus tree leaves by a hydrothermal treatment technique. Different heat treatment times and temperatures were used during the hydrothermal treatment technique. The optical, morphological, and compositional analyses of the green-synthesized GQDs were carried out. It can be noted that the product yield of GQDs showed the maximum yield at a reaction temperature of 300 °C. Further, it was noted that at a treatment period of 480 min, the greatest product yield of about 44.34% was attained. The quantum yields of prepared GQDs obtained after 480 min of treatment at 300 °C (named as GQD/300) were noted to be 0.069. Moreover, the D/G ratio of GQD/300 was noted to be 0.532 and this suggested that the GQD/300 developed has a nano-crystalline graphite structure. The TEM images demonstrated the development of GQD/300 with sizes between 2.0 to 5.0 nm. Furthermore, it was noted that the GQD/300 can detect Fe3+ in a very selective manner, and hence the developed GQD/300 was successfully used for the metal ion sensing application.

1. Introduction

With the world’s rapid development in several sectors, the possibility of excessive pollution ejection into the environment has increased [1,2,3]. As a result, environmental water pollution caused by harmful metal ions has developed as one of the most serious issues faced by humans [4,5]. In every aspect of daily life, water is crucial and as a result, it is extremely important to have access to clean and potable water. Wastewater such as produced water can be treated properly for various applications. Polluted water might contain hazardous metal ions such as Pb2+, Hg2+, Fe3+, Fe2+, Cu2+, Cd2+, Co2+, Ni2+, Al3+, and Ag+ [6,7,8,9,10], and these metal ions can combine to produce dangerous solution compounds which can turn out to be toxic [11,12,13]. These toxic compounds do not disintegrate and are non-degradable. As a result, these compounds keep building up in the soil and eventually reach human bodies [14,15]. Since these problems have become a serious issue, the field of toxic metal ion sensing has become more popular among the research communities [16,17,18]. To detect harmful metal ions, numerous types of sensing systems combined with various materials have been developed, and graphene-based materials are having good potential in sensing applications [19,20,21,22]. The main limitations of the conventional metal sensing methods include low selectivity, instability, difficult on-site sampling, and poor compatibility in the aqueous environment. Given that metal ions have a direct relationship with environmental water pollution, it is extremely desirable to design and develop effective as well as selective fluorescent sensors for metal ion detection applications [22]. Fe3+ ions have always drawn a lot of attention among the different heavy metal ions, as it is difficult to identify and remove them. In general, it is necessary to monitor and maintain a balanced iron proportion in biological as well as environmental systems [23]. As metal ions might cause adverse impacts on human as well as environmental health, the usage of GQDs as fluorescent probes to detect metal ions has received good research attention recently. The metal ion Fe3+ performs an important role in both biological as well as ecological processes. In biological systems, Fe3+ ions might coordinate with several types of regulatory proteins. Excessive Fe3+ ions may lead to cytotoxicity, creating their concentration a vital marker for Parkinson’s disease. Therefore, the proper monitoring of Fe3+ pollution is very crucial [24]. Hence, there is a requirement for highly sensitive and specific sensors of Fe3+ concentrations in both the environment and biological systems. Due to its fascinating characteristics, such as chemical inertness, low toxicity, high solubility, persistent photoluminescence (PL), good surface grafting, etc., graphene quantum dots (GQDs) have recently begun to be recognized as a material to develop an advanced novel type of fluorescence sensor. There are several studies carried out using GQDs for improving water treatment efficiency and works related to the environmental impact of nanomaterials [25,26].
GQD is a zero-dimensional (0-D) member in the carbon-based nanomaterials family and is generally considered a shredded fragment from a graphene sheet. This material has been studied widely since its unforeseen discovery in 2004, during the carbon nanotube’s purification, and it has a honeycomb structure with a single carbon layer [27]. It has exceptional electrical, structural, chemical, and tunable optical properties of PL as well as electrochemiluminescence. The better stability, biocompatibility, superior dispersibility, surface grafting, good solubility, non-toxicity, and inertness of the GQD materials enable its extensive potential in several applications [28]. In recent times, GQDs have been utilized to identify several analytes based on their luminescence characteristics. Many sensors have been based on the GQDs’ photoluminescence quenching process [29,30]. The application of optical characteristics of GQDs for dangerous metal ion sensing has gained the attention of researchers since they are very compatible in aquatic environments, inexpensive, straightforward, quick, efficient, and highly sensitive and selective [31,32,33]. Numerous studies on the integration of GQDs with the aforementioned optical methods have been completed successfully to efficiently identify harmful metal ions. These studies were motivated by the beneficial optical properties of GQDs [34,35,36,37,38,39]. Different preparation methodology has been utilized in the preparation of GQDs, and it can be categorized into two types, respectively, (1) top down and (2) bottom-up methods [40]. The top-down method involves the breaking down of bulk materials into small-sized nanostructured materials, and it included methods such as chemical exfoliation [41], solvothermal preparation [42], nanolithography [43], and electrochemical cutting [44]. Contrarily, the bottom-up method involving in the formation of bigger units from small units, and it comprises techniques such as the cage-opening technique [45], hydrothermal heating [46], microwave irradiation [47], or thermal combustion [48]. In spite of the fact that the above-mentioned techniques have several benefits, these preparation methods require complex purification techniques, harmful organic solvents, low quantum yield, elevated temperature, treatment with concentrated acid or alkali, and superior quality carbon precursors. Even though different naturally occurring carbon species such as carbon black [49], carbon fibers [50], coal [51], etc., can be used as a precursor for the preparation of GQDs, these materials are also associated with fossil fuels which are considered as a non-renewable source and may have the possibility that it will not be adequately available in future. Therefore, green chemistry methods are used recently for the preparation of nanomaterials, which contribute additional benefits such as the possibility of large-scale production, biocompatibility, unique morphologies, environmental-friendly preparation, economic nature, the vast availability of different carbon sources, and the possibility of recycling the waste products into beneficial products [52].
The green-synthesized GQDs from various carbon sources such as mango leaves [53], tea waste [54], flower extract [55], cow’s milk [56], etc., were already reported. The utilization of the above-mentioned precursors for the preparation of GQDs has several benefits such as non-toxicity, easy handling nature, and large-scale availability. The results from the above-mentioned studies encouraged us to investigate whether we can develop the GQDs from green plants, which are the foundation of the majority of the Earth’s ecologies. Certain biomolecules, proteins, polysaccharides, vitamins, and enzymes in plants have a good capability for performing reduction as well as capping of non-biocompatible materials. Plant-based materials are proven to be an exceptional source for the bio-based preparation of carbon nanomaterials since these materials have high carbon content for the preparation of carbon-based nanomaterials. In recent times, carbon nanotubes have been synthesized from natural sources such as turpentine and eucalyptus [57]. Eucalyptus trees are now grown all over the globe and used extensively for their medicinal properties. Eucalyptus tree leaves are a byproduct of the trees, and these byproduct leaves can be used for nanomaterial synthesis and hence turn out to be a significant research subject. The leaf extracts of the Eucalyptus tree are mostly comprised of hydrocarbons having low oxygen content [58], and this will enable the leaf extract to be an ideal precursor to synthesize the GQDs. Dubeey et al. [59] claimed the application of methanol extract of eucalyptus hybrid leaves for the extra-cellular green preparation of silver(Ag) nanoparticles. In research carried out by Weng et al. [60], hybrid nonoxidized graphene oxide or iron-based nanoparticles were prepared by means of a green synthesis technique, using a single-step process with the leaf extract of Eucalyptus trees. Moreover, Zhuang et al. [61] prepared reduced graphene oxide from Eucalyptus leaf extract as these leaves are abundant in nature and the oxidized products are eco-friendly. According to another research by Wang et al. [62], the group illustrated that the aqueous leaf extracts of Eucalyptus trees were employed as a stabilizing and non-oxidizing agent for the preparation of iron nanoparticles.
The above-mentioned studies encouraged us to select the eucalyptus tree leaves as the precursor for the green preparation of GQDs. To the best of our knowledge, the leaf extract of Eucalyptus trees has never been used as the precursor for the sustainable development of GQDs to be used in metal ion sensing applications. This is a very sustainable way of preparing the GQDs, where the ethanol used will be recovered during the GQD preparation stage itself. The formation of GQDs was confirmed by UV–Visible spectrophotometry and PL. The morphology, as well as the surface properties of green-synthesized GQDs, were characterized by different techniques such as transmission electron microscopy (TEM), Fourier transform infrared spectroscopy (FTIR), X-ray diffraction analysis (XRD), and Raman spectroscopy analysis. The major objectives of the present work were to: (1) prepare the GQDs from eucalyptus leaves (2) characterize the green synthesized GQDs using different techniques and (3) use the developed GQDs for metal sensing application.

