Next Article in Journal
Pharmaceutical Aspects of Nanocarriers for Smart Anticancer Therapy
Next Article in Special Issue
Fucoidan-Based Nanoparticles with Inherently Therapeutic Efficacy for Cancer Treatment
Previous Article in Journal
Optimized In Silico Modeling of Drug Absorption after Gastric Bypass: The Case of Metformin
Previous Article in Special Issue
Formulation, Optimization and Evaluation of Luteolin-Loaded Topical Nanoparticulate Delivery System for the Skin Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Natural Carbon Nanodots: Toxicity Assessment and Theranostic Biological Application

1
Genomics Research Center, Academia Sinica, Taipei 115, Taiwan
2
Department of Chemistry, National Taiwan University, Taipei 106, Taiwan
3
Nano Science and Technology Program, Taiwan International Graduate Program, Academia Sinica and National Taiwan University, Taipei 115, Taiwan
4
Department of Biochemistry, College of Medicine, Kaohsiung Medical University, Kaohsiung 807, Taiwan
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Pharmaceutics 2021, 13(11), 1874; https://doi.org/10.3390/pharmaceutics13111874
Submission received: 30 September 2021 / Revised: 27 October 2021 / Accepted: 29 October 2021 / Published: 5 November 2021
(This article belongs to the Special Issue Natural Nanoparticle for Cancer Diagnosis and Treatment)

Abstract

:
This review outlines the methods for preparing carbon dots (CDs) from various natural resources to select the process to produce CDs with the best biological application efficacy. The oxidative activity of CDs mainly involves photo-induced cell damage and the destruction of biofilm matrices through the production of reactive oxygen species (ROS), thereby causing cell auto-apoptosis. Recent research has found that CDs derived from organic carbon sources can treat cancer cells as effectively as conventional drugs without causing damage to normal cells. CDs obtained by heating a natural carbon source inherit properties similar to the carbon source from which they are derived. Importantly, these characteristics can be exploited to perform non-invasive targeted therapy on human cancers, avoiding the harm caused to the human body by conventional treatments. CDs are attractive for large-scale clinical applications. Water, herbs, plants, and probiotics are ideal carbon-containing sources that can be used to synthesize therapeutic and diagnostic CDs that have become the focus of attention due to their excellent light stability, fluorescence, good biocompatibility, and low toxicity. They can be applied as biosensors, bioimaging, diagnosis, and treatment applications. These advantages make CDs attractive for large-scale clinical application, providing new technologies and methods for disease occurrence, diagnosis, and treatment research.

1. Introduction

The materials studied in the past were at the micrometer scale; however, in the past few decades, nanometer-scale development has changed, which has become an essential direction of scientific and technological development. Nano-sized materials are widely used in various fields, such as medicine, biosensor development, energy research, and catalysis. As fluorescent nanomaterials can emit light, they have the potential for application in biomarking technology, such as semiconductor-based quantum dots containing cadmium sulfide (CdS) or cadmium selenide (CdSe). However, most materials with a high quantum yield (QY, ϕ) on the market include heavy metal components, which pose risks related to their biological toxicity and cytotoxicity. Contamination may occur during the synthesis process, and they are not adequately recycled after use; as such, the associated harm to the environment should not be underestimated. The above shortcomings limit the application range as semiconductor quantum dots.
In recent years, nanotechnology has been widely used in the field of biomedicine [1]. Generally, when the scale of a substance drops to the nanometer level, its physical and chemical properties also undergo tremendous changes, especially in terms of its optical properties, providing an ideal development space for applications in the field of biomarkers and optical imaging of diseases [2]. Since their inception, carbon nanomaterials have attracted attention from researchers in materials science and biomedicine, especially considering their excellent optical properties, facilitating their application in bioimaging, biomarkers, and sensors [3]. Fluorescent carbon dots (CDs) are another new type of carbon nanomaterial, with a typical particle size of about 10 nm. In addition to the advantages related to the particle size and wavelength-dependent luminescence characterizing quantum dots, fluorescent CDs also have high light stability and present no light scintillation phenomena. Their surface is easily functionalized and modified, and the preparation materials are widely available [4,5,6]. Wang et al. evaluated the toxicity of CDs and showed that they do not cause any abnormalities or damage to tissues and organs [7]. Kang et al. compared CDs, single-walled carbon tubes, and carbon dioxide [8]. After the cytotoxicities of silicon and zinc oxide were determined, CDs showed the lowest toxicity compared with the materials mentioned above. CDs’ unique optical properties and excellent biological properties confer outstanding advantages and good development prospects in biomedical optical imaging and tumor diagnosis. In addition, the ability to introduce multiple functional groups onto the surface of CDs also provides possibilities for their modification.
Green manufacturing is a modern manufacturing model that comprehensively considers the environmental impact and resource benefits. Due to the increasing awareness of environmental protection, green manufacturing is becoming an increasingly important process in all walks of life. The chemical synthesis of CDs can be similar to that produced from various natural resources in the nanoscale process. For instance, Hsu et al. used coffee grounds to synthesize CDs by calcining in 2012. The average particle size of the synthesized CDs was about 5 nm, and they were successfully applied to biological imaging. Unlike the previous need for complex processing procedures, a rapid and straightforward synthesis method was developed. Later, Crista et al. used different organic compounds to synthesize CDs, with an average particle size of about 2.6–7.9 nm, with many carboxyl and amine groups, proving that the quantum yield varies with functional groups [9]. In 2013, Qu et al. used citric acid and ethylenediamine to synthesize CDs, leading to a quantum yield as high as 60.2% [10]. It can be speculated that naturally synthesized carbon nanomaterials should have higher biocompatibility and lower toxicity. To date, many techniques for synthesizing CDs from natural sources have been developed, including laser ablation, arc discharge, electrochemistry, thermal decomposition, ultrasonic, and microwave methods, among others. In particular, the thermal decomposition method is simple, safe, fast, and effective. Research has found that carbon nanoparticles synthesized by this method have good fluorescent performance, although the quantum yield is not high [11,12,13]. The optical properties, stability, and biocompatibility are still powerful enough for use in bioapplications. This review introduces and simplifies the current carbon nanocomposite synthesis techniques. We provide a brief introduction to fluorescent CDs, mainly reviewing their structural characteristics, carbon source materials, preparation methods, luminescence mechanism, and applications in the field of biomedicine. Obtaining CDs from natural resources can produce nanomaterials that are more environmentally friendly. The considered technique intends to use available (in a daily sense) materials to synthesize low-toxicity fluorescent nanomaterials. The production process does not require the use of many precursor chemicals to build natural CDs. We hope that developing a safe and straightforward production process can unite fluorescent materials and natural CDs and contribute to understanding and exploring the application of nano-fluorescent materials.

2. Natural Carbon Nanodots

CDs are a new type of carbon nanomaterials with luminescence characteristics, quasi-zero dimension, relatively simple preparation, abundant sources of raw materials, easy surface functionalization, low toxicity, and good biocompatibility. The fluorescence wavelength can be adjusted, and the two-photon absorption area is large. In various literature reports, CDs have also been referred to as carbon quantum dots (CQDs), carbon nanodots (CNDs), Graphene quantum dots (GQDs), carbon nanocrystals, and so on. Due to their excellent performance, CDs show promising potential application prospects in sensing, biology, medicine, food, environment, catalysis, photoelectricity, energy, and so on. Researchers have conducted extensive scientific studies and made significant progress [14]. CDs synthesis methods can be mainly divided into two categories: One uses physical or chemical means to crack larger carbon structures (e.g., carbon nanotubes, graphite rods, graphene, carbon powder) into tiny CDs; these top-down methods include arc discharge methods, laser ablation methods, chemical oxidation methods, and so on. Small organic molecule precursors, such as sugar, citric acid, and amino acids, are used as carbon sources for these methods through functional group coupling to achieve polymerization to prepare the CDs [15,16,17]. The second category comprises bottom-up methods, such as electrochemical, hydrothermal, pyrolysis, microwave-assisted, and ultrasonic methods [18].
Other methods, such as template methods, neutralization reaction exothermic methods, and micro-fluidized bed technology methods, can also be used to prepare carbon dots [19,20,21]. The preparation of CDs well-embodies the concept of green chemistry, using cheap, environmentally friendly carbon source precursors and natural renewable, cheap raw materials as carbon sources. The resources for synthesizing CDs can be found in the natural environment, such as eggs, grass, tea leaves, leaves, silk, silkworm pupae, shrimp shells, grapefruit peel, peanut shells, coffee grounds, beer, and other materials. CDs were discovered and debuted for the first time due to their fluorescent luminescence characteristics. The luminescence mechanism of CDs has always been a critical research direction for researchers, considering factors such as the quantum size effect, surface state, functional group mechanisms, electron holes, and radiation. Their rearrangement theory has been studied in various aspects. Although a complete theoretical explanation system of the CDs fluorescence mechanism has not been formed, the absorption and fluorescence of CDs exhibit properties such as photoluminescence, chemiluminescence, electrochemiluminescence, and luminescence. The conversion of photoluminescence, peroxidase-like activity, non-toxicity, and other physical and chemical properties provides a solid and feasible theoretical basis for further research.
The main application research directions of CDs can be divided into the following categories (Figure 1):
  • Imaging: multicolor fluorescent images of mammalian cells, plant cells, and micro-organisms, and imaging in mice;
  • Photocatalysis: the degradation of organic molecules, the reduction in CO2, and water splitting;
  • Optoelectronic devices: LEDs and solar cells (sensitizer/co-sensitizer, transport layer, electrolyte, and/or co-catalyst for counter electrode);
  • Sensors: food quality and safety, drug analysis, environmental pollution determination, immunoassay, and other fields, such as detecting heavy metal ions, anions, pesticides, molecules, small organic molecules, and/or nucleic acids;
  • Electrocatalysis: mainly used in oxidation-reduction reactions, oxygen evolution reactions, hydrogen evolution reactions, and reduction reactions for carbon dioxide, and dual-function catalysts;
  • Biomedicine: photodynamic therapy, photosensitizers for cancer cell destruction, radiotherapy, the tracing and delivery of drugs or genes, drug release, and anticancer drugs.
Most carbon sources derive from exhausted petroleum resources or non-environmentally friendly manufacturing processes. Therefore, using recycled materials and bio-renewable resources to develop high-performance CDs is a critical green environmental issue. In recent years, the awareness of environmental protection has significantly risen, and, as such, naturally derived CDs have gradually received more attention.
The main factor is that they are derived from renewable and sustainable biological materials; for example, lignocellulose from dead wood, waste wood, rice straw, bagasse, wheat straw, etc. Suppose that such a raw material can be used as a carbon resource and converted into CDs with therapeutic and diagnostic value. In that case, we can imagine the result will not overwhelm the demand of the food supply chain. Still, it can also address waste disposal problems while applying the resource to produce high value-added products such as electronics, energy, and biomedicine. Table 1 summarizes the existing research on converting various natural resources into CDs, and the associated applications and emission spectra are described in detail.

3. Toxicity Evaluation of Natural Carbon Nanodots

CDs can be derived from a wide range of synthetic raw materials, low-cost, stable chemical properties, and non-toxic materials. The application of CDs in medicine and pharmacy was recently extensively studied. One of the most promising nanomaterials is carbon quantum dots (CQDs). For this purpose, the toxicity of CQDs was investigated in cells and living systems (Table 2). In 2019, L. Janus used human dermal fibroblasts to conduct a cytotoxicity study of N-doped chitosan-based CDs [129]. As shown in Figure 2, after 48 h, the cell viability was recorded as 94%. CDs were synthesized by utilizing non-toxic raw materials and removing unreactive residues during purification.
First, we start with the selection of carbon source and the adjustment of dopants that synthesize carbon dots, and at the same time list the test models of carbon dots. Secondly, according to the selective response of the carbon dot to the biological models, the corresponding toxicity was constructed, and the actual application was investigated. As mentioned in Table 2, CQDs are usually doped with nitrogen to enhance their fluorescence quantum yield and optical performance [130,131]. CDs doped with nitrogen (such as carbon nanotubes, graphene, hollow spheres, etc.) have unique properties. They can inject electrons into the carbon substrate to change the electron and transport characteristics such as sensors, nanogenerators, etc., and are widely used and have become a research trend. Moreover, green methods that utilize natural biomass/biowaste and micro-organisms without introducing toxic chemicals as CDs precursors have been widely used to synthesize CQDs [132].
Table 2. Toxicity evaluation methods for CDs.
Table 2. Toxicity evaluation methods for CDs.
MaterialSourcesConcentrationQY (%)Cells or Animal ModelsToxicityRef.
Carbon quantum dotsMedicinal mulberry leaves500 ug/mL9.7Human normal hepatic stellate cell line LX-2 cells and human HCC cell line HepG2 cellsAlmost non-cytotoxic[133]
Carbon dotsMango peel500 ug/L8.5A549 cellsRemained above 90%
Low toxicity
[134]
Nitrogen-doped carbon quantum dotsWatermelon juice300 ug/mL10.6HepG2 cellsRemained 90%
Low cytotoxicity
[135]
N-doped carbon quantum dotsBio-waste lignin100 mg/mL8.1Mouse macrophage cellsRemained 96.8%
Low toxicity
[136]
Carbon dotsRoast duck1 mg/mL38.05PC12 cells and C. elegansRemained 91.19%
Low toxicity
[122]
Nitrogen-doped carbon dotsP. acidus fruit juice200 ug/mL12.5Cells and C. elegansRemained 93%
Low cytotoxicity
[137]
Carbon quantum dotsSalvia
hispanica L. seeds
250 ug/mL17.8HEK293 cell lineRemained 91.7%
Low toxicity
[138]
Carbon dotsWheat straw0.8 mg/mL7.5HeLa cellsNegligible
cytotoxicity
[139]
Carbon dotsMalus floribunda fruit200 ug/mL19Cells and C. elegansRemained 93% and low toxicity[140]
Carbon quantum dotsBanana peel waste200 ug/mL20C. elegansLow toxicity[141]
Nitrogen and sulfur dual-doped carbon quantum dotsFungus fibers400 ug/mL28.11HepG2 cellsRemained over 95%
Low cytotoxicity
[142]
Carbon dotsSweet lemon peel500 ug/mLn/aMDA-MB231 cellsRemained above 75%
Low cytotoxicity
[143]
Carbon dotsLychee waste1.2 mg/mL23.5Skin melanoma cellsRemained above 89%
Low cytotoxicity
[144]
Nitrogen-doped carbon quantum dotsCitrus lemon2 mg/mL31Human breast adenocarcinoma cellsRemained above 88%
low cytotoxicity,
[145]
Carbon dotsDaucus carota subsp. sativus roots1 mg/mL7.6MCF-7 cellsRemained above 95%
Low toxicity
[146]
Carbon nanodotsCustard apple peel waste biomass100 ug/mLn/aHeLa and L929 cellsRemained above 85%
Low toxicity
[147]
Carbon quantum dotsPineapple peel1 mg/mL42HeLa and MCF-7 cellsRemained 84%
Low toxicity
[148]
N-carbon dotsJackfruit seeds2 mg/mL17.91A549 cellsRemained 96% and less toxic[149]
S. Cong et al. [122] used PC12 cells for a cytotoxicity study of CQDs obtained from a roast duck. After 36 h, the cell viability of PC12 was recorded as 91.19% at the concentration of 1 mg/mL. In addition, using CDs at a concentration of 15 mg/mL for treating nematodes did not lead to any death for 24 h. These results indicate the low toxicity of CDs, even after a long period of exposure at high concentrations. In 2021, R. Atchudan et al. [141] used C. elegans as an animal model for their toxicity evaluation of CQDs. As displayed in Figure 3, the CQDs synthesized from banana peel were shown to have low nematode toxicity, even at a high concentration of 200 ug/mL. These results can be explained by their utilizing non-toxic raw materials without adding any passivates or additives.