2. Experimental Section

2.1. Materials

Eucalyptus tree leaves were acquired from the Eucalyptus tree on the university grounds of Qatar University, Qatar. The ethanol was procured from Sigma Aldrich, St. Louis, MO, USA. Distilled water and ethanol used were obtained from Merck, Kenilworth, NJ, USA. The salts AgCl, FeCl3, NaCl, LiCl, AlCl3, CaCO3, CuSO4, CoCl2, CrCl3, MgSO4, SrCl2, MnCl2, MoCl2, FeCl2, ZnCl2 were supplied by Sigma Aldrich, St. Louis, MO, USA. Quinine sulfate and sulfuric acid were purchased locally from Sulfur chemicals, Doha, Qatar.

2.2. GQD Synthesis from Leaves of Eucalyptus Tree Employing Ethanol

Eucalyptus tree leaves were collected, dried in oven for 5 h, and then ball milled the leaves to obtain fine powder. Subsequently, 10 gms of Eucalyptus tree leaves powder was mixed with pure ethanol solution, and this mixture was placed for continuous stirring for a time of 4 h at room temperature, and the resulting extract was centrifuged for 10 min at 8000 rpm for achieving a fine supernatant. Further, the obtained extract was filtered using a 0.22 micrometer filter and consequently concentrated by the evaporation of ethanol using a rotary evaporator till the residue slurry was obtained. This slurry was blended with a less sum of milli-Q water and heated in an oven at different temperatures (240, 260, 300, 320 °C) for different treatment times (360, 400, 440, 480, and 520 min), and subsequently, all sample residue was dispersed in absolute ethanol for properly dispersing the GQDs. The obtained dispersion was then filtered out using the syringe filter (0.22 μm) for obtaining the pure GQDs. These GQDs were then permitted for drying for 24 h at 65 °C to obtain the dried powder. The graphical diagram for the GQD synthesis from Eucalyptus tree leaves is shown in Figure 1.

2.3. Characterization of the Prepared GQDs

The FTIR instrument used in the current study was 760 Nicolet, and it permitted the identification of inorganic and organic groups present in the sample, based on their particular IR frequency. The X-ray diffraction pattern of the developed GQDs was obtained using XRD: PANalytical, EMPYREAN using Cu/Ka radiation with a 1.54-angstrom wavelength. To examine the degree of structural defects and crystallite size of GQDs, Raman analysis was carried out, with a laser power of 10 mW and wavelength of 532 nm. A fluorescence spectrophotometer was used to investigate the photo-luminescence behavior of GQDs. (Horiba FluoroMax-4 Spectrofluorometer), and absorption spectra were recorded in a UV−Visible spectrophotometer (Biochrom). The GQD morphology was examined employing TEM (HT 770, Hitachi, Japan).

2.4. Product Yield and Quantum Yield of the Prepared GQDs

The product yield of the developed GQDs was determined using Equation (1), below.
Y i e l d ( % ) = W e i g h t   o f   D r i e d   G Q D   o b t a i n e d W e i g h t   o f   S l u r r y   × 100
The quantum yield (QY) of the developed GQDs was determined using Equation (2), below.
Φ = Φ r I ( A r ) n 2 I r ( A ) n r 2
where Φr and Φ are the QYs of the standard reference and sample, Ir and I are the integrated photoluminescence intensities of the reference and sample, A and Ar are the absorbance values of sample and reference, and n and nr are the refractive indices of the sample and reference, respectively.

2.5. Application of GQD/300 as PL Metal Ion Sensor

In the present research work, the selective sensing of Fe3+ was carried out by 55 μg mL−1 GQD/300 solution employing 100 µM concentration of distinct metal ions: 100 µM concentrations of the following metal ions to accomplish the selective sensing of Fe3+: Co2+, Zn2+, Ag+, Na+, Mg2+, Mo2+, Sr2+, Fe3+, Li+, Ca2+, Mn2+, Al3+, Cr3+, Fe2+.

3. Results and Discussion

In the following section, we discuss the results of the optical, morphological, and structural characterization of GQDs, developed from the Eucalyptus tree leaves. Moreover, the product yield and quantum yield of the developed GQDs were also determined. Additionally, the efficiency of the developed GQDs in the metal sensing application.

3.1. Optical Characterization of GQDs

In the current study, a detailed study was performed at various temperatures for determining the effective work parameters (240–320 °C) and processing times (360 min) by observing the products’ PL emission and optical characteristics. To compare the PL intensities of the GQDs produced at various temperatures, same concentration of 55 g mL−1 was adopted. The PL emission intensities of the developed GQD at various temperatures (240, 260, 300, and 320 °C) at a predetermined period of time of 120 min was investigated and presented in Figure 2a. It was noted that the products obtained at 260 °C temperature showed less intense optical emission, indicating that below this temperature the cutting of carbon domains to develop GQDs will not induce good results. Therefore, the temperature was further increased to 320 °C [63,64]. It was observed in Figure 2a that the PL emission intensity enhanced with temperature up to 300 °C, whereas it declined with an additional increase in temperature to 320 °C. Additionally, the PL emission intensities of the developed GQD at different heat-treatment durations (360, 400, 440, 480, and 520 min) at a constant temperature of 300 °C are shown in Figure 2b. It was noted that the PL intensity enhanced with an increment in the treatment time up to 480 min and reduced thereafter (Figure 2b). These results suggest that increasing the temperature and the processing time can speed up the transformation of carbon domains into nanosized GQDs and enhance optical emission. Exceeding a certain temperature and processing time cutoff point might lead to GQD’s surface degradation and structural degradation and reduced PL emission. According to these results, it can be confirmed that the optimal conditions for the preparation of good quality GQDs with superior optical properties are a temperature value of 300 °C and a processing time of 480 min.
The PL spectra of the developed GQD/300 are shown in Figure 2c. Examining the photoluminescence spectra of the generated nanostructured material at various excitation wavelengths between 320 and 400 nm revealed that the emission intensity increased to 360 nm and then decreased. The photoluminescence emission intensity decreased in direct proportion to the increase in excitation wavelength. Since the excitation wavelength increases from 300 to 400 nm, the photoluminescence peaks change to larger wavelengths, indicating a redshift (430 nm to 480 nm) [65]. At 360 nm wavelength, the excitation-dependent photoluminescence of GQDs has been observed. The highest PL intensity of GQD/300 is 434 nm, with a vibrational relaxation or dissipation of the wavelength at 174 nm. All of the GQD fluorescence studies are consistent with what has been reported in the previous findings [24,66]. The presence of conjugated aromatic hydrocarbons, existence of hydroxyl and other functional groups containing oxygen, emission of inhibited zigzag edge with carbine-like triple ground state as well as the emission that is trapped on the surfaces are considered to be the major reasons for the fluorescence emission mechanism of GQDs [67,68,69]. The PL characteristic of GQDs at excited state might be due to optical selection of GQDs at different sizes and defects of GQDs on the surface level [70,71]. According to a study [72], the key reason for the fluorescence in excited stage is that when the emission occurs from the carbon backbone of sp2, the sp2 conjugated domain of GQDs is adequate to have a limited energy gap in the band due to the effect quantum confinement.
To additionally understand, the optical properties of GQDs prepared at 300 °C after 480 min of treatment (named GQD/300), these developed GQDs were examined thoroughly by UV–Vis absorption as well as photoluminescence emission spectroscopies. The optical properties of synthesized GQD/300 were explored by conducting the UV–Vis absorption. Usually, the absorption spectrum of graphene quantum dots appears in the UV region, and the tail extends toward the visible region. UV–Visible spectrum of GQD/300 as in Figure 3 shows a significant absorption at 300–320 nm which can be related to the π-π * and n-π * transition arising from aromatic C=C bonds (sp2 domain) and C=O groups, respectively [73]. The results suggested that the UV absorbance in the GQDs is associated with their surface oxygenated (C=O) states established at the time of reaction [74].