4. Theranostic Application of Natural Carbon Nanodots

Therefore, fluorescent materials are expected to be critical in biological applications. CDs have become the focus of attention as new nanomaterials due to their excellent light stability, fluorescence, good biocompatibility, and low toxicity. Surgery, radiotherapy, and chemotherapy are inevitable treatments for most cancer patients, but these processes have considerable side effects on the human body. However, fluorescent nanomaterials have the advantages of high fluorescence stability, low biotoxicity, and good biocompatibility. The most important thing is that they can be used to perform non-invasive targeted therapy on human lesions, exploiting their characteristics to avoid harm to the human body caused by the abovementioned treatments.
This paragraph covers the luminescence mechanism of CDs and their applications in biology, focusing on applying natural CDs in biological diagnosis and treatment. We discuss the combination of CDs with specific targeting molecules to form CD-based probes for detecting fluorescent signals. With the help of advanced optical imaging technology, real-time dynamic monitoring of molecules in cells and organisms can be carried out, and rapid immunofluorescence analysis of primary infectious disease sources can be carried out. They can provide new technologies and methods for disease occurrence, diagnosis, and treatment research [150].

4.1. Bioimaging

The discrete and diverse microstates of CDs lead to broad excitation and emission ranges [151,152]. CDs have many unique properties, and their excellent photostability can provide fluorescent information in a biological environment. The surface modification of functional groups can make CDs more helpful in applying biomarkers (Table 3). In this section, we compare bioimaging from cells to living animals, emphasizing the biological diversity to prove the generalized safety of CDs. In cell imaging applications, CDs are usually applied to HepG2 cells, HeLa cells, T24 cells, and so on [27,102,120,124]. In addition, normal cell lines have been used in trials, showing good cell compatibility [24,71]. Alam et al. treated HaCaT cells with cabbage-derived CDs and showed that, at 500 μg/mL of CDs, the cell viability was 100%. Furthermore, CDs’ tunable excitation and emission show promise for normal cell imaging under the irradiation of confocal fluorescence microscopy [71]. In a cell imaging experiment, Mehta et al. considered CDs originating from apple juice and three fungal (M. tuberculosis, P. aeruginosa, and M. oryzae) sources. Cell compatibility was shown by feeding more than 100 mg/mL of CDs. The germination of M. oryzae spores also strongly indicated how CDs are good biocompatible nanocomposites at high concentrations (i.e., 400 mg/mL). The fungi appeared red, green, and blue in confocal laser microscopic images [62].
Kasibabu et al. provided pictures of Bacillus subtilis and Aspergillus aculeatus after incubating with CDs derived from papaya juice for 1–6 h. The uptake ability was shown by observing well-dispersed CDs in the cytoplasm of the fungi [90]. In vivo tests are critical standards for investigating the potential and toxicity of CDs in animals. Atchudan et al. explored colorful nitrogen-doped CDs (NCDs) derived from gooseberry by hydrothermal methods in C. elegans imaging. C. elegans presented blue, green, and red in the whole body when excited under 400 nm, 470 nm, and 550 nm. The cell viability was over 97% after incubating C. elegans in NCDs for 24 h from 0 to 200 μg/mL (Figure 4a–f) [96]. Cong et al., using roast duck as the source and pyrolyzing at 200 °C, 250 °C, and 300 °C, synthesized single-fluorescence CDs. C. elegans were treated with 15 mg/mL of the CDs pyrolyzed at 300 °C (300-CDs) for 24 h. Compared with the wild-type group, the accumulation and uptake of 300-CDs made the intestine appear blue under UV light exposure (Figure 4g–j) [122]. The murine model has also been widely used for determining the efficiency of CDs.
Ding et al. subcutaneously injected 100 μL of red-emitting CDs (R-CDs) into nude mice. Strong fluorescence at 700 nm was detected under an excitation wavelength of 535 nm, indicating the excellent penetration ability from tissues to skin. Furthermore, the mice were still alive after 10 days (Figure 5a) [22]. Liu et al. investigated how the accumulation of carbonized polymer dots (CPDs) varied with time. Most CPDs remained in the lung and liver in the early period, with negligible dispersion to the brain and heart. The metabolism of CPDs was confirmed after 4 h, as their fluorescence decreased sharply (Figure 5b) [23]. Ding et al. compared subcutaneous and intravenous injections of near-infrared emissive CDs (NIR-CDs). Through the use of subcutaneous methods, NIR-CDs were distributed into mouse skin and tissues. As for intravenously injected nude mice, fluorescence was seen in the bladder, indicating NIR-CDs elimination via urine (Figure 5c) [26]. Liu et al. tracked CDs (Fn-CDs) in Kunming mice for different periods.
They concluded that, with the assistance of PL, Fn-CDs show strong fluorescence in the bladder and are eliminated after 7 h (Figure 5d) [28]. Therefore, CDs can perform well as imaging nanoparticles and be adapted to different cell lines and living animals. In addition, their multiple emission ranges, low toxicity, and small size confer their high potential in future clinical applications.

4.2. Sensors

Inorganic ions are critical for creatures, not only for enhancing the efficiency of catalytic reactions in bio-systems but also for maintaining our fundamental life functions. However, it is harmful to have too many metal cations, which are highly toxic to human beings. Therefore, developing sensors for inorganic ions is a simple beneficial method to collate the concentration and standard. Owing to the multiple PL of CDs, we can observe the intensity variation and chelation quenching effect at other peaks in the PL image. For natural CDs sensors, Fe3+, Hg2+, and Cu2+ are the most examined targets. Shen et al. assessed HepG2 and HeLa cell images after incubating with CDs and Fe3+.
According to the decreased blue fluorescence under 405 nm irradiated light, Fe3+ had a practical quenching effect on the CDs (Figure 6a–d) [33]. Hu et al. prepared double--emitted biomass nitrogen co-doped CDs (B-NCdots) for Cu2+ probing in T24 cells. Similarly, the quenching effect was still available for Cu2+, causing a decreased intensity of blue and green, as shown in Figure 6e,f [117]. Furthermore, bio-related molecules, including peptides and drug-containing cells, are even more crucial. Some researchers designed chemically sensitive CDs to assist in directly resolving the effects of molecules by monitoring the decrease or increase in fluorescence. Liang et al. added 0.5 mM and 1 mM of glutathione and 150 μg/mL rose-red fluorescence CDs (wCDs) in L929 cells, HeLa cells, and HepG2 cells. An intense quenching effect of glutathione was observed in L929 cells and HeLa cells. However, no apparent variation occurred in HepG2 cells, which implies a distinct response of wCDs to glutathione (GSH) in various cells (Figure 7) [24].
Wang et al. constructed a glutathione assay composed of eggshell-derived CDs and Cu2+. The authors quantify without other indicators by plotting the fluorescence ratio versus the glutathione concentration (Figure 8a) [107]. Wang et al. examined how the fluorescence intensity of Shiitake mushroom-derived CDs (MCDs) varied with pH in dexamethasone-induced HeLa cell apoptosis. At higher concentrations of dexamethasone, the fluorescence under excitation at 405 nm and 488 nm was stronger, indicating a positive relationship on MCDs with increasing intracellular acidification (Figure 8c–e) [66]. As for drug probing, Zhu et al. analyzed doxorubicin (DOX), an anthracycline-based anticancer medicine, by taking advantage of the PL of plum-based carbon quantum dots (PCQDs). The dual-emitted property at the wavelengths of 491 nm and 591 nm provided a ratiometric calibration curve as a function of the DOX concentration. They also confirmed the accuracy by analyzing urine and serum samples (Figure 8b) [27]. CDs are suitable for detecting different kinds of molecules and ions. The intensity changed at a single emission peak, but the amplitude ratio of two emission peaks is valid for sensing experiments. Due to their intrinsic properties, CDs show promise in the bio-sensing field and are applied to cancer therapy.

4.3. Antibacterial Activity

Bacteria are well-known as the origins of various diseases. Recently, super bacteria have appeared globally, which cause incurable illnesses due to the abuse of antibiotics. In addition, people have come to pay more attention to the side effects of antibiotics and wish to avoid unexpected risks. Nanomedicines, especially CDs, have been taken into consideration as substitute methods. E. coli and S. aureus are the most common types of bacteria for investigating how nanomedicines or antibiotics work to induce apoptosis in bacteria. Wang et al. used CDs (ACDs) derived from Artemisia argyi leaves to treat E. coli and S. aureus cultures. According to the SEM images (Figure 9a–h), it can be seen that the cell walls of E. coli were destroyed; however, there was no distinct difference between treated and untreated S. aureus. This means that ACDs are selective to Gram-negative bacteria due to the structural properties of their cell walls [123]. Sun et al. synthesized chlorhexidine gluconate CDs from large to small (l-CGCDs, m-CGCD, and s-CGCDs) to determine the relationship between size and antibacterial activity. From the SEM imagery (Figure 9i), it can be seen that the rigidity of cell walls was the strongest in the control group and decreased from l- to s-CGCDs groups. These results revealed that CGCDs lead to frustration in the walls and membranes of E. coli and S. aureus. Bacterial death can be controlled by tuning the size of the CGCDs [76]. Ma et al. tested three kinds of CDs, including osmanthus leaves-derived CDs (OCDs), milk vetch-derived CDs (MCDs), and tea leaves-derived CDs (TCDs). In Figure 10, 80% of E. coli and S. aureus were killed by OCDs at a 1 mg/mL concentration, while 70% of bacteria survived in the MCDs group. In addition, E. coli had stronger resistivity than S. aureus among these CDs. CDs are internalized into bacteria. The outer surface of bacteria is attached to CDs leading to indirect proliferating inhibition [153,154,155]. These results prove the natural sources are essential for the synthesis of CDs [79]. As mentioned above, the tunability of raw material and diameters primarily affect the antibacterial efficiency and selectivity of the resultant CDs. Treating the patient’s wounds after surgery with CDs with editable properties can tremendously decrease the associated risks.

4.4. Anticancer Activity

At present, cancers are prevalent within all age ranges. Cancers may be fatal due to unexpected syndromes as well as the disorder of living functions. Furthermore, conventional cancer therapies are long-term processes. Surgeries are straightforward methods, but recovery typically poses a challenge for patients. Even though chemotherapy seems safer, the currently used drugs lack selectivity and affinities to specific tumors. Some targeted therapies have been developed in recent years. However, they are expensive and only valid for certain types of cancers. CDs can provide great theranostic nanomedicines in cancer treatments. Scientists have attempted to eliminate cancer cells through photothermal therapy (PTT) [81,82] and photodynamic therapy (PDT) [83,87,129] to fulfill tumor targeting. Li et al. tested NIR-II emitted (900–1200 nm) CDs (CDs), adapted for 808 nm laser photothermal therapy. According to the in vivo test (Figure 11), the temperature increased to 50 °C in the intratumoral environment after intravenous injection. Tumor inhibition and volume shrinkage were observed within 6 days, compared with the PBS group. No detectable damage to tissues or weight loss after the treatment confirmed the high biocompatibility of the CDs [156]. Jia et al. prepared red-light absorbing (610 nm) CDs from Hypocrella Bambusa (HBCUs). They found that HBCDs highly generate 1O2 under 635 laser irradiation. The reactive radicals induced apoptosis of cancer cells, which is helpful in the hypoxia tumor environment.
As shown in an in vivo experiment (Figure 12), due to the synergistic effect of PDT and PTT, the temperature at the tumor site increased to 56.4 °C in 10 min. Secondly, a good tumor inhibition effect was found after 14 days of therapy, even though the tumor could not be depleted thoroughly. No harmful phenomena were observed in other organs, indicating the safety of the treatment [82]. Xue et al. conducted modification with polyethylene glycol diamine (H2N-PEG-NH2), chlorin e6 (Ce6), and transferrin (Tf) on natural biomass CDs (NBCDs) to increase the targeting efficacy.
The resulting products, NBCD-PEG-Ce6-Tf, were shown to remain within the tumor environment for 120 h using a real-time NIR fluorescence image (Figure 13a). The mice were irradiated daily under a 650 nm laser to generate 1O2. During the 21-day process, tumor growth was stopped, and the tumors were ablated, indicating no conflict between the NBCDs and modifications (Figure 13b,c) [126]. Li et al. synthesized reactive oxygen species (ROS)-generating CDs from ginger. The CDs were harvested from HepG2 tumor inoculating mice; next, the tumor regressions were observed in the C-dot (440 μg) treated group; the tumor growth was prominently delayed, which attained only 3.7 ± 0.2 mg. In contrast, tumors in the PBS group grew up to 104 mg [84]. Boobalan et al. added 30 μg/mL of CDs into P. aeruginosa. They observed the destruction of cell walls due to ROS attack, in agreement with the results of ROS fluorescence detection using a fluorogenic dye, 2′,7′-dichlorofluorescein diacetate (DCFDA) (Figure 14a–c). MDA-MB-231 breast cancer cells were treated with CDs (3.34 μg/mL). Cell apoptosis staining, acridine orange and ethidium bromide (AO/EtBr), and nuclear staining (Hoechst 33342) were applied to the cells. The presence of orange colors and blue dots indicate cell fragmentation due to CDs (Figure 14d–g) [80]. CDs have a powerful potential in the anticancer field. Their flexibility is because the CDs are modified with various molecules, which can improve the uptake by tumor cells and increase the tumor-killing ability of the nanohybrids.

5. Discussion and Conclusions

Carbon nanomaterials have been widely used in various scientific, engineering, and commercial fields, due to their high catalytic activity and good stability. Among them, the new “zero-dimensional” carbon nanomaterials, CDs, have unique optical properties, such as stable fluorescence signals, no light scintillation, adjustable excitation and emission wavelength, and low biological toxicity and biocompatibility. These advantages gradually led to the popularity of researching carbon nanomaterials, widely used in bioimaging, natural cell labeling, sensors, photocatalysis, solar cells, and light-emitting elements. This article mainly reviewed the different synthesis methods of CDs (including top-down and bottom-up methods) and their applications. Their luminescence properties can be adjusted through surface modification. They have been applied in many fields and have great potential. The function of CDs can also be modified by using various surface functional groups, allowing them to act, for example, as detectors and cleaners for different heavy metal ions or by doping with other ions. By controlling the surface light energy groups, they can be better used in the required fields.
Green chemistry is a discipline that has gradually received attention in recent years. The core concept focuses on the development of environmentally friendly chemical technologies. At the technical level, chemical technologies and methods are applied to reduce or eliminate the use and generation of hazardous substances in chemical synthesis and analysis, and recovery and reuse technologies are combined with increasing energy and material use efficiency. Green chemistry and nanotechnology have become emerging technology research and development directions in recent years. With the deepening of the concept of sustainability, combining the advantages of the two and accelerating the expansion of their research and development applications has become a top priority. Laboratories are committed to determining the complementary relationships between green chemistry practices and nanotechnology and applying them to materials development, chemical analysis, energy, environmental, and other related fields. The preparation of CDs well embodies the concept of green chemistry. Cheap, environmentally friendly carbon source precursors and natural renewable raw materials as carbon sources for preparation, such as eggs, grass, leaves, silk, coffee grounds, beer, and other materials, have become carbon sources for the synthesis of CDs.
Excellent performance and a unique structure provide natural CDs unlimited charm and various changes. Natural CDs, combined with biological and pharmaceutical molecules of interest through surface modification, seem to be an emerging platform for imaging probes that are both diagnostic and therapeutic. The next generation of nano-molecular probes integrates a variety of fluorescent dyes, drugs, and multifunctional nanomaterials into a single nanoprobe, providing superior signal contrast, controllable transmission, and targeted drug delivery capabilities. However, before this kind of multifunctional imaging probe was used in diagnosis and treatment, there were still many challenges, such as long-term safety, risk-benefit, biocompatibility, and biodistribution, to be evaluated. In the future, we need to solve several critical scientific problems in the research of natural CDs. First of all, the uncertain chemical groups on the surface indicate that natural-synthetic CDs are a kind of unsure material. It means the method of natural mass production of high-quality CDs is still a big challenge. Secondly, due to the different sources of naturally transformed CDs, the luminescence centers of CDs are also dissimilar. Finding a suitable luminescence position is also essential to research content. Third, categorizing different natural CDs and conducting a systematic comparative analysis will be beneficial research methods. Nanotoxicology is the emerging study of potential adverse effects derived from the interaction between nanomaterials and biological systems, and it is bound to become more critical. To further clarify its physical toxicity and adjust the size and structure accordingly, it will significantly improve CDs’ application performance while expanding a more comprehensive range of applications. The scale and complexity of biomedical issues have always been an enormous challenge for researchers. It is more necessary to conduct cross-disciplinary research in chemistry, physics, materials, biomedical engineering, toxicology, public health, and clinical medicine. As more research on natural CDs continues to develop, the topic will achieve breakthroughs and progress in a short period.

Author Contributions

Conceptualization, M.-H.C., B.-G.C., R.-S.L. and M.H.; writing—original draft preparation, M.-H.C., B.-G.C. and L.T.N.; writing—review and editing, M.-H.C., B.-G.C., L.T.N., W.-T.H. and C.-H.L.; supervision, R.-S.L. and M.H.; project administration, R.-S.L. and M.H. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by Genomics Research Center, Academia Sinica, Taiwan, to Michael Hsiao. This work was also financially supported by the Ministry of Science and Technology in Taiwan (MOST 109-2113-M-002-020-MY3) to Ru-Shi Liu. Academia Sinica Outstanding Postdoctoral Fellowship supported to Ming-Hsien Chan.

Acknowledgments

The authors would like to express their gratitude to support from Genomics Research Center, Academia Sinica, Taiwan.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AO/EtBrAcridine orange and ethidium bromide
B-NCdotsBiomass nitrogen co-doped carbon dots
c-CDsCamellia bee pollen carbon dots
CDsCarbon dots
CNDsCarbon Nanodots
CPDsCarbonized polymer dots
CQDsCarbon quantum dots
CdSeCadmium selenide
CdSCadmium sulfide
ACDsCDs derived from Artemisia argyi leaves
Ce6Chlorin e6
l-CGCDsLarge chlorhexidine gluconate carbon dots
m-CGCDsMedium chlorhexidine gluconate carbon dots
s-CGCDsSmall chlorhexidine gluconate carbon dots
DCFDA2′,7′-dichlorofluorescein diacetate
DOXDoxorubicin
DCDsDual emission carbon dots
EtOHEthanol
Fn-CDsF. nucleatum-carbon dots
HBCUsHypocrella Bambusa
HBCDsHypocrella Bambusa CDs
GSHGlutathione
GQDsGraphene quantum dots
l-CDsLotus bee pollen carbon dots
MCDsMilk vetch-derived CDs
NCDsNitrogen-doped CDs
NBCDsNatural biomass CDs
NIR-CDsNear-infrared emissive CDs
PDTPhotodynamic therapy
PLPhotoluminescence
PTTPhotothermal therapy
PCQDsPlum-based carbon quantum dots
H2N-PEG-NH2Polyethylene glycol diamine
OCDsOsmanthus leaves-derived CDs
QY, ϕQuantum yield
ROSReactive oxygen species
R-CDsRed-emitting CDs
wCDsRose-red fluorescence CDs
SCDsSingle-emission carbon dots
MCDsShiitake mushroom derived CDs
TCDsTea leaves-derived CDs
TfTransferrin