3.2. Product Yield of GQDs

A thorough investigation was carried out to examine the impact of temperature and reaction time on the GQD product yield. In accordance with the precursor of the eucalyptus extract, the product yield for each treatment was determined, and the results are summarized in Table 1. It can be noted that the product yield of GQDs showed the maximum yield at a reaction temperature of 300 °C. This GQD of reaction temperature of 300 °C was checked at different reaction times (Table 2), and it was noted that at a treatment period of 480 min, the greatest product yield of about 44.34% is attained.
According to these findings, greater temperatures may cause carbon domains to cut into GQDs more quickly. These are the greatest yield figures for products that have been reported thus far for the preparation of GQDs from biomass waste to the best of the authors’ knowledge [75,76,77]. The yield of GQDs made from the various biomass-based precursors covered in the prior studies is summarized in the table.

3.3. Quantum Yield of GQDs

Quantum yield is considered to be a significant aspect related to the PL of GQDs. In most cases, the QY depends on the synthesis techniques as well as surface chemistry. By comparing the integrated PL intensities and absorbance values of GQDs and quinine sulfate, the quantum yields of GQDs were estimated. A quantum yield of 0.53 was shown by dissolved quinine sulfate in 0.1 M H2SO4. To prevent re-absorption effects, as-prepared GQD/300 was dissolved in water with concentrations adjusted to yield an absorbance value lower than 0.1. Slit widths for both excitation and emission were fixed at 4.0 nm. The quantum yields of developed GQD/300 obtained after 480 min of treatment were determined to be 0.069. Research works confirmed that although oxygen-containing functional groups present in GQDs make the material hydrophilic and offer sites for additional chemical functionalization, they also act as emissive traps resulting in reduced quantum yield.

3.4. Structural and Morphological Characterization of GQDs

The TEM images of the synthesized GQD/300 are shown in Figure 4a,b. The structure of GQD/300 examined by TEM confirmed that the GQD/300 were noted to have a size in the range of 2 to 5 nm. Figure 4c shows the particle size distribution of GQD/300. This is in line with the results obtained by a research work completed by Kumawat et al. [53]. The synthesis of the GQD/300 may be because of the carbonization of the solution during the heat treatment in the autoclave. The material’s carbonization degree will support controlling the size of the GQD/300 developed. The resultant GQDs are single dispersed sphere particles, as observed. The mean size of the GQD/300 particle is (Figure 4c) around 3 nm, with a comparatively tapered size distribution between 3–5 nm (based on statistic evaluation of over 100 quantum dots).
The XRD profiles in Figure 4d show that the GQD/300 have single broad diffraction peaks centered around 21°, which is attributed to the (002) lattice spacing of carbon-based materials with amorphous nature, a graphitic structure, suggesting that carbonizing eucalyptus powder produces graphitic structures. The peak is broad due to the small size of the GQD/300, and it is a typical band corresponding to highly disordered amorphous graphene quantum dots [35,36,78]. In several circumstances, the XRD patterns of GQDs are comparable to the bio-mass precursor peak with a minimal shift, as the main component of these materials is carbon, the only variation that matters are the peak intensity and broadness of the signal. The key difference between the precursor and resulting materials is peak sharpness and the XRD result shows the peak widening, which suggests the development of an amorphous nanoscale GQD structure.
The chemical bonding states of GQD/300 were investigated and characterized using FTIR spectra (Figure 4e). The FTIR spectra of GQD/300 exhibited strong absorption bands as seen in Figure 4e. The GQD/300 exhibited stretching vibrations of the carbonyl group –C=O at 1633 cm−1, hydroxyl group –OH at 3390.0 cm−1, and C=C stretching vibrations at 1498 cm−1 and –CH2 stretching at 2929 cm−1. Thus, it can be suggested that during hydrothermal polymerization, a carbonization reaction has occurred. The C–O bond stretching caused the peak to occur at 1211 cm−1. All of the findings matched with the results observed in earlier investigations [79,80,81].
GQD/300 Raman spectroscopy result is shown in Figure 4f and it revealed two distinct bands called D and G bands. The band D at 1354 cm−1 is linked to the crystalline characteristic of the compound and the vibrational characteristic of carbon atoms with dangling bonds. However, the band G at 1577 cm−1 is assigned to the crystalline nature of the compound as well as E2g vibration on photon mode of sp2 hybridization of carbon atom in the 2D hexagonal lattice of graphite structure(D,G) [82,83]. The ratio of the intensities of the disordered D band and the amorphous G band (D/G) is a typical approach to evaluate the homogeneity (degree of disorder or graphitization) of a quantum dot sample. The D/G ratio of the amorphous quantum-dots sample is greater. A reduced D/G ratio confirms a greater degree of graphitization in the specimen. In the current study, the D/G ratio of GQD/300 was noted to be 0.532 and this suggested that graphene quantum dots have a nano-crystalline graphite structure, which is nearly the same as the results formerly published [38,84,85].