References

  1. Wu, D.; Zhou, J.J.; Creyer, M.N.; Yim, W.; Chen, Z.; Messersmith, P.B.; Jokerst, J.V. Phenolic-enabled nanotechnology: Versatile particle engineering for biomedicine. Chem. Soc. Rev. 2021, 50, 4432–4483. [Google Scholar] [CrossRef]
  2. Mauro, N.; Utzeri, M.A.; Drago, S.E.; Buscarino, G.; Cavallaro, G.; Giammona, G. Carbon Nanodots as Functional Excipient to Develop Highly Stable and Smart PLGA Nanoparticles Useful in Cancer Theranostics. Pharmaceutics 2020, 12, 1012. [Google Scholar] [CrossRef] [PubMed]
  3. Nair, A.; Haponiuk, J.T.; Thomas, S.; Gopi, S. Natural carbon-based quantum dots and their applications in drug delivery: A review. Biomed. Pharmacother. 2020, 132. [Google Scholar] [CrossRef]
  4. Zhao, X.Q.; Wang, L.; Ren, S.M.; Hu, Z.; Wang, Y.M. One-pot synthesis of Forsythia@carbon quantum dots with natural anti-wood rot fungus activity. Mater. Des. 2021, 206. [Google Scholar] [CrossRef]
  5. Song, J.Q.; Zhao, N.; Qu, Y.; Zhao, L.S. Natural deep eutectic solvent-assisted preparation of nitrogen-doped carbon dots for ratiometric determination of pirimicarb and pH. Dyes Pigment. 2021, 193. [Google Scholar] [CrossRef]
  6. Gagic, M.; Kociova, S.; Smerkova, K.; Michalkova, H.; Setka, M.; Svec, P.; Pribyl, J.; Masilko, J.; Balkova, R.; Heger, Z.; et al. One-pot synthesis of natural amine-modified biocompatible carbon quantum dots with antibacterial activity. J. Colloid Interf. Sci. 2020, 580, 30–48. [Google Scholar] [CrossRef]
  7. Wang, C.X.; Pan, C.W.; Wei, X.R.; Yang, F.; Wu, W.J.; Mao, L.Q. Emissive carbon dots derived from natural liquid fuels and its biological sensing for copper ions. Talanta 2020, 208. [Google Scholar] [CrossRef]
  8. Kang, L.X.; Hu, Y.; Liu, L.L.; Wu, J.X.; Zhang, S.C.; Zhao, Q.C.; Ding, F.; Li, Q.W.; Zhang, J. Growth of Close-Packed Semiconducting Single-Walled Carbon Nanotube Arrays Using Oxygen-Deficient TiO2 Nanoparticles as Catalysts. Nano Lett. 2015, 15, 403–409. [Google Scholar] [CrossRef] [PubMed]
  9. Crista, D.M.A.; El Mragui, A.; Algarra, M.; da Silva, J.C.G.E.; Luque, R.; da Silva, L.P. Turning Spent Coffee Grounds into Sustainable Precursors for the Fabrication of Carbon Dots. Nanomaterials 2020, 10, 1209. [Google Scholar] [CrossRef]
  10. Qu, D.; Zheng, M.; Du, P.; Zhou, Y.; Zhang, L.G.; Li, D.; Tan, H.Q.; Zhao, Z.; Xie, Z.G.; Sun, Z.C. Highly luminescent S, N co-doped graphene quantum dots with broad visible absorption bands for visible light photocatalysts. Nanoscale 2013, 5, 12272–12277. [Google Scholar] [CrossRef]
  11. Gu, S.Y.; Hsieh, C.T.; Gandomi, Y.A.; Li, J.L.; Yue, X.X.; Chang, J.K. Tailoring fluorescence emissions, quantum yields, and white light emitting from nitrogen-doped graphene and carbon nitride quantum dots. Nanoscale 2019, 11, 16553–16561. [Google Scholar] [CrossRef]
  12. Han, Y.; Tang, B.J.; Wang, L.; Bao, H.; Lu, Y.H.; Guan, C.T.; Zhang, L.; Le, M.Y.; Liu, Z.; Wu, M.H. Machine-Learning-Driven Synthesis of Carbon Dots with Enhanced Quantum Yields. ACS Nano 2020, 14, 14761–14768. [Google Scholar] [CrossRef]
  13. Chao, D.Y.; Chen, J.X.; Dong, Q.; Wu, W.W.; Qi, D.S.; Dong, S.J. Ultrastable and ultrasensitive pH-switchable carbon dots with high quantum yield for water quality identification, glucose detection, and two starch-based solid-state fluorescence materials. Nano Res. 2020, 13, 3012–3018. [Google Scholar] [CrossRef]
  14. Manzur, A.; Oluwasanmi, A.; Moss, D.; Curtis, A.; Hoskins, C. Nanotechnologies in Pancreatic Cancer Therapy. Pharmaceutics 2017, 9, 39. [Google Scholar] [CrossRef] [Green Version]
  15. Messina, M.M.; Barrionuevo, S.D.; Coustet, M.E.; Kreuzer, M.P.; Saccone, F.D.; Claro, P.C.D.; Ibanez, F.J. Graphene and Carbon Dots for Photoanodes with Enhanced Performance. ACS Appl. Nano Mater. 2021, 4, 7309–7318. [Google Scholar] [CrossRef]
  16. Akbar, K.; Moretti, E.; Vomiero, A. Carbon Dots for Photocatalytic Degradation of Aqueous Pollutants: Recent Advancements. Adv. Opt. Mater. 2021, 9. [Google Scholar] [CrossRef]
  17. Shen, J.L.; Chen, W.F.; Yang, Z.H.; Lv, G.; Cao, J.; Li, D.Y.; Liu, X. A Critical Review of Graphene Quantum Dots: Synthesis and Application in Biosensors. Nano 2021, 16. [Google Scholar] [CrossRef]
  18. 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]
  19. Michaud, V.; Pracht, J.; Schilfarth, F.; Damm, C.; Platzer, B.; Haines, P.; Harreiss, C.; Guldi, D.M.; Spiecker, E.; Peukert, W. Well-separated water-soluble carbon dots via gradient chromatography. Nanoscale 2021, 13, 13116–13128. [Google Scholar] [CrossRef]
  20. Kurniawan, D.; Chiang, W.H. Microplasma-enabled colloidal nitrogen-doped graphene quantum dots for broad-range fluorescent pH sensors. Carbon 2020, 167, 675–684. [Google Scholar] [CrossRef]
  21. Ru, Y.; Ai, L.; Jia, T.T.; Liu, X.J.; Lu, S.Y.; Tang, Z.Y.; Yang, B. Recent advances in chiral carbonized polymer dots: From synthesis and properties to applications. Nano Today 2020, 34. [Google Scholar] [CrossRef]
  22. Ding, H.; Ji, Y.; Wei, J.-S.; Gao, Q.-Y.; Zhou, Z.-Y.; Xiong, H.-M. Facile synthesis of red-emitting carbon dots from pulp-free lemon juice for bioimaging. J. Mater. Chem. B 2017, 5, 5272–5277. [Google Scholar] [CrossRef] [PubMed]
  23. Liu, J.; Geng, Y.; Li, D.; Yao, H.; Huo, Z.; Li, Y.; Zhang, K.; Zhu, S.; Wei, H.; Xu, W.; et al. Deep Red Emissive Carbonized Polymer Dots with Unprecedented Narrow Full Width at Half Maximum. Adv. Mater. 2020, 32, 1906641. [Google Scholar] [CrossRef] [PubMed]
  24. Liang, C.; Xie, X.; Zhang, D.; Feng, J.; Lu, S.; Shi, Q. Biomass carbon dots derived from Wedelia trilobata for the direct detection of glutathione and their imaging application in living cells. J. Mater. Chem. B 2021, 9, 5670–5681. [Google Scholar] [CrossRef]
  25. Wu, L.; Long, R.; Li, T.; Tang, C.; Tong, X.; Guo, Y.; Shi, S.; Xiang, H.; Tong, C. One-pot fabrication of dual-emission and single-emission biomass carbon dots for Cu2+ and tetracycline sensing and multicolor cellular imaging. Anal. Bioanal. Chem. 2020, 412, 7481–7489. [Google Scholar] [CrossRef] [PubMed]
  26. Ding, H.; Zhou, X.; Qin, B.; Zhou, Z.; Zhao, Y. Highly fluorescent near-infrared emitting carbon dots derived from lemon juice and its bioimaging application. J. Lumin. 2019, 211, 298–304. [Google Scholar] [CrossRef]
  27. Zhu, J.; Chu, H.; Shen, J.; Wang, C.; Wei, Y. Green preparation of carbon dots from plum as a ratiometric fluorescent probe for detection of doxorubicin. Opt. Mater. 2021, 114, 110941. [Google Scholar] [CrossRef]
  28. Liu, L.; Zhang, S.; Zheng, X.; Li, H.; Chen, Q.; Qin, K.; Ding, Y.; Wei, Y. Carbon dots derived from Fusobacterium nucleatum for intracellular determination of Fe3+ and bioimaging both in vitro and in vivo. Anal. Methods 2021, 13, 1121–1131. [Google Scholar] [CrossRef]
  29. Wang, W.; Chen, J.; Wang, D.; Shen, Y.; Yang, L.; Zhang, T.; Ge, J. Facile synthesis of biomass waste-derived fluorescent N, S, P co-doped carbon dots for detection of Fe3+ ions in solutions and living cells. Anal. Methods 2021, 13, 789–795. [Google Scholar] [CrossRef]
  30. Wang, N.; Wang, Y.; Guo, T.; Yang, T.; Chen, M.; Wang, J. Green preparation of carbon dots with papaya as carbon source for effective fluorescent sensing of Iron (III) and Escherichia coli. Biosens. Bioelectron. 2016, 85, 68–75. [Google Scholar] [CrossRef]
  31. Edison, T.N.J.I.; Atchudan, R.; Shim, J.-J.; Kalimuthu, S.; Ahn, B.-C.; Lee, Y.R. Turn-off fluorescence sensor for the detection of ferric ion in water using green synthesized N-doped carbon dots and its bio-imaging. J. Photochem. Photobiol. B Biol. 2016, 158, 235–242. [Google Scholar] [CrossRef]
  32. Yang, X.; Zhuo, Y.; Zhu, S.; Luo, Y.; Feng, Y.; Dou, Y. Novel and green synthesis of high-fluorescent carbon dots originated from honey for sensing and imaging. Biosens. Bioelectron. 2014, 60, 292–298. [Google Scholar] [CrossRef] [PubMed]
  33. Shen, J.; Shang, S.; Chen, X.; Wang, D.; Cai, Y. Facile synthesis of fluorescence carbon dots from sweet potato for Fe3+ sensing and cell imaging. Mater. Sci. Eng. C 2017, 76, 856–864. [Google Scholar] [CrossRef]
  34. Song, P.; Zhang, L.; Long, H.; Meng, M.; Liu, T.; Yin, Y.; Xi, R. A multianalyte fluorescent carbon dots sensing system constructed based on specific recognition of Fe(iii) ions. RSC Adv. 2017, 7, 28637–28646. [Google Scholar] [CrossRef] [Green Version]
  35. Liu, W.; Zhang, R.; Kang, Y.; Zhang, X.Y.; Wang, H.J.; Li, L.H.; Diao, H.P.; Wei, W.L. Preparation of nitrogen-doped carbon dots with a high fluorescence quantum yield for the highly sensitive detection of Cu2+ ions, drawing anti-counterfeit patterns and imaging live cells. New Carbon Mater. 2019, 34, 390–401. [Google Scholar] [CrossRef]
  36. Atchudan, R.; Edison, T.N.J.I.; Perumal, S.; Vinodh, R.; Sundramoorthy, A.K.; Babu, R.S.; Lee, Y.R. Leftover Kiwi Fruit Peel-Derived Carbon Dots as a Highly Selective Fluorescent Sensor for Detection of Ferric Ion. Chemosensors 2021, 9, 166. [Google Scholar] [CrossRef]
  37. Liu, R.; Zhang, J.; Gao, M.; Li, Z.; Chen, J.; Wu, D.; Liu, P. A facile microwave-hydrothermal approach towards highly photoluminescent carbon dots from goose feathers. RSC Adv. 2015, 5, 4428–4433. [Google Scholar] [CrossRef]
  38. Zulfajri, M.; Gedda, G.; Chang, C.-J.; Chang, Y.-P.; Huang, G.G. Cranberry Beans Derived Carbon Dots as a Potential Fluorescence Sensor for Selective Detection of Fe3+ Ions in Aqueous Solution. ACS Omega 2019, 4, 15382–15392. [Google Scholar] [CrossRef] [Green Version]
  39. Xu, J.; Zhou, Y.; Liu, S.; Dong, M.; Huang, C. Low-cost synthesis of carbon nanodots from natural products used as a fluorescent probe for the detection of ferrum(iii) ions in lake water. Anal. Methods 2014, 6, 2086. [Google Scholar] [CrossRef]
  40. Venkatesan, G.; Rajagopalan, V.; Chakravarthula, S.N. Boswellia ovalifoliolata bark extract derived carbon dots for selective fluorescent sensing of Fe3+. J. Environ. Chem. Eng. 2019, 7, 103013. [Google Scholar] [CrossRef]
  41. Zhou, J.; Ge, M.; Han, Y.; Ni, J.; Huang, X.; Han, S.; Peng, Z.; Li, Y.; Li, S. Preparation of Biomass-Based Carbon Dots with Aggregation Luminescence Enhancement from Hydrogenated Rosin for Biological Imaging and Detection of Fe3+. ACS Omega 2020, 5, 11842–11848. [Google Scholar] [CrossRef] [PubMed]
  42. Sachdev, A.; Gopinath, P. Green synthesis of multifunctional carbon dots from coriander leaves and their potential application as antioxidants, sensors and bioimaging agents. Analyst 2015, 140, 4260–4269. [Google Scholar] [PubMed]
  43. Raveendran, V.; Suresh Babu, A.R.; Renuka, N.K. Mint leaf derived carbon dots for dual analyte detection of Fe(iii) and ascorbic acid. RSC Adv. 2019, 9, 12070–12077. [Google Scholar] [CrossRef] [Green Version]
  44. Hu, Y.; Li, J.; Li, X. Leek-derived co-doped carbon dots as efficient fluorescent probes for dichlorvos sensitive detection and cell multicolor imaging. Anal. Bioanal. Chem. 2019, 411, 7879–7887. [Google Scholar] [CrossRef] [PubMed]
  45. Liao, J.; Cheng, Z.; Zhou, L. Nitrogen-Doping Enhanced Fluorescent Carbon Dots: Green Synthesis and Their Applications for Bioimaging and Label-Free Detection of Au3+ Ions. ACS Sustain. Chem. Eng. 2016, 4, 3053–3061. [Google Scholar] [CrossRef]
  46. Miao, H.; Wang, L.; Zhuo, Y.; Zhou, Z.; Yang, X. Label-free fluorimetric detection of CEA using carbon dots derived from tomato juice. Biosens. Bioelectron. 2016, 86, 83–89. [Google Scholar] [CrossRef]
  47. Lu, H.; Li, C.; Wang, H.; Wang, X.; Xu, S. Biomass-Derived Sulfur, Nitrogen Co-Doped Carbon Dots for Colorimetric and Fluorescent Dual Mode Detection of Silver (I) and Cell Imaging. ACS Omega 2019, 4, 21500–21508. [Google Scholar] [CrossRef] [Green Version]
  48. Singh, A.K.; Singh, V.K.; Singh, M.; Singh, P.; Khadim, S.R.; Singh, U.; Koch, B.; Hasan, S.H.; Asthana, R.K. One pot hydrothermal synthesis of fluorescent NP-carbon dots derived from Dunaliella salina biomass and its application in on-off sensing of Hg (II), Cr (VI) and live cell imaging. J. Photochem. Photobiol. A 2019, 376, 63–72. [Google Scholar] [CrossRef]
  49. Tsai, S.R.; Yin, R.; Huang, Y.Y.; Sheu, B.C.; Lee, S.C.; Hamblin, M.R. Low-level light therapy potentiates NPe6-mediated photodynamic therapy in a human osteosarcoma cell line via increased ATP. Photodiagn. Photodyn. 2015, 12, 123–130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Lu, W.; Qin, X.; Liu, S.; Chang, G.; Zhang, Y.; Luo, Y.; Asiri, A.M.; Al-Youbi, A.O.; Sun, X. Economical, Green Synthesis of Fluorescent Carbon Nanoparticles and Their Use as Probes for Sensitive and Selective Detection of Mercury(II) Ions. Anal. Chem. 2012, 84, 5351–5357. [Google Scholar] [CrossRef]
  51. Huang, H.; Lv, J.-J.; Zhou, D.-L.; Bao, N.; Xu, Y.; Wang, A.-J.; Feng, J.-J. One-pot green synthesis of nitrogen-doped carbon nanoparticles as fluorescent probes for mercury ions. RSC Adv. 2013, 3, 21691. [Google Scholar] [CrossRef]
  52. Wang, C.; Sun, D.; Zhuo, K.; Zhang, H.; Wang, J. Simple and green synthesis of nitrogen-, sulfur-, and phosphorus-co-doped carbon dots with tunable luminescence properties and sensing application. RSC Adv. 2014, 4, 54060–54065. [Google Scholar] [CrossRef]
  53. Xie, Y.; Cheng, D.; Liu, X.; Han, A. Green Hydrothermal Synthesis of N-doped Carbon Dots from Biomass Highland Barley for the Detection of Hg2+. Sensors 2019, 19, 3169. [Google Scholar] [CrossRef] [Green Version]
  54. Tyagi, A.; Tripathi, K.M.; Singh, N.; Choudhary, S.; Gupta, R.K. Green synthesis of carbon quantum dots from lemon peel waste: Applications in sensing and photocatalysis. RSC Adv. 2016, 6, 72423–72432. [Google Scholar] [CrossRef]
  55. Ghereghlou, M.; Esmaeili, A.A.; Darroudi, M. Green Synthesis of Fluorescent Carbon Dots from Elaeagnus angustifolia and its Application as Tartrazine Sensor. J. Fluoresc. 2021, 31, 185–193. [Google Scholar] [CrossRef] [PubMed]
  56. Xu, H.; Yang, X.; Li, G.; Zhao, C.; Liao, X. Green Synthesis of Fluorescent Carbon Dots for Selective Detection of Tartrazine in Food Samples. J. Agric. Food Chem. 2015, 63, 6707–6714. [Google Scholar] [CrossRef]
  57. Purbia, R.; Paria, S. A simple turn on fluorescent sensor for the selective detection of thiamine using coconut water derived luminescent carbon dots. Biosens. Bioelectron. 2016, 79, 467–475. [Google Scholar] [CrossRef]
  58. Amjadi, M.; Hallaj, T.; Mayan, M.A. Green synthesis of nitrogen-doped carbon dots from lentil and its application for colorimetric determination of thioridazine hydrochloride. RSC Adv. 2016, 6, 104467–104473. [Google Scholar] [CrossRef]
  59. Akhgari, F.; Samadi, N.; Farhadi, K.; Akhgari, M. A green one-pot synthesis of nitrogen and sulfur co-doped carbon quantum dots for sensitive and selective detection of cephalexin. Can. J. Chem. 2017, 95, 641–648. [Google Scholar] [CrossRef]
  60. Liu, Y.; Zhao, Y.; Zhang, Y. One-step green synthesized fluorescent carbon nanodots from bamboo leaves for copper(II) ion detection. Sens. Actuators B Chem. 2014, 196, 647–652. [Google Scholar] [CrossRef]
  61. Sha, Y.; Lou, J.; Bai, S.; Wu, D.; Liu, B.; Ling, Y. Hydrothermal synthesis of nitrogen-containing carbon nanodots as the high-efficient sensor for copper(II) ions. Mater. Res. Bull. 2013, 48, 1728–1731. [Google Scholar] [CrossRef]
  62. Mehta, V.N.; Jha, S.; Basu, H.; Singhal, R.K.; Kailasa, S.K. One-step hydrothermal approach to fabricate carbon dots from apple juice for imaging of mycobacterium and fungal cells. Sens. Actuators B Chem. 2015, 213, 434–443. [Google Scholar] [CrossRef]
  63. Arul, V.; Edison, T.N.J.I.; Lee, Y.R.; Sethuraman, M.G. Biological and catalytic applications of green synthesized fluorescent N-doped carbon dots using Hylocereus undatus. J. Photochem. Photobiol. B Biol. 2017, 168, 142–148. [Google Scholar]
  64. Mehta, V.N.; Jha, S.; Kailasa, S.K. One-pot green synthesis of carbon dots by using Saccharum officinarum juice for fluorescent imaging of bacteria (Escherichia coli) and yeast (Saccharomyces cerevisiae) cells. Mater. Sci. Eng. C 2014, 38, 20–27. [Google Scholar] [CrossRef] [PubMed]
  65. Song, Y.; Yan, X.; Li, Z.; Qu, L.; Zhu, C.; Ye, R.; Li, S.; Du, D.; Lin, Y. Highly photoluminescent carbon dots derived from linseed and their applications in cellular imaging and sensing. J. Mater. Chem. B 2018, 6, 3181–3187. [Google Scholar] [CrossRef]
  66. Wang, W.-J.; Xia, J.-M.; Feng, J.; He, M.-Q.; Chen, M.-L.; Wang, J.-H. Green preparation of carbon dots for intracellular pH sensing and multicolor live cell imaging. J. Mater. Chem. B 2016, 4, 7130–7137. [Google Scholar] [CrossRef] [PubMed]
  67. Zhao, X.J.; Zhang, W.L.; Zhou, Z.Q. Sodium hydroxide-mediated hydrogel of citrus pectin for preparation of fluorescent carbon dots for bioimaging. Colloids Surf. B Biointerfaces 2014, 123, 493–497. [Google Scholar] [CrossRef]
  68. Liu, Y.; Liu, Y.; Park, M.; Park, S.-J.; Zhang, Y.; Akanda, M.R.; Park, B.-Y.; Kim, H.Y. Green synthesis of fluorescent carbon dots from carrot juice for in vitro cellular imaging. Carbon Lett. 2017, 21, 61–67. [Google Scholar]
  69. Atchudan, R.; Edison, T.N.J.I.; Perumal, S.; Muthuchamy, N.; Lee, Y.R. Hydrophilic nitrogen-doped carbon dots from biowaste using dwarf banana peel for environmental and biological applications. Fuel 2020, 275, 117821. [Google Scholar] [CrossRef]
  70. Du, F.; Zhang, M.; Li, X.; Li, J.; Jiang, X.; Li, Z.; Hua, Y.; Shao, G.; Jin, J.; Shao, Q. Economical and green synthesis of bagasse-derived fluorescent carbon dots for biomedical applications. Nanotechnology 2014, 25, 315702. [Google Scholar] [CrossRef]
  71. Alam, A.-M.; Park, B.-Y.; Ghouri, Z.K.; Park, M.; Kim, H.-Y. Synthesis of carbon quantum dots from cabbage with down- and up-conversion photoluminescence properties: Excellent imaging agent for biomedical applications. Green Chem. 2015, 17, 3791–3797. [Google Scholar] [CrossRef]
  72. Ding, Z.; Li, F.; Wen, J.; Wang, X.; Sun, R. Gram-scale synthesis of single-crystalline graphene quantum dots derived from lignin biomass. Green Chem. 2018, 20, 1383–1390. [Google Scholar] [CrossRef]
  73. D’Souza, S.L.; Deshmukh, B.; Bhamore, J.R.; Rawat, K.A.; Lenka, N.; Kailasa, S.K. Synthesis of fluorescent nitrogen-doped carbon dots from dried shrimps for cell imaging and boldine drug delivery system. RSC Adv. 2016, 6, 12169–12179. [Google Scholar] [CrossRef]
  74. John, T.S.; Yadav, P.K.; Kumar, D.; Singh, S.K.; Hasan, S.H. Highly fluorescent carbon dots from wheat bran as a novel drug delivery system for bacterial inhibition. Luminescence 2020, 35, 913–923. [Google Scholar] [CrossRef] [PubMed]
  75. Yuan, Y.; Guo, B.; Hao, L.; Liu, N.; Lin, Y.; Guo, W.; Li, X.; Gu, B. Doxorubicin-loaded environmentally friendly carbon dots as a novel drug delivery system for nucleus targeted cancer therapy. Colloids Surf. B Biointerfaces 2017, 159, 349–359. [Google Scholar] [CrossRef] [PubMed]
  76. Sun, B.; Wu, F.; Zhang, Q.; Chu, X.; Wang, Z.; Huang, X.; Li, J.; Yao, C.; Zhou, N.; Shen, J. Insight into the effect of particle size distribution differences on the antibacterial activity of carbon dots. J. Colloid Interface Sci. 2021, 584, 505–519. [Google Scholar] [CrossRef]
  77. Saravanan, A.; Maruthapandi, M.; Das, P.; Luong, J.H.T.; Gedanken, A. Green Synthesis of Multifunctional Carbon Dots with Antibacterial Activities. Nanomaterials 2021, 11, 369. [Google Scholar] [CrossRef]
  78. Eskalen, H.; Çeşme, M.; Kerli, S.; Özğan, Ş. Green synthesis of water-soluble fluorescent carbon dots from rosemary leaves: Applications in food storage capacity, fingerprint detection, and antibacterial activity. J. Chem. Res. 2021, 45, 428–435. [Google Scholar] [CrossRef]
  79. Ma, Y.; Zhang, M.; Wang, H.; Wang, B.; Huang, H.; Liu, Y.; Kang, Z. N-doped carbon dots derived from leaves with low toxicity via damaging cytomembrane for broad-spectrum antibacterial activity. Mater. Today Commun. 2020, 24, 101222. [Google Scholar] [CrossRef]
  80. Boobalan, T.; Sethupathi, M.; Sengottuvelan, N.; Kumar, P.; Balaji, P.; Gulyás, B.; Padmanabhan, P.; Selvan, S.T.; Arun, A. Mushroom-Derived Carbon Dots for Toxic Metal Ion Detection and as Antibacterial and Anticancer Agents. ACS Appl. Nano Mater. 2020, 3, 5910–5919. [Google Scholar] [CrossRef]
  81. Castaing, V.; Sontakke, A.D.; Xu, J.; Fernandez-Carrion, A.J.; Genevois, C.; Tanabe, S.; Allix, M.; Viana, B. Persistent energy transfer in ZGO:Cr3+,Yb3+: A new strategy to design nano glass-ceramics featuring deep red and near infrared persistent luminescence. Phys. Chem. Chem. Phys. 2019, 21, 19458–19468. [Google Scholar] [CrossRef]
  82. Jia, Q.; Zheng, X.; Ge, J.; Liu, W.; Ren, H.; Chen, S.; Wen, Y.; Zhang, H.; Wu, J.; Wang, P. Synthesis of carbon dots from Hypocrella bambusae for bimodel fluorescence/photoacoustic imaging-guided synergistic photodynamic/photothermal therapy of cancer. J. Colloid Interface Sci. 2018, 526, 302–311. [Google Scholar] [CrossRef] [PubMed]
  83. Kim, D.; Jo, G.; Chae, Y.; Subramani, S.; Lee, B.Y.; Kim, E.J.; Ji, M.-K.; Sim, U.; Hyun, H. Bioinspired Camellia japonica carbon dots with high near-infrared absorbance for efficient photothermal cancer therapy. Nanoscale 2021, 13, 14426–14434. [Google Scholar] [CrossRef] [PubMed]
  84. Li, C.-L.; Ou, C.-M.; Huang, C.-C.; Wu, W.-C.; Chen, Y.-P.; Lin, T.-E.; Ho, L.-C.; Wang, C.-W.; Shih, C.-C.; Zhou, H.-C.; et al. Carbon dots prepared from ginger exhibiting efficient inhibition of human hepatocellular carcinoma cells. J. Mater. Chem. B 2014, 2, 4564. [Google Scholar] [CrossRef] [PubMed]
  85. Zhao, S.; Lan, M.; Zhu, X.; Xue, H.; Ng, T.-W.; Meng, X.; Lee, C.-S.; Wang, P.; Zhang, W. Green Synthesis of Bifunctional Fluorescent Carbon Dots from Garlic for Cellular Imaging and Free Radical Scavenging. ACS Appl. Mater. Inter. 2015, 7, 17054–17060. [Google Scholar] [CrossRef]
  86. Chen, W.; Li, D.; Tian, L.; Xiang, W.; Wang, T.; Hu, W.; Hu, Y.; Chen, S.; Chen, J.; Dai, Z. Synthesis of graphene quantum dots from natural polymer starch for cell imaging. Green Chem. 2018, 20, 4438–4442. [Google Scholar] [CrossRef]
  87. Sahu, S.; Behera, B.; Maiti, T.K.; Mohapatra, S. Simple one-step synthesis of highly luminescent carbon dots from orange juice: Application as excellent bio-imaging agents. Chem. Commun. 2012, 48, 8835. [Google Scholar] [CrossRef] [PubMed]
  88. Zhang, J.; Yuan, Y.; Liang, G.; Yu, S.-H. Scale-Up Synthesis of Fragrant Nitrogen-Doped Carbon Dots from Bee Pollens for Bioimaging and Catalysis. Adv. Sci. 2015, 2, 1500002. [Google Scholar] [CrossRef]
  89. Liang, Q.; Ma, W.; Shi, Y.; Li, Z.; Yang, X. Easy synthesis of highly fluorescent carbon quantum dots from gelatin and their luminescent properties and applications. Carbon 2013, 60, 421–428. [Google Scholar] [CrossRef]
  90. Kasibabu, B.S.B.; D’Souza, S.L.; Jha, S.; Kailasa, S.K. Imaging of Bacterial and Fungal Cells Using Fluorescent Carbon Dots Prepared from Carica papaya Juice. J. Fluoresc. 2015, 25, 803–810. [Google Scholar] [CrossRef]
  91. Yu, C.; Xuan, T.; Chen, Y.; Zhao, Z.; Sun, Z.; Li, H. A facile, green synthesis of highly fluorescent carbon nanoparticles from oatmeal for cell imaging. J. Mater. Chem. C 2015, 3, 9514–9518. [Google Scholar] [CrossRef]