3.5. Application of GQD/300 as PL Metal Ion Sensor

By examining the variations in fluorescence intensity with different metal ion concentrations (1 μM, 6 μM, 11 μM, 17 μM, 23 μM, 28 μM, 32 μM, 37 μM, the variation in fluorescence intensity of GQD/300 was tracked. The variation in intensity was used to estimate the sensitive response of GQD/300. The analytical performance of the GQD/300 sensing system was performed in order to show the method’s sensitivity as well as for the quantitative study of the PL response of the GQD/300 towards different metal ions. The affinity (coordination) of several metal ions toward GQD/300 was assessed thoroughly under comparable experimental conditions. Figure 5b demonstrates that among sixteen different metal ions, the Fe3+ ions have the largest affinity (Figure 5b) for GQD/300, indicating the possibility of selective sensing by GQD/300.
When illuminated, charge carriers in GQD/300 moved from the highest occupied molecular orbital to the lower unoccupied molecular orbital and then fluorescence appeared. Additionally, GQD/300’s primary dependence for their PL emission wavelength was particle size, which showed its effect on energy gap variation. GQDs that show smaller bandgap will be greater in size and vice versa [23,86]. For GQD/300 containing metal ion adsorbent, photoluminescence measurements were made at room temperature with an excitation wavelength of 360 nm. Interestingly, it was discovered that the maximal PL intensity of the GQD/300/Fe3+ combination significantly decreased upon increasing the metal ion concentration as reported by many authors. The generated histogram clearly demonstrated that the addition of metal ions caused PL alterations in GQD/300. With the inclusion of 1 µM of the metal ion Fe3+, a maximum 24% reduction in PL intensity for the GQD/300/Fe3+ combination is seen. The intensity of the PL peak was reduced by 37% by raising the Fe3+ quencher ion concentration from 1 µM to 6 µM. This demonstrated that the GQD/300 is extremely sensitive to Fe3+ even at very low quencher concentrations (1 µM). The GQD/300 displays significant PL emission and distinctive oxygenated surface functionalization in their as-synthesized state. Because of the presence of hydrophilic functional groups, the GQD/300 has good solubility in water. These GQDs are anticipated to be the best option for fluorescence sensing due to the existence of these functional groups as well as strong emissions. Several scientists used the quenching of carbon QDs by the addition of Fe3+ to accomplish this method. Fe3+ concentration and quenching % are correlated. In order to track how metal ions affect the intensity of the PL, a 360 nm wavelength was chosen. Different metal ions were added to the GQD/300 PL intensity to record them. According to Figure 5b, only Fe3+ ions out of a total of sixteen different metal ions significantly reduced the PL of GQDs when compared to the control (blank) sample. These findings imply that the GQD/300 can detect Fe3+ in a very selective manner. These findings can be explained as the functional groups located on the margins and base of GQDs’ increasing binding affinity. When ligands such as carboxyl and hydroxyl groups are present, they serve as a choice coordination site for the metal ions to diffuse and adhere to the GQD/300 and due to variations in charge transfer, changes in the local density of states, and changes in optical characteristics, this will result in a narrowing of the energy gap between GQD/300 and quencher. This results in static PL quenching without spectra shift which can be observed in Figure 5a. The significant affinity of Fe3+ ions toward the hydroxyl/carboxyl groups of GQD/300, which results in a stable complex, is connected to the mechanism of PL quenching. Figure 5c presents the fluorescent quenching mechanism of the GQD/300 in presence of the Fe3+ metal ions.
In general, the bare GQDs are very sensitive to several metal ions, and some transition or heavy metals could selectively quench fluorescence [87]. Other than the surface functionalized GQDs (that generate a specific combination of Fe3+), bare GQDs are also very selective for Fe3+ ions. Research works on the selective detection of Fe3+ have concentrated mainly on how Fe3+ quenches GQDs fluorescence. The unique coordination interaction between the phenolic hydroxy groups of the GQDs and Fe3+ ions or the energy/electron transfer process has been extensively used to describe the basis of fluorescence quenching [88]. Furthermore, few authors have explained this situation taking into account the pH values of GQDs solutions. Nevertheless, comprehensive justifications have not been suggested for why metal ions can be selectively detected employing the bare GQDs. Such a particular mechanism needs to be studied further. The majority of studies have recommended that the recognition mechanism arises from the specific coordination interaction between phenolic hydroxyl groups on the GQD surface and Fe3+ [89]. The subsequent electron-transfer process between the GQDs and Fe3+ brings fluorescence quenching dropping their fluorescence lifetime. Apart from inducing dynamic quenching, the GQDs could also develop complexes with Fe3+ for inducing the static quenching. Even though certain studies have employed this sensing mechanism, the idea that Fe3+ could be specially identified from other metal ions still lacks systematic research as well as experimental support [90].
During illumination, the metal ion was bonded to the GQD/300 surface by functional groups with electrostatic attraction or non-covalent bonds. Following the identification of particular metal ions (Fe3+), the distance between GQDs and Fe3+ ions was reduced, strengthening the GQD/300-Fe3+ interaction. This strongly promotes charge transfer between GQD/300 and Fe3+ ions, hence reducing GQD fluorescence. This quenching effect results from the inherent emission effect of excited state electron transfer by the absorption of photons between the metal ions and the fluorophore, which promotes electron–hole pair formation. The concentration of the target analyte affects the PL quenching of GQD/300 as well, allowing for the sensitive detection of deleterious metal ions. It has already been mentioned that until they are close to each other, fluorescence quenching of the fluorophore with the presence of metal ions is feasible in both the conjugated system and the unbonded state. A charge transfer-based sensing technique will have great sensitivity due to the high quenching efficiency of metal ions to the fluorescence of GQD/300. The probability of achieving a nonradiative transition increases with decreasing the distance between the lowest unoccupied (LUMO) and highest occupied (HOMO) molecular orbitals which results in a smaller emission gap [91]. Furthermore, because of structural relaxation during photoexcitation, the emission gap shrinks and overtakes the absorption gap. On the surface of the GQD/300, functional groups such as carboxyl, carbonyl, and hydroxyl may greatly promote the nonradiative transition through structural vibrations [92].

4. Conclusions

In the current study, we demonstrated a rapid, accessible, and inexpensively viable green preparation of GQDs using Eucalyptus tree leaves as the precursor. The current study reported on the synthesis of GQDs from the ethanolic extracts of eucalyptus tree leaves by a hydrothermal treatment technique at different heat treatment times and temperatures. This is a very sustainable way of preparing the GQDs, where the ethanol used will be recovered during the GQD preparation stage itself. The optical, morphological, and compositional analyses of the green-synthesized GQDs were carried out. It was observed that the product yield of GQDs showed the maximum yield at a reaction temperature of 300 °C. Furthermore, it was noted that at a treatment period of 480 min, the greatest product yield of about 44.34% was attained. The quantum yields of prepared GQD/300 obtained after 480 min of treatment at 300 °C were calculated to be 0.069. Moreover, the D/G ratio of GQD/300 was noted to be 0.532 and this suggested that the GQD/300 developed has a nano-crystalline graphite structure. The TEM images demonstrated the development of GQD/300 with sizes between 2.0 to 5.0 nm. Moreover, it was noted that the GQD/300 can detect Fe3+ in a very selective manner, and hence the developed GQDs were successfully used for the metal ion sensing application. The emerging field of GQD sensors for metal ion species should be studied more, with a perspective on the future of this extremely versatile material.

Author Contributions

Conceptualization, A.S., N.M. and S.J.Z.; methodology, A.S.; software, A.S.; validation, A.S., H.S. and S.J.Z.; formal analysis, A.S.; investigation, H.S.; resources, H.R.S.; data curation, A.A.S.; writing—original draft preparation, A.S.; writing—review and editing, H.S.; visualization, A.S.; supervision, S.J.Z.; project administration, H.S.; funding acquisition, S.J.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Qatar University, grant number QUEX-CAM-QP-PW-18/19.

Data Availability Statement

Not Applicable.