  92. Zhang, Z.; Sun, W.; Wu, P. Highly Photoluminescent Carbon Dots Derived from Egg White: Facile and Green Synthesis, Photoluminescence Properties, and Multiple Applications. ACS Sustain. Chem. Eng. 2015, 3, 1412–1418. [Google Scholar] [CrossRef]
  93. Wei, J.; Zhang, X.; Sheng, Y.; Shen, J.; Huang, P.; Guo, S.; Pan, J.; Feng, B. Dual functional carbon dots derived from cornflour via a simple one-pot hydrothermal route. Mater. Lett. 2014, 123, 107–111. [Google Scholar] [CrossRef]
  94. 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]
  95. Wang, G.; Guo, Q.; Chen, D.; Liu, Z.; Zheng, X.; Xu, A.; Yang, S.; Ding, G. Facile and Highly Effective Synthesis of Controllable Lattice Sulfur-Doped Graphene Quantum Dots via Hydrothermal Treatment of Durian. ACS Appl. Mater. Inter. 2018, 10, 5750–5759. [Google Scholar] [CrossRef]
  96. Atchudan, R.; Jebakumar Immanuel Edison, T.N.; Perumal, S.; Lee, Y.R. Indian Gooseberry-Derived Tunable Fluorescent Carbon Dots as a Promise for In Vitro/In Vivo Multicolor Bioimaging and Fluorescent Ink. ACS Omega 2018, 3, 17590–17601. [Google Scholar] [CrossRef]
  97. Hou, J.; Wang, W.; Zhou, T.Y.; Wang, B.; Li, H.Y.; Ding, L. Synthesis and formation mechanistic investigation of nitrogen-doped carbon dots with high quantum yields and yellowish-green fluorescence. Nanoscale 2016, 8, 11185–11193. [Google Scholar] [CrossRef]
  98. Meena, R.; Singh, R.; Marappan, G.; Kushwaha, G.; Gupta, N.; Meena, R.; Gupta, J.P.; Agarwal, R.R.; Fahmi, N.; Kushwaha, O.S. Fluorescent carbon dots driven from ayurvedic medicinal plants for cancer cell imaging and phototherapy. Heliyon 2019, 5, e02483. [Google Scholar] [CrossRef] [Green Version]
  99. Zhang, X.; Wang, H.; Ma, C.; Niu, N.; Chen, Z.; Liu, S.; Li, J.; Li, S. Seeking value from biomass materials: Preparation of coffee bean shell-derived fluorescent carbon dots via molecular aggregation for antioxidation and bioimaging applications. Mater. Chem. Front. 2018, 2, 1269–1275. [Google Scholar] [CrossRef]
  100. Desai, M.L.; Jha, S.; Basu, H.; Singhal, R.K.; Park, T.-J.; Kailasa, S.K. Acid Oxidation of Muskmelon Fruit for the Fabrication of Carbon Dots with Specific Emission Colors for Recognition of Hg2+ Ions and Cell Imaging. ACS Omega 2019, 4, 19332–19340. [Google Scholar] [CrossRef] [Green Version]
  101. D’Angelis do ES Barbosa, C.; Corrêa, J.R.; Medeiros, G.A.; Barreto, G.; Magalhães, K.G.; De Oliveira, A.L.; Spencer, J.; Rodrigues, M.O.; Neto, B.A.D. Carbon Dots (C-dots) from Cow Manure with Impressive Subcellular Selectivity Tuned by Simple Chemical Modification. Chem. Eur. J. 2015, 21, 5055–5060. [Google Scholar] [CrossRef] [PubMed]
  102. Park, S.Y.; Lee, H.U.; Park, E.S.; Lee, S.C.; Lee, J.-W.; Jeong, S.W.; Kim, C.H.; Lee, Y.-C.; Huh, Y.S.; Lee, J. Photoluminescent Green Carbon Nanodots from Food-Waste-Derived Sources: Large-Scale Synthesis, Properties, and Biomedical Applications. ACS Appl. Mater. Inter. 2014, 6, 3365–3370. [Google Scholar] [CrossRef]
  103. Oza, G.; Oza, K.; Pandey, S.; Shinde, S.; Mewada, A.; Thakur, M.; Sharon, M.; Sharon, M. A Green Route Towards Highly Photoluminescent and Cytocompatible Carbon dot Synthesis and its Separation Using Sucrose Density Gradient Centrifugation. J. Fluoresc. 2015, 25, 9–14. [Google Scholar] [CrossRef] [PubMed]
  104. Dehvari, K.; Liu, K.Y.; Tseng, P.-J.; Gedda, G.; Girma, W.M.; Chang, J.-Y. Sonochemical-assisted green synthesis of nitrogen-doped carbon dots from crab shell as targeted nanoprobes for cell imaging. J. Taiwan Inst. Chem. Eng. 2019, 95, 495–503. [Google Scholar] [CrossRef]
  105. Feng, J.; Wang, W.-J.; Hai, X.; Yu, Y.-L.; Wang, J.-H. Green preparation of nitrogen-doped carbon dots derived from silkworm chrysalis for cell imaging. J. Mater. Chem. B 2016, 4, 387–393. [Google Scholar] [CrossRef] [PubMed]
  106. Ramanan, V.; Thiyagarajan, S.K.; Raji, K.; Suresh, R.; Sekar, R.; Ramamurthy, P. Outright Green Synthesis of Fluorescent Carbon Dots from Eutrophic Algal Blooms for In Vitro Imaging. ACS Sustain. Chem. Eng. 2016, 4, 4724–4731. [Google Scholar] [CrossRef]
  107. Wang, Q.; Liu, X.; Zhang, L.; Lv, Y. Microwave-assisted synthesis of carbon nanodots through an eggshell membrane and their fluorescent application. Analyst 2012, 137, 5392. [Google Scholar] [CrossRef] [PubMed]
  108. Qin, X.; Lu, W.; Asiri, A.M.; Al-Youbi, A.O.; Sun, X. Microwave-assisted rapid green synthesis of photoluminescent carbon nanodots from flour and their applications for sensitive and selective detection of mercury(II) ions. Sens. Actuators B Chem. 2013, 184, 156–162. [Google Scholar] [CrossRef]
  109. Liu, X.; Li, T.; Hou, Y.; Wu, Q.; Yi, J.; Zhang, G. Microwave synthesis of carbon dots with multi-response using denatured proteins as carbon source. RSC Adv. 2016, 6, 11711–11718. [Google Scholar] [CrossRef]
  110. Feng, Y.; Zhong, D.; Miao, H.; Yang, X. Carbon dots derived from rose flowers for tetracycline sensing. Talanta 2015, 140, 128–133. [Google Scholar] [CrossRef] [PubMed]
  111. Bankoti, K.; Rameshbabu, A.P.; Datta, S.; Das, B.; Mitra, A.; Dhara, S. Onion derived carbon nanodots for live cell imaging and accelerated skin wound healing. J. Mater. Chem. B 2017, 5, 6579–6592. [Google Scholar] [CrossRef]
  112. Xue, M.; Zou, M.; Zhao, J.; Zhan, Z.; Zhao, S. Green preparation of fluorescent carbon dots from lychee seeds and their application for the selective detection of methylene blue and imaging in living cells. J. Mater. Chem. B 2015, 3, 6783–6789. [Google Scholar] [CrossRef]
  113. Hsu, P.-C.; Shih, Z.-Y.; Lee, C.-H.; Chang, H.-T. Synthesis and analytical applications of photoluminescent carbon nanodots. Green Chem. 2012, 14, 917. [Google Scholar] [CrossRef]
  114. Essner, J.B.; Laber, C.H.; Ravula, S.; Polo-Parada, L.; Baker, G.A. Pee-dots: Biocompatible fluorescent carbon dots derived from the upcycling of urine. Green Chem. 2016, 18, 243–250. [Google Scholar] [CrossRef] [Green Version]
  115. Zhou, J.; Sheng, Z.; Han, H.; Zou, M.; Li, C. Facile synthesis of fluorescent carbon dots using watermelon peel as a carbon source. Mater. Lett. 2012, 66, 222–224. [Google Scholar] [CrossRef]
  116. Teng, X.; Ma, C.; Ge, C.; Yan, M.; Yang, J.; Zhang, Y.; Morais, P.C.; Bi, H. Green synthesis of nitrogen-doped carbon dots from konjac flour with “off–on” fluorescence by Fe3+ and l-lysine for bioimaging. J. Mater. Chem. B 2014, 2, 4631. [Google Scholar] [CrossRef] [PubMed]
  117. Hu, Y.; Chen, Z.; Lai, F.; Li, J. Biomass-codoped carbon dots: Efficient fluorescent probes for isocarbophos ultrasensitive detection and for living cells dual-color imaging. J. Mater. Sci. 2019, 54, 8627–8639. [Google Scholar] [CrossRef]
  118. Murugan, N.; Sundramoorthy, A.K. Green synthesis of fluorescent carbon dots from Borassus flabellifer flowers for label-free highly selective and sensitive detection of Fe3+ ions. New J. Chem. 2018, 42, 13297–13307. [Google Scholar] [CrossRef]
  119. Ma, X.; Dong, Y.; Sun, H.; Chen, N. Highly fluorescent carbon dots from peanut shells as potential probes for copper ion: The optimization and analysis of the synthetic process. Mater. Today Chem. 2017, 5, 1–10. [Google Scholar] [CrossRef]
  120. Baruah, U.; Gogoi, N.; Konwar, A.; Jyoti Deka, M.; Chowdhury, D.; Majumdar, G. Carbon Dot Based Sensing of Dopamine and Ascorbic Acid. J. Nanoparticles 2014, 2014, 1–8. [Google Scholar] [CrossRef]
  121. Xue, M.; Zhan, Z.; Zou, M.; Zhang, L.; Zhao, S. Green synthesis of stable and biocompatible fluorescent carbon dots from peanut shells for multicolor living cell imaging. New J. Chem. 2016, 40, 1698–1703. [Google Scholar] [CrossRef]
  122. Cong, S.; Liu, K.; Qiao, F.; Song, Y.; Tan, M. Biocompatible fluorescent carbon dots derived from roast duck for in vitro cellular and in vivo C. elegans bio-imaging. Methods 2019, 168, 76–83. [Google Scholar] [CrossRef]
  123. Wang, H.; Zhang, M.; Ma, Y.; Wang, B.; Shao, M.; Huang, H.; Liu, Y.; Kang, Z. Selective inactivation of Gram-negative bacteria by carbon dots derived from natural biomass: Artemisia argyi leaves. J. Mater. Chem. B 2020, 8, 2666–2672. [Google Scholar] [CrossRef]
  124. Chung, H.K.; Wongso, V.; Sambudi, N.S.; Isnaeni. Biowaste-derived carbon dots/hydroxyapatite nanocomposite as drug delivery vehicle for acetaminophen. J. Sol-Gel Sci. Technol. 2020, 93, 214–223. [Google Scholar] [CrossRef]
  125. Wang, X.; Zhang, Y.; Kong, H.; Cheng, J.; Zhang, M.; Sun, Z.; Wang, S.; Liu, J.; Qu, H.; Zhao, Y. Novel mulberry silkworm cocoon-derived carbon dots and their anti-inflammatory properties. Artif. Cells Nanomed. Biotechnol. 2020, 48, 68–76. [Google Scholar] [CrossRef] [PubMed]
  126. Xue, M.; Zhao, J.; Zhan, Z.; Zhao, S.; Lan, C.; Ye, F.; Liang, H. Dual functionalized natural biomass carbon dots from lychee exocarp for cancer cell targetable near-infrared fluorescence imaging and photodynamic therapy. Nanoscale 2018, 10, 18124–18130. [Google Scholar] [CrossRef] [PubMed]
  127. Fahmi, M.Z.; Haris, A.; Permana, A.J.; Nor Wibowo, D.L.; Purwanto, B.; Nikmah, Y.L.; Idris, A. Bamboo leaf-based carbon dots for efficient tumor imaging and therapy. RSC Adv. 2018, 8, 38376–38383. [Google Scholar] [CrossRef] [Green Version]
  128. Cheng, C.; Shi, Y.; Li, M.; Xing, M.; Wu, Q. Carbon quantum dots from carbonized walnut shells: Structural evolution, fluorescence characteristics, and intracellular bioimaging. Mater. Sci. Eng. C 2017, 79, 473–480. [Google Scholar] [CrossRef]
  129. Janus, Ł.; Piątkowski, M.; Radwan-Pragłowska, J.; Bogdał, D.; Matysek, D. Chitosan-based carbon quantum dots for biomedical applications: Synthesis and characterization. Nanomaterials 2019, 9, 274. [Google Scholar] [CrossRef] [Green Version]
  130. Qu, D.; Zheng, M.; Zhang, L.; Zhao, H.; Xie, Z.; Jing, X.; Haddad, R.E.; Fan, H.; Sun, Z. Formation mechanism and optimization of highly luminescent N-doped graphene quantum dots. Sci. Rep. 2014, 4, 1–11. [Google Scholar] [CrossRef]
  131. Zhang, R.; Chen, W. Nitrogen-doped carbon quantum dots: Facile synthesis and application as a “turn-off” fluorescent probe for detection of Hg2+ ions. Biosens. Bioelectron. 2014, 55, 83–90. [Google Scholar] [CrossRef] [PubMed]
  132. Kurian, M.; Paul, A. Recent trends in the use of green sources for carbon dot synthesis-A short review. Carbon Trends 2021, 100032. [Google Scholar] [CrossRef]
  133. Lai, Z.; Guo, X.; Cheng, Z.; Ruan, G.; Du, F. Green synthesis of fluorescent carbon dots from cherry tomatoes for highly effective detection of trifluralin herbicide in soil samples. ChemistrySelect 2020, 5, 1956–1960. [Google Scholar]
  134. Jiao, X.-Y.; Li, L.-s.; Qin, S.; Zhang, Y.; Huang, K.; Xu, L. The synthesis of fluorescent carbon dots from mango peel and their multiple applications. Colloids Surfaces A Physicochem. Eng. Asp. 2019, 577, 306–314. [Google Scholar] [CrossRef]
  135. Lu, M.; Duan, Y.; Song, Y.; Tan, J.; Zhou, L. Green preparation of versatile nitrogen-doped carbon quantum dots from watermelon juice for cell imaging, detection of Fe3+ ions and cysteine, and optical thermometry. J. Mol. Liq. 2018, 269, 766–774. [Google Scholar] [CrossRef]
  136. Shi, Y.; Liu, X.; Wang, M.; Huang, J.; Jiang, X.; Pang, J.; Xu, F.; Zhang, X. Synthesis of N-doped carbon quantum dots from bio-waste lignin for selective irons detection and cellular imaging. Int. J. Biol. Macromol. 2019, 128, 537–545. [Google Scholar] [CrossRef]
  137. Atchudan, R.; Edison, T.N.J.I.; Perumal, S.; Selvam, N.C.S.; Lee, Y.R. Green synthesized multiple fluorescent nitrogen-doped carbon quantum dots as an efficient label-free optical nanoprobe for in vivo live-cell imaging. J. Photochem. Photobiol. A 2019, 372, 99–107. [Google Scholar] [CrossRef]
  138. Marouzi, S.; Darroudi, M.; Hekmat, A.; Sadri, K.; Oskuee, R.K. One-pot hydrothermal synthesis of carbon quantum dots from Salvia hispanica L. seeds and investigation of their biodistribution, and cytotoxicity effects. J. Environ. Chem. Eng. 2021, 9, 105461. [Google Scholar] [CrossRef]
  139. Liu, S.; Liu, Z.; Li, Q.; Xia, H.; Yang, W.; Wang, R.; Li, Y.; Zhao, H.; Tian, B. Facile synthesis of carbon dots from wheat straw for colorimetric and fluorescent detection of fluoride and cellular imaging. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2021, 246, 118964. [Google Scholar]
  140. Atchudan, R.; Edison, T.N.J.I.; Perumal, S.; Vinodh, R.; Lee, Y.R. Multicolor-emitting carbon dots from Malus floribunda and their interaction with Caenorhabditis elegans. Mater. Lett. 2020, 261, 127153. [Google Scholar] [CrossRef]
  141. Atchudan, R.; Edison, T.N.J.I.; Shanmugam, M.; Perumal, S.; Somanathan, T.; Lee, Y.R. Sustainable synthesis of carbon quantum dots from banana peel waste using hydrothermal process for in vivo bioimaging. Phys. E Low-Dimens. Syst. Nanostruct. 2021, 126, 114417. [Google Scholar] [CrossRef]
  142. Shi, C.; Qi, H.; Ma, R.; Sun, Z.; Xiao, L.; Wei, G.; Huang, Z.; Liu, S.; Li, J.; Dong, M. N, S-self-doped carbon quantum dots from fungus fibers for sensing tetracyclines and for bioimaging cancer cells. Mater. Sci. Eng. C 2019, 105, 110132. [Google Scholar] [CrossRef] [PubMed]
  143. Ghosh, S.; Ghosal, K.; Mohammad, S.A.; Sarkar, K. Dendrimer functionalized carbon quantum dot for selective detection of breast cancer and gene therapy. Chem. Eng. J. 2019, 373, 468–484. [Google Scholar] [CrossRef]
  144. Sahoo, N.K.; Jana, G.C.; Aktara, M.N.; Das, S.; Nayim, S.; Patra, A.; Bhattacharjee, P.; Bhadra, K.; Hossain, M. Carbon dots derived from lychee waste: Application for Fe3+ ions sensing in real water and multicolor cell imaging of skin melanoma cells. Mater. Sci. Eng. C 2020, 108, 110429. [Google Scholar] [CrossRef] [PubMed]
  145. Tadesse, A.; Hagos, M.; RamaDevi, D.; Basavaiah, K.; Belachew, N. Fluorescent-nitrogen-doped carbon quantum dots derived from citrus lemon juice: Green synthesis, mercury (II) ion sensing, and live cell imaging. ACS omega 2020, 5, 3889–3898. [Google Scholar]
  146. D’souza, S.L.; Chettiar, S.S.; Koduru, J.R.; Kailasa, S.K. Synthesis of fluorescent carbon dots using Daucus carota subsp. sativus roots for mitomycin drug delivery. Optik 2018, 158, 893–900. [Google Scholar] [CrossRef]
  147. Babu, P.J.; Saranya, S.; Singh, Y.D.; Venkataswamy, M.; Raichur, A.M.; Doble, M. Photoluminescence carbon nano dots for the conductivity based optical sensing of dopamine and bioimaging applications. Opt. Mater. 2021, 117, 111120. [Google Scholar] [CrossRef]
  148. Vandarkuzhali, S.A.A.; Natarajan, S.; Jeyabalan, S.; Sivaraman, G.; Singaravadivel, S.; Muthusubramanian, S.; Viswanathan, B. Pineapple peel-derived carbon dots: Applications as sensor, molecular keypad lock, and memory device. ACS Omega 2018, 3, 12584–12592. [Google Scholar]
  149. Raji, K.; Ramanan, V.; Ramamurthy, P. Facile and green synthesis of highly fluorescent nitrogen-doped carbon dots from jackfruit seeds and its applications towards the fluorimetric detection of Au3+ ions in aqueous medium and in in vitro multicolor cell imaging. New J. Chem. 2019, 43, 11710–11719. [Google Scholar] [CrossRef]
  150. Kumari, S.; Sharma, N.; Sahi, S.V. Advances in Cancer Therapeutics: Conventional Thermal Therapy to Nanotechnology-Based Photothermal Therapy. Pharmaceutics 2021, 13, 1174. [Google Scholar] [CrossRef]
  151. Molaei, M.J. Principles, mechanisms, and application of carbon quantum dots in sensors: A review. Anal. Methods 2020, 12, 1266–1287. [Google Scholar] [CrossRef]
  152. Xu, Q.; Zhou, Q.; Hua, Z.; Xue, Q.; Zhang, C.; Wang, X.; Pan, D.; Xiao, M. Single-Particle Spectroscopic Measurements of Fluorescent Graphene Quantum Dots. ACS Nano 2013, 7, 10654–10661. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Wang, L.L.; Hu, C.; Shao, L.Q. The antimicrobial activity of nanoparticles: Present situation and prospects for the future. Int. J. Nanomed. 2017, 12, 1227–1249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Dong, X.L.; Liang, W.X.; Meziani, M.J.; Sun, Y.P.; Yang, L.J. Carbon Dots as Potent Antimicrobial Agents. Theranostics 2020, 10, 671–686. [Google Scholar] [CrossRef] [PubMed]
  155. Ghirardello, M.; Ramos-Soriano, J.; Galan, M.C. Carbon Dots as an Emergent Class of Antimicrobial Agents. Nanomaterials 2021, 11, 1877. [Google Scholar] [CrossRef] [PubMed]
  156. Li, Y.B.; Bai, G.X.; Zeng, S.J.; Hao, J.H. Theranostic Carbon Dots with Innovative NIR-II Emission for in Vivo Renal-Excreted Optical Imaging and Photothermal Therapy. ACS Appl. Mater. Inter. 2019, 11, 4737–4744. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The main application research directions of CDs. This review focuses on summarizing the biological applications of CDs, such as cell imaging, photocatalysis, optoelectronic devices, molecules sensors, electrocatalysis, and biomedicines, comprehensively. Finally, current challenges, research emphasis, and prospects of this field are also discussed.
Figure 1. The main application research directions of CDs. This review focuses on summarizing the biological applications of CDs, such as cell imaging, photocatalysis, optoelectronic devices, molecules sensors, electrocatalysis, and biomedicines, comprehensively. Finally, current challenges, research emphasis, and prospects of this field are also discussed.
Pharmaceutics 13 01874 g001
Figure 2. TXT assay study on human dermal fibroblasts. Adapted from [129], MDPI, 2019.
Figure 2. TXT assay study on human dermal fibroblasts. Adapted from [129], MDPI, 2019.
Pharmaceutics 13 01874 g002
Figure 3. (a) Toxicity assay of nematode incubation under different concentrations of synthesized CQDs. Multicolor imaging of in vivo model nematode incubated with CQDs (100 μg/mL) under the excitation wavelengths of (b) 400 nm, (c) 470 nm, (d) 550 nm, (e) bright-field (BF), and (f) merge (overlap). Live nematodes were immobilized using 0.05% sodium azide (NaN3) for imaging under fluorescence filters. Adapted with permission from [141], Elsevier, 2021.
Figure 3. (a) Toxicity assay of nematode incubation under different concentrations of synthesized CQDs. Multicolor imaging of in vivo model nematode incubated with CQDs (100 μg/mL) under the excitation wavelengths of (b) 400 nm, (c) 470 nm, (d) 550 nm, (e) bright-field (BF), and (f) merge (overlap). Live nematodes were immobilized using 0.05% sodium azide (NaN3) for imaging under fluorescence filters. Adapted with permission from [141], Elsevier, 2021.
Pharmaceutics 13 01874 g003
Figure 4. C. elegans confocal imaging excited under the wavelengths of (a) 400 nm, (b) 470 nm, and (c) 550 nm; as well as (d) bight field; (e) the merged image; and (f) cell viability test under different concentrations of NCDs, adapted with permission from [96], ACS Publications, 2018. Bright-field (g) wild-type C. elegans imaging and (i) 300-CDs treated C. elegans imaging; UV-exposed (h) wild-type C. elegans imaging and (j) 300-CDs treated C. elegans imaging, adapted with permission from [122], Elsevier, 2019.
Figure 4. C. elegans confocal imaging excited under the wavelengths of (a) 400 nm, (b) 470 nm, and (c) 550 nm; as well as (d) bight field; (e) the merged image; and (f) cell viability test under different concentrations of NCDs, adapted with permission from [96], ACS Publications, 2018. Bright-field (g) wild-type C. elegans imaging and (i) 300-CDs treated C. elegans imaging; UV-exposed (h) wild-type C. elegans imaging and (j) 300-CDs treated C. elegans imaging, adapted with permission from [122], Elsevier, 2019.
Pharmaceutics 13 01874 g004
Figure 5. In vivo images and accumulation of CDs: (a) in vivo images of R-CDs, adapted with permission from [22], Royal Society of Chemistry, 2017; (b) metabolism of CPDs, adapted with permission from [23], Wiley, 2020; (c) subcutaneous and intravenous injection of NIR-CDs, adapted with permission from [26], Elsevier, 2019; and (d) metabolism of Fn-CDs, adapted with permission from [28], Royal Society of Chemistry, 2021.
Figure 5. In vivo images and accumulation of CDs: (a) in vivo images of R-CDs, adapted with permission from [22], Royal Society of Chemistry, 2017; (b) metabolism of CPDs, adapted with permission from [23], Wiley, 2020; (c) subcutaneous and intravenous injection of NIR-CDs, adapted with permission from [26], Elsevier, 2019; and (d) metabolism of Fn-CDs, adapted with permission from [28], Royal Society of Chemistry, 2021.
Pharmaceutics 13 01874 g005
Figure 6. Confocal images collected at 405 nm of (a) HeLa cell; (c) HepG2 cell incubated with CDs; adding Fe3+ and CDs to (b) HeLa cell; and (d) HepG2 cell, adapted with permission from [33], Elsevier, 2017. Confocal laser images of T24 cells excited at 405 nm and 488 nm, (e) treated with B-NCdots and (f) treated with Cu2+ and B-NCdots, adapted with permission from [117], Springer Nature, 2019.
Figure 6. Confocal images collected at 405 nm of (a) HeLa cell; (c) HepG2 cell incubated with CDs; adding Fe3+ and CDs to (b) HeLa cell; and (d) HepG2 cell, adapted with permission from [33], Elsevier, 2017. Confocal laser images of T24 cells excited at 405 nm and 488 nm, (e) treated with B-NCdots and (f) treated with Cu2+ and B-NCdots, adapted with permission from [117], Springer Nature, 2019.
Pharmaceutics 13 01874 g006
Figure 7. Confocal laser images of: (a) L929 cells, (c) HeLa cells, and (e) HepG2 cells in 0.5 mM GSH and 150 μg/mL wCDs; and (b) L929 cells, (d) HeLa cells, (f) and HepG2 cells in 1 mM GSH and 150 μg/mL wCDs, adapted with permission from [24], Royal Society of Chemistry, 2021.
Figure 7. Confocal laser images of: (a) L929 cells, (c) HeLa cells, and (e) HepG2 cells in 0.5 mM GSH and 150 μg/mL wCDs; and (b) L929 cells, (d) HeLa cells, (f) and HepG2 cells in 1 mM GSH and 150 μg/mL wCDs, adapted with permission from [24], Royal Society of Chemistry, 2021.
Pharmaceutics 13 01874 g007
Figure 8. (a) Linear calibration of GSH probing, adapted with permission from [107], Royal Society of Chemistry, 2012. (b) The fluorescence spectra of PCQDs with varying DOX concentrations, adapted with permission from [27], Elsevier, 2021. (c) Confocal images of HeLa cells treated with MCDs, (d) adding 10 μM dexamethasone or (e) 100 μM dexamethasone, at excitation wavelengths of (c1,d1,e1) 405 nm, (c2,d2,e2) 488 nm, and (c3,d3,e3) bright-field, adapted with permission from [66], Royal Society of Chemistry, 2016.
Figure 8. (a) Linear calibration of GSH probing, adapted with permission from [107], Royal Society of Chemistry, 2012. (b) The fluorescence spectra of PCQDs with varying DOX concentrations, adapted with permission from [27], Elsevier, 2021. (c) Confocal images of HeLa cells treated with MCDs, (d) adding 10 μM dexamethasone or (e) 100 μM dexamethasone, at excitation wavelengths of (c1,d1,e1) 405 nm, (c2,d2,e2) 488 nm, and (c3,d3,e3) bright-field, adapted with permission from [66], Royal Society of Chemistry, 2016.
Pharmaceutics 13 01874 g008
Figure 9. SEM images without ACDs of (a) E. coli, (e) S. aureus, and ACDs-treated (c) E. coli, and (g) S. aureus. Magnified SEM images in the red square (b) E. coli, (f) S. aureus; and ACDs-treated (d) E. coli, and (h) S. aureus, adapted with permission from [123], Royal Society of Chemistry, 2020. SEM images of (i) E. coli and S. aureus untreated and treated with 75 μg/mL and 50 μg/mL of s-CGCDs, m-CGCDs, and l-CGCDs in Luria–Bertani broth medium for 6 h, adapted with permission from [76], Elsevier, 2021.