Acknowledgments

The author would like to thank the support from the project QUEX-CAM-QP-PW-18/19 for this research. The TEM analysis was accomplished in the Central Laboratories unit, Qatar University. The findings achieved herein are solely the responsibility of the authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Moossa, B.; Trivedi, P.; Saleem, H.; Zaidi, S.J. Desalination in the GCC countries-a review. J. Clean. Prod. 2022, 357, 131717. [Google Scholar] [CrossRef]
  2. Saleem, H.; Trabzon, L.; Kilic, A.; Zaidi, S.J. Recent advances in nanofibrous membranes: Production and applications in water treatment and desalination. Desalination 2020, 478, 114178. [Google Scholar] [CrossRef]
  3. Dawoud, H.D.; Saleem, H.; Alnuaimi, N.A.; Zaidi, S.J. Characterization and Treatment Technologies Applied for Produced Water in Qatar. Water 2021, 13, 3573. [Google Scholar] [CrossRef]
  4. Saleem, H.; Zaidi, S.J. Nanoparticles in reverse osmosis membranes for desalination: A state of the art review. Desalination 2020, 475, 114171. [Google Scholar] [CrossRef]
  5. Soliman, M.N.; Guen, F.Z.; Ahmed, S.A.; Saleem, H.; Zaidi, S.J. Environmental Impact Assessment of Desalination Plants in the Gulf Region. In Water-Energy-Nexus in the Ecological Transition; Springer: Cham, Switzerland, 2002; pp. 173–177. [Google Scholar]
  6. Ruiyi, L.; Huahua, Z.; Zaijun, L. Switchable Two-Color Graphene Quantum Dot as a Promising Fluorescence Probe to Highly Sensitive PH Detection and Bioimaging. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2022, 275, 121028. [Google Scholar] [CrossRef]
  7. Yao, Q.; Wu, H.; Jin, Y.; Wang, C.; Zhang, R.; Lin, Y.; Wu, S.; Hu, Y. One-Pot Synthesis of Fluorescent Nitrogen-Doped Graphene Quantum Dots for Portable Detection of Iron Ion. Curr. Appl. Phys. 2022, 41, 191–199. [Google Scholar] [CrossRef]
  8. Luhana, C.; Moyo, I.; Tshenkeng, K.; Mashazi, P. In-Sera Selectivity Detection of Catecholamine Neurotransmitters Using Covalent Composite of Cobalt Phthalocyanine and Aminated Graphene Quantum Dots. Microchem. J. 2022, 180, 107605. [Google Scholar] [CrossRef]
  9. Zaidi, S.J.; Saleem, H. Reverse Osmosis Systems: Design, Optimization and Troubleshooting Guide; Elsevier: Amsterdam, The Netherlands, 2021. [Google Scholar]
  10. Ding, J.; Zhou, X.; Huang, Y.; Chen, B.; Chen, S.; Jin, Y.; Yang, Y.; Pan, N.; Xu, C.; Chen, J.; et al. An Innovative Strategy for Construction of PH-Responsive Supramolecular Hydrogel from Graphene Quantum Dots Clusters toward Integration of Detection and Removal of Uranium. Appl. Surf. Sci. 2022, 583, 152492. [Google Scholar] [CrossRef]
  11. Saleem, H.; Zaidi, S.J. Recent developments in the application of nanomaterials in agroecosystems. Nanomaterials 2020, 10, 2411. [Google Scholar] [CrossRef] [PubMed]
  12. Ghazi, Z.M.; Rizvi, S.W.F.; Shahid, W.M.; Abdulhameed, A.M.; Saleem, H.; Zaidi, S.J. An overview of water desalination systems integrated with renewable energy sources. Desalination 2022, 542, 116063. [Google Scholar] [CrossRef]
  13. Saleem, H.; Zaidi, S.J.; Ismail, A.F.; Goh, P.S.; Vinu, A. Recent advances in the application of carbon nitrides for advanced water treatment and desalination technology. Desalination 2022, 542, 116061. [Google Scholar] [CrossRef]
  14. Bakly, S.; Ibrar, I.; Saleem, H.; Yadav, S.; Al-Juboori, R.; Naji, O.; Zaidi, S.J. Polymer-based nano-enhanced forward osmosis membranes. In Advancement in Polymer-Based Membranes for Water Remediation; Elsevier: Amsterdam, The Netherlands, 2022; pp. 471–501. [Google Scholar]
  15. Saud, A.; Saleem, H.; Zaidi, S.J. Progress and Prospects of Nanocellulose-Based Membranes for Desalination and Water Treatment. Membranes 2022, 12, 462. [Google Scholar] [CrossRef]
  16. Shen, Y.; Rong, M.; Qu, X.; Zhao, B.; Zou, J.; Liu, Z.; Bao, Y.; He, Y.; Li, S.; Wang, X.; et al. Graphene Oxide-Assisted Synthesis of N, S Co-Doped Carbon Quantum Dots for Fluorescence Detection of Multiple Heavy Metal Ions. Talanta 2022, 241, 123224. [Google Scholar] [CrossRef]
  17. Zhu, Q.; Li, R.; Sun, X.; Zaijun, L. Highly Sensitive and Selective Electrochemical Aptasensor with Gold-Aspartic Acid, Glycine Acid-Functionalized and Boron-Doped Graphene Quantum Dot Nanohybrid for Detection of α-Amanitin in Blood. Anal. Chim. Acta 2022, 1219, 340033. [Google Scholar] [CrossRef]
  18. Mahajan, M.R.; Patil, P.O. Design of Zero-Dimensional Graphene Quantum Dots Based Nanostructures for the Detection of Organophosphorus Pesticides in Food and Water: A Review. Inorg. Chem. Commun. 2022, 144, 109883. [Google Scholar] [CrossRef]
  19. Shangguan, Q.; Chen, Z.; Yang, H.; Cheng, S.; Yang, W.; Yi, Z.; Wu, X.; Wang, S.; Yi, Y.; Wu, P. Design of ultra-narrow band graphene refractive index sensor. Sensors 2022, 22, 6483. [Google Scholar] [CrossRef]
  20. Cheng, Z.; Liao, J.; He, B.; Zhang, F.; Zhang, F.; Huang, X.; Zhou, L. One-step fabrication of graphene oxide enhanced magnetic composite gel for highly efficient dye adsorption and catalysis. ACS Sustain. Chem. Eng. 2015, 3, 1677–1685. [Google Scholar] [CrossRef]
  21. Chen, H.; Chen, Z.; Yang, H.; Wen, L.; Yi, Z.; Zhou, Z.; Dai, B.; Zhang, J.; Wu, X.; Wu, P. Multi-mode surface plasmon resonance absorber based on dart-type single-layer graphene. RSC Adv. 2022, 12, 7821–7829. [Google Scholar] [CrossRef]
  22. Zhang, Z.; Cai, R.; Long, F.; Wang, J. Development and application of tetrabromobisphenol A imprinted electrochemical sensor based on graphene/carbon nanotubes three-dimensional nanocomposites modified carbon electrode. Talanta 2015, 134, 435–442. [Google Scholar] [CrossRef]
  23. Nair, R.V.; Thomas, R.T.; Sankar, V.; Muhammad, H.; Dong, M.; Pillai, S. 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. [Google Scholar] [CrossRef]
  24. Ananthanarayanan, A.; Wang, X.; Routh, P.; Sana, B.; Lim, S.; Kim, D.H.; Lim, K.H.; Li, J.; Chen, P. Facile synthesis of graphene quantum dots from 3D graphene and their application for Fe3+ sensing. Adv. Funct. Mater. 2014, 24, 3021–3026. [Google Scholar] [CrossRef]
  25. Saleem, H.; Zaidi, S.J. Developments in the application of nanomaterials for water treatment and their impact on the environment. Nanomaterials 2020, 10, 1764. [Google Scholar] [CrossRef] [PubMed]
  26. Saleem, H.; Zaidi, S.J.; Ismail, A.F.; Goh, P.S. Advances of nanomaterials for air pollution remediation and their impacts on the environment. Chemosphere 2022, 287, 132083. [Google Scholar] [CrossRef] [PubMed]
  27. Saleem, H.; Saud, A.; Munira, N.; Goh, P.S.; Ismail, A.F.; Siddiqui, H.R.; Zaidi, S.J. Improved Forward Osmosis Performance of Thin Film Composite Membranes with Graphene Quantum Dots Derived from Eucalyptus Tree Leaves. Nanomaterials 2022, 12, 3519. [Google Scholar] [CrossRef]
  28. Saleem, H.; Goh, P.S.; Saud, A.; Khan, M.A.W.; Munira, N.; Ismail, A.F.; Zaidi, S.J. Graphene Quantum Dot-Added Thin-Film Composite Membrane with Advanced Nanofibrous Support for Forward Osmosis. Nanomaterials 2022, 12, 4154. [Google Scholar] [CrossRef]
  29. Li, B.; Xiao, X.; Hu, M.; Wang, Y.; Wang, Y.; Yan, X.; Huang, Z.; Servati, P.; Huang, L.; Tang, J. Mn, B, N Co-Doped Graphene Quantum Dots for Fluorescence Sensing and Biological Imaging. Arab. J. Chem. 2022, 15, 103856. [Google Scholar] [CrossRef]
  30. Wang, Y.; He, Q.; Zhao, X.; Yuan, J.; Zhao, H.; Wang, G.; Li, M. Synthesis of Corn Straw-Based Graphene Quantum Dots (GQDs) and Their Application in PO43- Detection. J. Environ. Chem. Eng. 2022, 10, 107150. [Google Scholar] [CrossRef]
  31. Quyen, T.T.B.; My, N.N.T.; Pham, D.T.; Thien, D.V.H. Synthesis of TiO2 Nanosheets/Graphene Quantum Dots and Its Application for Detection of Hydrogen Peroxide by Photoluminescence Spectroscopy. Talanta Open 2022, 5, 100103. [Google Scholar] [CrossRef]
  32. Tran, H.L.; Dang, V.D.; Dega, N.K.; Lu, S.M.; Huang, Y.F.; Doong, R. An Ultrasensitive Detection of Breast Cancer Cells with a Lectin-Based Electrochemical Sensor Using N-Doped Graphene Quantum Dots as the Sensing Probe. Sens. Actuators B Chem. 2022, 368, 132233. [Google Scholar] [CrossRef]
  33. Li, K.; Tu, J.; Zhang, Y.; Jin, D.; Li, T.; Li, J.; Ni, W.; Xiao, M.M.; Zhang, Z.Y.; Zhang, G.J. Ultrasensitive Detection of Exosomal MiRNA with PMO-Graphene Quantum Dots-Functionalized Field-Effect Transistor Biosensor. Iscience 2022, 25, 104522. [Google Scholar] [CrossRef]
  34. Liang, Z.; Kang, M.; Payne, G.F.; Wang, X.; Sun, R. Probing Energy and Electron Transfer Mechanisms in Fluorescence Quenching of Biomass Carbon Quantum Dots. ACS Appl. Mater. Interfaces 2016, 8, 17478–17488. [Google Scholar] [CrossRef] [PubMed]
  35. Wu, F.; Su, H.; Wang, K.; Wong, W.K.; Zhu, X. Facile Synthesis of N-Rich Carbon Quantum Dots from Porphyrins as Efficient Probes for Bioimaging and Biosensing in Living Cells. Int. J. Nanomed. 2017, 12, 7375–7391. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Bourlinos, A.B.; Bakandritsos, A.; Kouloumpis, A.; Gournis, D.; Krysmann, M.; Giannelis, E.P.; Polakova, K.; Safarova, K.; Hola, K.; Zboril, R. Gd(III)-Doped Carbon Dots as a Dual Fluorescent-MRI Probe. J. Mater. Chem. 2012, 22, 23327–23330. [Google Scholar] [CrossRef]
  37. Chhabra, V.A.; Kaur, R.; Kumar, N.; Deep, A.; Rajesh, C.; Kim, K.H. Synthesis and Spectroscopic Studies of Functionalized Graphene Quantum Dots with Diverse Fluorescence Characteristics. RSC Adv. 2018, 8, 11446–11453. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Mewada, A.; Pandey, S.; Shinde, S.; Mishra, N.; Oza, G.; Thakur, M.; Sharon, M.; Sharon, M. Green Synthesis of Biocompatible Carbon Dots Using Aqueous Extract of Trapa Bispinosa Peel. Mater. Sci. Eng. C 2013, 33, 2914–2917. [Google Scholar] [CrossRef]
  39. Freitas, F.S.; Gonçalves, A.S.; De Morais, A.; Benedetti, J.E.; Nogueira, A.F. Graphene-like MoS2 as a Low-Cost Counter Electrode Material for Dye-Sensitized Solar Cells. J. NanoGe J. Energy Sustain. 2012, 1, 11002–11003. [Google Scholar]
  40. Biswas, M.C.; Islam, M.T.; Nandy, P.K.; Hossain, M.M. Graphene quantum dots (GQDs) for bioimaging and drug delivery applications: A review. ACS Mater. Lett. 2021, 3, 889–911. [Google Scholar] [CrossRef]
  41. Khaledian, S.; Nalaini, F.; Mehrbakhsh, M.; Abdoli, M.; Zahabi, S.S. Applications of novel quantum dots derived from layered materials in cancer cell imaging. FlatChem 2021, 27, 100246. [Google Scholar] [CrossRef]
  42. Lei, Y.; Wang, Y.; Du, P.; Wu, Y.; Li, C.; Du, B.; Zou, B. Preparation and photoelectric properties of nitrogen-doped graphene quantum dots modified SnO2 composites. Mater. Sci. Semicond. Process. 2022, 141, 106416. [Google Scholar] [CrossRef]
  43. Manna, A.K.; Gilbert, S.J.; Joshi, S.R.; Komesu, T.; Dowben, P.A.; Varma, S. Tuning photo-response and electronic behavior of graphene quantum dots synthesized via ion irradiation. Phys. B: Condens. Matter 2021, 613, 412978. [Google Scholar] [CrossRef]
  44. Qiang, R.; Sun, W.; Hou, K.; Li, Z.; Zhang, J.; Ding, Y.; Yang, S. Electrochemical Trimming of Graphene Oxide Affords Graphene Quantum Dots for Fe3+ Detection. ACS Appl. Nano Mater. 2021, 4, 5220–5229. [Google Scholar] [CrossRef]
  45. Kalkal, A.; Kadian, S.; Pradhan, R.; Manik, G.; Packirisamy, G. Recent advances in graphene quantum dot-based optical and electrochemical (bio) analytical sensors. Mater. Adv. 2021, 2, 5513–5541. [Google Scholar] [CrossRef]
  46. Zhu, Q.; Mao, H.; Li, J.; Hua, J.; Wang, J.; Yang, R.; Li, Z. A glycine-functionalized graphene quantum dots synthesized by a facile post-modification strategy for a sensitive and selective fluorescence sensor of mercury ions. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2021, 247, 119090. [Google Scholar] [CrossRef] [PubMed]
  47. Veeresh, S.; Ganesh, H.; Nagaraju, Y.S.; Vandana, M.; Ashokkumar, S.P.; Vijeth, H.; Devendrappa, H. UV-irradiated hydrothermal synthesis of reduced graphene quantum dots for electrochemical applications. Diam. Relat. Mater. 2021, 114, 108289. [Google Scholar] [CrossRef]
  48. Ettefaghi, E.; Rashidi, A.; Ghobadian, B.; Najafi, G.; Ghasemy, E.; Khoshtaghaza, M.H.; Mazlan, M. Bio-nano emulsion fuel based on graphene quantum dot nanoparticles for reducing energy consumption and pollutants emission. Energy 2021, 218, 119551. [Google Scholar] [CrossRef]
  49. Dong, Y.; Chen, C.; Zheng, X.; Gao, L.; Cui, Z.; Yang, H.; Li, C.M. One-step and high yield simultaneous preparation of single-and multi-layer graphene quantum dots from CX-72 carbon black. J. Mater. Chem. 2012, 22, 8764–8766. [Google Scholar] [CrossRef]
  50. Peng, J.; Gao, W.; Gupta, B.K.; Liu, Z.; Romero-Aburto, R.; Ge, L.; Ajayan, P.M. Graphene quantum dots derived from carbon fibers. Nano Lett. 2012, 12, 844–849. [Google Scholar] [CrossRef]
  51. Ye, R.; Xiang, C.; Lin, J.; Peng, Z.; Huang, K.; Yan, Z.; Tour, J.M. Coal as an abundant source of graphene quantum dots. Nat. Commun. 2013, 4, 2943. [Google Scholar] [CrossRef] [Green Version]
  52. Jlassi, K.; Mallick, S.; Eribi, A.; Chehimi, M.M.; Ahmad, Z.; Touati, F.; Krupa, I. Facile preparation of NS co-doped graphene quantum dots (GQDs) from graphite waste for efficient humidity sensing. Sens. Actuators B Chem. 2021, 328, 129058. [Google Scholar] [CrossRef]
  53. Kumawat, M.K.; Srivastava, R.; Thakur, M.; Gurung, R.B. Graphene Quantum Dots from Mangifera Indica: Application in near-Infrared Bioimaging and Intracellular Nanothermometry. ACS Sustain. Chem. Eng. 2017, 5, 1382–1391. [Google Scholar] [CrossRef]
  54. Abbas, A.; Tabish, T.A.; Bull, S.J.; Lim, T.M.; Phan, A.N. High yield synthesis of graphene quantum dots from biomass waste as a highly selective probe for Fe3+ sensing. Sci. Rep. 2020, 10, 21262. [Google Scholar] [CrossRef] [PubMed]
  55. Tak, K.; Sharma, R.; Dave, V.; Jain, S.; Sharma, S. Clitoria ternatea Mediated Synthesis of Graphene Quantum Dots for the Treatment of Alzheimer’s Disease. ACS Chem. Neurosci. 2020, 11, 3741–3748. [Google Scholar] [CrossRef] [PubMed]
  56. Thakur, M.; Mewada, A.; Pandey, S.; Bhori, M.; Singh, K.; Sharon, M.; Sharon, M. Milk-derived multi-fluorescent graphene quantum dot-based cancertheranostic system. Mater. Sci. Eng. C 2016, 67, 468–477. [Google Scholar] [CrossRef] [PubMed]
  57. Ghosh, P.; Afre, R.A.; Soga, T.; Jimbo, T. A simple method of producing single-walled carbon nanotubes from a natural precursor: Eucalyptus oil. Mater. Lett. 2007, 61, 3768–3770. [Google Scholar] [CrossRef]
  58. Fadel, H.; Marx, F.; El-Sawy, A.; El-Ghorab, A. Effect of extraction techniques on the chemical composition and antioxidant activity of Eucalyptus camaldulensis var. brevirostris leaf oils. Z. Für Lebensm. Und-Forsch. A 1999, 208, 212–216. [Google Scholar] [CrossRef]
  59. Dubey, M.; Bhadauria, S.; Kushwah, B.S. Green synthesis of nanosilver particles from extract of Eucalyptus hybrida (safeda) leaf. Dig. J. Nanomater. Biostruct. 2009, 4, 537–543. [Google Scholar]
  60. Weng, X.; Lin, Z.; Xiao, X.; Li, C.; Chen, Z. One-step biosynthesis of hybrid reduced graphene oxide/iron-based nanoparticles by eucalyptus extract and its removal of dye. J. Clean. Prod. 2018, 203, 22–29. [Google Scholar] [CrossRef]
  61. Li, C.; Zhuang, Z.; Jin, X.; Chen, Z. A facile and green preparation of reduced graphene oxide using Eucalyptus leaf extract. Appl. Surf. Sci. 2017, 422, 469–474. [Google Scholar] [CrossRef]
  62. Wang, T.; Lin, J.; Chen, Z.; Megharaj, M.; Naidu, R. Green synthesized iron nanoparticles by green tea and eucalyptus leaves extracts used for removal of nitrate in aqueous solution. J. Clean. Prod. 2014, 83, 413–419. [Google Scholar] [CrossRef]
  63. Xie, J.D.; Lai, G.W.; Huq, M.M. Hydrothermal Route to Graphene Quantum Dots: Effects of Precursor and Temperature. Diam. Relat. Mater. 2017, 79, 112–118. [Google Scholar] [CrossRef]
  64. Li, C.; Yue, Y. Fluorescence spectroscopy of graphene quantum dots: Temperature effect at different excitation wavelengths. Nanotechnology 2014, 25, 435703. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Luo, Y.; Xu, Y.; Li, M.; Sun, L.; Hu, G.; Tang, T.; Wen, J.; Li, X. Tuning the Photoluminescence of Graphene Quantum Dots by Fluorination. J. Nanomater. 2017, 2017, 9682846. [Google Scholar] [CrossRef]
  66. Wang, W.; Wang, Z.; Liu, J.; Peng, Y.; Yu, X.; Wang, W.; Zhang, Z.; Sun, L. One-Pot Facile Synthesis of Graphene Quantum Dots from Rice Husks for Fe3+ Sensing. Ind. Eng. Chem. Res. 2018, 57, 9144–9150. [Google Scholar] [CrossRef]
  67. Lin, L.; Zhang, S. Creating High Yield Water Soluble Luminescent Graphene Quantum Dots via Exfoliating and Disintegrating Carbon Nanotubes and Graphite Flakes. Chem. Commun. 2012, 48, 10177–10179. [Google Scholar] [CrossRef] [PubMed]
  68. Li, Y.; Hu, Y.; Zhao, Y.; Shi, G.; Deng, L.; Hou, Y.; Qu, L. An Electrochemical Avenue to Green-Luminescent Graphene Quantum Dots as Potential Electron-Acceptors for Photovoltaics. Adv. Mater. 2011, 23, 776–780. [Google Scholar] [CrossRef]
  69. Routh, P.; Das, S.; Shit, A.; Bairi, P.; Das, P.; Nandi, A.K. Graphene Quantum Dots from a Facile Sono-Fenton Reaction and Its Hybrid with a Polythiophene Graft Copolymer toward Photovoltaic Application. ACS Appl. Mater. Interfaces 2013, 5, 12672–12680. [Google Scholar] [CrossRef] [PubMed]
  70. Zhu, S.; Zhang, J.; Qiao, C.; Tang, S.; Li, Y.; Yuan, W.; Li, B.; Tian, L.; Liu, F.; Hu, R.; et al. Strongly Green-Photoluminescent Graphene Quantum Dots for Bioimaging Applications. Chem. Commun. 2011, 47, 6858–6860. [Google Scholar] [CrossRef] [PubMed]
  71. De, B.; Karak, N. A Green and Facile Approach for the Synthesis of Water Soluble Fluorescent Carbon Dots from Banana Juice. RSC Adv. 2013, 3, 8286–8290. [Google Scholar] [CrossRef]
  72. Liu, Y. Photoluminescence Mechanism and Applications of Graphene Quantum Dots. Ph.D. Thesis, University of Kentucky, Lexington, KY, USA, 2017; p. 118. [Google Scholar]
  73. Fan, T.; Zeng, W.; Tang, W.; Yuan, C.; Tong, S.; Cai, K.; Liu, Y.; Huang, W.; Min, Y.; Epstein, A.J. Controllable size-selective method to prepare graphene quantum dots from graphene oxide. Nanoscale Res. Lett. 2015, 10, 1–8. [Google Scholar] [CrossRef] [Green Version]
  74. Shen, K.; Xue, X.; Wang, X.; Hu, X.; Tian, H.; Zheng, W. One-step synthesis of band-tunable N, S co-doped commercial TiO 2/graphene quantum dots composites with enhanced photocatalytic activity. RSC Adv. 2017, 7, 23319–23327. [Google Scholar] [CrossRef] [Green Version]
  75. Roy, P.; Periasamy, A.P.; Chuang, C.; Liou, Y.R.; Chen, Y.F.; Joly, J.; Liang, C.T.; Chang, H.T. Plant Leaf-Derived Graphene Quantum Dots and Applications for White LEDs. New J. Chem. 2014, 38, 4946–4951. [Google Scholar] [CrossRef]
  76. Wang, R.; Xia, G.; Zhong, W.; Chen, L.; Chen, L.; Wang, Y.; Min, Y.; Li, K. Direct Transformation of Lignin into Fluorescence-Switchable Graphene Quantum Dots and Their Application in Ultrasensitive Profiling of a Physiological Oxidant. Green Chem. 2019, 21, 3343–3352. [Google Scholar] [CrossRef]
  77. Wang, L.; Li, W.; Wu, B.; Li, Z.; Wang, S.; Liu, Y.; Pan, D.; Wu, M. Facile Synthesis of Fluorescent Graphene Quantum Dots from Coffee Grounds for Bioimaging and Sensing. Chem. Eng. J. 2016, 300, 75–82. [Google Scholar] [CrossRef]
  78. Siddique, A.B.; Pratap Singh, V.; Chatterjee, S.; Kumar Pramanik, A.; Ray, M. Facile Synthesis and Versatile Applications of Amorphous Carbon Dot. Mater. Today Proc. 2018, 5, 10077–10083. [Google Scholar] [CrossRef]
  79. Jeong, S.; Pinals, R.L.; Dharmadhikari, B.; Song, H.; Kalluri, A.; Debnath, D.; Wu, Q.; Ham, M.H.; Patra, P.; Landry, M.P. Graphene Quantum Dot Oxidation Governs Noncovalent Biopolymer Adsorption. Sci. Rep. 2020, 10, 7074. [Google Scholar] [CrossRef]
  80. Yang, J.S.; Pai, D.Z.; Chiang, W.H. Microplasma-enhanced synthesis of colloidal graphene quantum dots at ambient conditions. Carbon 2019, 153, 315–319. [Google Scholar] [CrossRef]
  81. Liu, H.; Lv, X.; Li, C.; Qian, Y.; Wang, X.; Hu, L.; Wang, Y.; Lin, W.; Wang, H. Direct Carbonization of Organic Solvents toward Graphene Quantum Dots. Nanoscale 2020, 12, 10956–10963. [Google Scholar] [CrossRef]
  82. Cayuela, A.; Soriano, M.L.; Carrión, M.C.; Valcárcel, M. Functionalized Carbon Dots as Sensors for Gold Nanoparticles in Spiked Samples: Formation of Nanohybrids. Anal. Chim. Acta 2014, 820, 133–138. [Google Scholar] [CrossRef]
  83. Tiede, K.; Tear, S.P.; David, H.; Boxall, A.B.A. Imaging of Engineered Nanoparticles and Their Aggregates under Fully Liquid Conditions in Environmental Matrices. Water Res. 2009, 43, 3335–3343. [Google Scholar] [CrossRef]
  84. Hu, Q.; Gong, X.; Liu, L.; Choi, M.M.F. Characterization and Analytical Separation of Fluorescent Carbon Nanodots. J. Nanomater. 2017, 2017, 30–37. [Google Scholar] [CrossRef] [Green Version]
  85. Shi, W.; Fan, H.; Ai, S.; Zhu, L. Preparation of Fluorescent Graphene Quantum Dots from Humic Acid for Bioimaging Application. New J. Chem. 2015, 39, 7054–7059. [Google Scholar] [CrossRef]
  86. Anas, N.A.A.; Fen, Y.W.; Omar, N.A.S.; Daniyal, W.M.E.M.M.; Ramdzan, N.S.M.; Saleviter, S. Development of Graphene Quantum Dots-Based Optical Sensor for Toxic Metal Ion Detection. Sensors 2019, 19, 3850. [Google Scholar] [CrossRef] [PubMed]
  87. Zong, J.; Yang, X.; Trinchi, A.; Hardin, S.; Cole, I.; Zhu, Y.; Li, C.; Muster, T.; Wei, G. Carbon dots as fluorescent probes for “off–on” detection of Cu2+ and l-cysteine in aqueous solution. Biosens. Bioelectron. 2014, 51, 330–335. [Google Scholar] [CrossRef] [PubMed]
  88. Qu, S.; Chen, H.; Zheng, X.; Cao, J.; Liu, X. Ratiometric fluorescent nanosensor based on water soluble carbon nanodots with multiple sensing capacities. Nanoscale 2013, 5, 5514–5518. [Google Scholar] [CrossRef] [PubMed]
  89. Miao, P.; Tang, Y.; Han, K.; Wang, B. Facile synthesis of carbon nanodots from ethanol and their application in ferric (III) ion assay. J. Mater. Chem. A 2015, 3, 15068–15073. [Google Scholar] [CrossRef]
  90. Zhu, X.; Zhang, Z.; Xue, Z.; Huang, C.; Shan, Y.; Liu, C.; Qin, X.; Yang, W.; Chen, X.; Wang, T. 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. [Google Scholar] [CrossRef] [PubMed]
  91. Shen, J.; Zhu, Y.; Yang, X.; Li, C. Graphene Quantum Dots: Emergent Nanolights for Bioimaging, Sensors, Catalysis and Photovoltaic Devices. Chem. Commun. 2012, 48, 3686–3699. [Google Scholar] [CrossRef]
  92. Zhou, L.; Geng, J.; Liu, B. Graphene Quantum Dots from Polycyclic Aromatic Hydrocarbon for Bioimaging and Sensing of Fe3+ and Hydrogen Peroxide. Part. Part. Syst. Charact. 2013, 30, 1086–1092. [Google Scholar] [CrossRef]
Figure 1. Graphical diagram for the preparation of GQDs from Eucalyptus tree leaves.
Figure 1. Graphical diagram for the preparation of GQDs from Eucalyptus tree leaves.
Nanomaterials 13 00148 g001
Figure 2. (a) PL of GQD at different temperatures (240, 260, 300, and 320 °C), (b) PL of GQD at different heat-treatment durations (360, 400, 440, 480, and 520 min) at a constant temperature of 300 °C (c) PL of GQD prepared at 300 °C after 480 min of treatment.
Figure 2. (a) PL of GQD at different temperatures (240, 260, 300, and 320 °C), (b) PL of GQD at different heat-treatment durations (360, 400, 440, 480, and 520 min) at a constant temperature of 300 °C (c) PL of GQD prepared at 300 °C after 480 min of treatment.
Nanomaterials 13 00148 g002aNanomaterials 13 00148 g002b
Figure 3. UV–Vis of GQD/300.
Figure 3. UV–Vis of GQD/300.
Nanomaterials 13 00148 g003
Figure 4. (a) TEM images of GQD/300 at 10 nm magnification (b) TEM images of GQD/300 at 5 nm magni-fication (c) particle size distribution of GQD/300 (d) XRD spectrum of GQD/300 (e) FTIR spectrum of GQD/300 (f) Raman spectrum of GQD/300.
Figure 4. (a) TEM images of GQD/300 at 10 nm magnification (b) TEM images of GQD/300 at 5 nm magni-fication (c) particle size distribution of GQD/300 (d) XRD spectrum of GQD/300 (e) FTIR spectrum of GQD/300 (f) Raman spectrum of GQD/300.
Nanomaterials 13 00148 g004aNanomaterials 13 00148 g004b
Figure 5. (a) PL spectra of GQD/300 in the presence of various concentrations of Fe3+ ions (b) comparison of the photoluminescence intensities of 55 g mL−1 GQD/300 solution containing various metal ions at an excitation wavelength of 360 nm. (c) Shows the schematic of fluorescence quenching. (d) Fluorescent quenching mechanism of the GQD/300 in presence of the Fe3+ metal ions.
Figure 5. (a) PL spectra of GQD/300 in the presence of various concentrations of Fe3+ ions (b) comparison of the photoluminescence intensities of 55 g mL−1 GQD/300 solution containing various metal ions at an excitation wavelength of 360 nm. (c) Shows the schematic of fluorescence quenching. (d) Fluorescent quenching mechanism of the GQD/300 in presence of the Fe3+ metal ions.
Nanomaterials 13 00148 g005aNanomaterials 13 00148 g005b
Table 1. Product yield of GQDs at different reaction temperatures and constant heat treatment time.
Table 1. Product yield of GQDs at different reaction temperatures and constant heat treatment time.
Temperature (°C)Treatment Time (min)Product Yield (%)
24012010
26012013.24
30012022.33
32012019.25
Table 2. Product yield of GQDs at different treatment times and at a constant temperature of 300 °C.
Table 2. Product yield of GQDs at different treatment times and at a constant temperature of 300 °C.
Temperature (°C)Treatment Time (min)Product Yield (%)
30036022.33
30040027.00
30044035.88
30048044.34
30052039.67
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Saud, A.; Saleem, H.; Munira, N.; Shahab, A.A.; Rahman Siddiqui, H.; Zaidi, S.J. Sustainable Preparation of Graphene Quantum Dots for Metal Ion Sensing Application. Nanomaterials 2023, 13, 148. https://doi.org/10.3390/nano13010148

AMA Style

Saud A, Saleem H, Munira N, Shahab AA, Rahman Siddiqui H, Zaidi SJ. Sustainable Preparation of Graphene Quantum Dots for Metal Ion Sensing Application. Nanomaterials. 2023; 13(1):148. https://doi.org/10.3390/nano13010148

Chicago/Turabian Style

Saud, Asif, Haleema Saleem, Nazmin Munira, Arqam Azad Shahab, Hammadur Rahman Siddiqui, and Syed Javaid Zaidi. 2023. "Sustainable Preparation of Graphene Quantum Dots for Metal Ion Sensing Application" Nanomaterials 13, no. 1: 148. https://doi.org/10.3390/nano13010148

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

Article Metrics

Back to TopTop