Figure 9. SEM images without ACDs of (a) E. coli, (e) S. aureus, and ACDs-treated (c) E. coli, and (g) S. aureus. Magnified SEM images in the red square (b) E. coli, (f) S. aureus; and ACDs-treated (d) E. coli, and (h) S. aureus, adapted with permission from [123], Royal Society of Chemistry, 2020. SEM images of (i) E. coli and S. aureus untreated and treated with 75 μg/mL and 50 μg/mL of s-CGCDs, m-CGCDs, and l-CGCDs in Luria–Bertani broth medium for 6 h, adapted with permission from [76], Elsevier, 2021.
Pharmaceutics 13 01874 g009
Figure 10. Graphs of E. coli and S. aureus incubated with varying concentrations of OCDs, TCDs, and MCDs for 24 h; alive ratio of E. coli and S. aureus calculated using UV–Vis spectroscopy methods, adapted with permission from [79], Elsevier, 2020.
Figure 10. Graphs of E. coli and S. aureus incubated with varying concentrations of OCDs, TCDs, and MCDs for 24 h; alive ratio of E. coli and S. aureus calculated using UV–Vis spectroscopy methods, adapted with permission from [79], Elsevier, 2020.
Pharmaceutics 13 01874 g010
Figure 11. (a) Schematic diagram of photothermal therapy of cancer in vivo; (b) infrared thermal images of tumor-bearing mice with intravenous (i.v.) or intratumoral (i.t.) injection of CDs (50 μL, 20 mg/mL), and PBS (50 μL); (c) tumor volume after treatment; and (d,e) images of tumor-bearing mice and harvested tumors after 6 days, adapted with permission from [156], ACS publications, 2019.
Figure 11. (a) Schematic diagram of photothermal therapy of cancer in vivo; (b) infrared thermal images of tumor-bearing mice with intravenous (i.v.) or intratumoral (i.t.) injection of CDs (50 μL, 20 mg/mL), and PBS (50 μL); (c) tumor volume after treatment; and (d,e) images of tumor-bearing mice and harvested tumors after 6 days, adapted with permission from [156], ACS publications, 2019.
Pharmaceutics 13 01874 g011
Figure 12. (a) IR thermal images post-injection; (b) temperature increase trends; (c) images of mice with different therapeutic methods; and (d) growth of the tumor during 14 days, adapted with permission from [82], Elsevier, 2018.
Figure 12. (a) IR thermal images post-injection; (b) temperature increase trends; (c) images of mice with different therapeutic methods; and (d) growth of the tumor during 14 days, adapted with permission from [82], Elsevier, 2018.
Pharmaceutics 13 01874 g012
Figure 13. (a) Real-time NIR fluorescence image of mice tumors under NBCD-PEG-Ce6-Tf treatment at different time points; tumor propagating images for (b) NBCD-PEG-Ce6-Tf + laser-treated and (d) control groups; (c) time-dependent tumor growth curves after other treatments, adapted with permission from [126], Royal Society of Chemistry, 2018.
Figure 13. (a) Real-time NIR fluorescence image of mice tumors under NBCD-PEG-Ce6-Tf treatment at different time points; tumor propagating images for (b) NBCD-PEG-Ce6-Tf + laser-treated and (d) control groups; (c) time-dependent tumor growth curves after other treatments, adapted with permission from [126], Royal Society of Chemistry, 2018.
Pharmaceutics 13 01874 g013
Figure 14. SEM images of (a) untreated and (b) treated P. aeruginosa with 30 μg/mL of CDs; (c) DCFDA probing of ROS generation; fluorescent microscopic images of MDA-MB-231 cells (d,f) CDs (3.34 μg/mL) treated and (e,g) untreated with nuclear staining assays (AO and Hoechst 33342), adapted with permission from [80], ACS publications, 2020.
Figure 14. SEM images of (a) untreated and (b) treated P. aeruginosa with 30 μg/mL of CDs; (c) DCFDA probing of ROS generation; fluorescent microscopic images of MDA-MB-231 cells (d,f) CDs (3.34 μg/mL) treated and (e,g) untreated with nuclear staining assays (AO and Hoechst 33342), adapted with permission from [80], ACS publications, 2020.
Pharmaceutics 13 01874 g014
Table 1. Natural CDs and their applications.
Table 1. Natural CDs and their applications.
SourceMethodsApplied Ex./Emit. (nm)QY(%)ApplicationApplied
Concentration
Ref.
Lemon juiceSolvothermal440–600/575–65028%In vivo bioimaging100 μL (20 μg/mL)[22]
Taxus leafSolvothermal380–580/67359%In vivo bioimaging20 mg/kg[23]
Wedelia trilobataSolvothermal370–470/483–520, 654H2O: 10.52%
EtOH: 18.16%
Glutathione sensing and cell imaging50–400 μg/mL[24]
LeekSolvothermal390–450; 325–385/440–500, 676; 440DCD: 1.7%
SCD: 1.14%
Cell imaging, Cu2+ and tetracycline sensing0.5 mg/ mL[25]
Lemon juiceSolvothermal237, 279, 570–670/70431%In vivo bioimaging50 μL (30 μg/mL)[26]
PlumSolvothermal328–418/450–5500.54%Doxorubicin sensing200 μL (1.0 mg/mL)[27]
F. nucleatumHydrothermal300–400/450–4709.9%In vivo bioimaging and Fe3+ sensing10 μL/ g (0.7 mg/mL)[28]
Green pepperHydrothermal310–380/400–4608.7%Fe3+ sensing and cell imaging50 mg/mL[29]
PapayaHydrothermal350–490/445–530H2O: 18.98%
EtOH: 18.39%
Fe3+ sensing and cell imaging100, 175 μL (0.94 mg/mL)
100, 175 μL (1.17 mg/mL)
[30]
P. aviumHydrothermal280–360/411–43013%Fe3+ sensing and cell imaging0–40 μL[31]
HoneyHydrothermal320–410/410–47519.8%Fe3+ sensing and cell imaging40 μL (1 mg/mL)[32]
Sweet potatoHydrothermal300–410/406–4868.64%Fe3+ sensing and cell imaging0–100 μg/mL[33]
Black teaHydrothermal290–420/398–490n/aFe3+ sensing990 μL (8 μg/mL)[34]
Fish-scaleHydrothermal220–390/425–4556.9%Fe3+ sensing5 mg/mL[35]
KiwiHydrothermal300–450/432–50014%/19%Fe3+ sensing0.5 mL[36]
Goose featherHydrothermal300–500/410–56017.1%Fe3+ sensing1 mL[37]
CranberryHydrothermal300–500/410–54010.85%Fe3+ sensingn/a[38]
PotatoHydrothermal323/40515%Fe3+ sensingn/a[39]
Boswellia ovalifoliolata barkHydrothermal275–440/400–53510.2%Fe3+ sensing20 μg/mL[40]
RosinHydrothermal290–370/425–4751.22%Fe3+ sensing and cell imaging1.25–160 μg/mL[41]
Coriander leafHydrothermal320–480/400–5106.48%Fe3+ sensing, cell imaging, and antioxidant0–1.0 mg/mL[42]
Mint leafHydrothermal330–420/410–5007.64%Fe3+ and ascorbic acid sensingn/a[43]
LeekHydrothermal300–460/449–534n/aDDVP sensing and cell imaging0–300 mg/mL[44]
Peach gumHydrothermal330–450/327–50528.46%Au3+ sensing and cell imaging0.5 mL (20 mg/mL)[45]
TomatoHydrothermal367/440n/aCarcinoembryonic antigen and aptamer sensing1 μg/mL[46]
Bean pod and onionHydrothermal310–380/410–4505.55%Ag1+ sensing and cell imaging200 μg/mL[47]
D. SalinaHydrothermal310–400/400–4758%Hg2+ and Cr4+ sensing and cell imaging0–75 μg/mL[48]
Chinese yamHydrothermal280–440/400–5259.3%6-mercaptopurine and Hg2+ sensingn/a[49]
Pomelo peelHydrothermal365/4446.9%Hg2+ sensingn/a[50]
StrawberryHydrothermal344–440/427–5006.3%Hg2+ sensing75 μL[51]
CucumberHydrothermal418–518/514–5713.25%Hg2+ sensingn/a[52]
Highland barleyHydrothermal340–480/450–52514.4%Hg2+ sensing0.05 mg/mL[53]
Lemon peelHydrothermal300–540/441–60514%Cr6+ sensingn/a[54]
Elaeagnus angustifoliaHydrothermal310–410/290–45016.8%Tartrazine sensingn/a[55]
AloeHydrothermal370–480/443–52510.37%Tartrazine sensingn/a[56]
Coconut waterHydrothermal340–450/430–5002.8%Thiamine sensing and cell imagingn/a[57]
LentilHydrothermal310–390/400–46010%Thioridazine hydrochloride sensing200 μL[58]
Pomegranate juiceHydrothermal280–350/350–6004.8%Cephalexin sensing30 μL (1.0 mg/mL)[59]
Bamboo leafHydrothermal365–525/440–5407.1%Cu2+ sensingn/a[60]
Pipe tobaccoHydrothermal310–430/425–5153.2%Cu2+ sensingn/a[61]
Apple juiceHydrothermal300–540/465–5654.27%Cell imaging10 μg/mL[62]
Hylocereus undatusHydrothermal275–380/400–450n/aCell imaging0–50 μL/mL[63]
Saccharum officinarumHydrothermal300–540/450–5505.76%Cell imaging0–400 mg/mL[64]
LinseedHydrothermal350–450/50361%Cell imaging0.04 mg/mL[65]
Shiitake mushroomHydrothermal330–450/410–5005.5%Cell imaging and pH sensing2 mg/mL[66]
CitrusHydrothermal360–500/460–5541.1%Cell imaging30 μL (1.0 mg/mL)[67]
CarrotHydrothermal360–520/442–5655.16%Cell imaging700 μg/mL[68]
Dwarf bananaHydrothermal310–460/395–50523%Cell imaging0–200 μg/mL[69]
BagasseHydrothermal330–510/450–55012.3%Cell imaging and biolabeling100 μg/mL[70]
CabbageHydrothermal276, 320/432–58416.5%Cell imaging100 μL (20–1000 μg/mL)[71]
Alkali ligninHydrothermal280–450/410–51021%Cell imaging0–100 μg/mL[72]
ShrimpHydrothermal360–530/430–55054%Cell imaging and drug delivery10–500 μg/mL[73]
Wheat branHydrothermal360–540/460–60033.23%Cell imaging and drug delivery2 mg/mL[74]
MilkHydrothermal360/450n/aCell imaging and anticancer drug delivery100–600 μg/mL[75]
Chlorhexidine gluconateHydrothermal360–560/480–600s-CGCDs: 2.6%
m-CGCDs:
11.3%
l-CGCDs:
8.0%
Antibacterial and cell imaging0–150 μg/mL[76]
Turmeric leafHydrothermal310–470/429–520n/aAnti-bacterial0–1.0 mg/mL[77]
RosemaryHydrothermal332–422/424–500n/aAnti-bacterial12 μg/mL[78]
Osmanthus leaf, tea leaf, and milk vetchHydrothermal450/530n/aAntibacterial and cell imaging0–1000 μg/mL[79]
MushroomHydrothermal300–500/372–545n/aAnti-bacterial, anti-cancer, and Pb2+ sensing0–25 μg/mL[80]
WatermelonHydrothermal808/900–12000.4%Photothermal therapy and cell imaging0–20 mg/mL[81]
Hypocrella BambusaHydrothermal540–590/600–650n/aPhotodynamic and photothermal therapy0–200 μg/mL[82]
Camellia japonicaHydrothermal360/400–700n/aPhotodynamic and photothermal therapy45 μg/mL[83]
GingerHydrothermal325–445/400–50013.4%Cancer inhibition and cell imaging440 μg[84]
GarlicHydrothermal320–580/380–60017.5%Cell imaging0–1 mg/mL[85]
StarchHydrothermal340–500/452–54521.7%Cell imaging0.078–1.250 mg/mL[86]
Orange juiceHydrothermal360–450/441–51026%Cell imaging0–200 μg/mL[87]
Bee pollenHydrothermal340–450/425–505c-CDs:
8.9%
l-CDs:
6.1%
Cell imaging0.5 mg/mL[88]
GelatinHydrothermal300–500/430–58031.6%Cell imaging5.o mL (0.8 mg/mL)[89]
PapayaHydrothermal300–500/450–5507.0%Cell imaging16.2–500 μg/mL[90]
OatmealHydrothermal280–460/410–50437.4%Cell imaging1 mg/mL[91]
Egg whiteHydrothermal290–450/415–54061%Cell imaging0.04 mg/mL[92]
Corn flourHydrothermal320–500/401–5537.7%Cell imaging and Cu2+ sensing0–640 μg/mL[93]
Humic acidHydrothermal320–520/440–5405.7%Cell imaging0.2 mg/mL[94]
DurianHydrothermal400–560/60579%Cell imaging0–500 μg/mL[95]
GooseberryHydrothermal300–500/406–54513.5%C. elegans bioimaging50 μg/mL[96]
Rice huskHydrothermal310–340/360–4408.1%Cell imaging50 μg/mL[97]
AyurvedicChemical ablation430/518n/aCell imaging and phototherapy0.5 mg/mL[98]
Coffee bean shellChemical ablation280–520/368–557n/aIn vivo bioimaging and antioxidant0–400 μg/mL[99]
MuskmelonChemical ablation342, 415, 425/432, 515, 5547.07%/26.9%/14.3%Hg2+ sensing and Cell imaging0.25–1.00 mg/mL[100]
Cow manureChemical ablation320–450/400–5300.65Cell imaging2.5 mg/mL[101]
Food wasteUltrasound irradiation330–405/400–4702.85%Cell imaging0–4 mg/mL[102]
Citrus limone juiceUltrasound irradiation230–450/325–53812.1%/15%Cell imaging2–100 mM[103]
Crab shellUltrasound irradiation330–390/410–45014.5%Cell imaging0–1000 μg/mL[104]
SilkwormMicrowave300–400/350–55046%Cell imaging0–15 mg/mL[105]
Algal bloomMicrowave300–500/400–55013%Cell imaging10–1000 μg/mL[106]
EggshellMicrowave275/45014%Glutathione sensingn/a[107]
FlourMicrowave360–500/438–5505.4%Hg2+ sensing4 μg/mL[108]
ProteinMicrowave300–420/380–48014%Ag+ sensingn/a[109]
Rose flowerMicrowave330–410/390–43513.45%Tetracycline sensingn/a[110]
Onion peelMicrowave300–470/520n/aSkin wound healing1.5 mg/mL[111]
LycheePyrolysis365/44010.6%Cell imaging0–1000 μg/mL[112]
CoffeePyrolysis350–500/400–6003.8%Cell imaging1.2 mg/mL[113]
UrinePyrolysis275–625/450–65014%Cell imaging0.05–1.5 mg/mL[114]
Watermelon peelPyrolysis310–550/490–5807.1%Cell imagingn/a[115]
Konjac flourPyrolysis400–700/575–76013%Fe3+ and L-lysine sensing and cell imaging200 μg/mL[116]
Soybean and broccoliPyrolysis300–460/425–50012.8%Cu2+ sensing and cell imaging0–300 μg/mL[117]
Borassus flabelliferPyrolysis300–400/350–40311.73%/13.97%/10.83%Fe3+ sensingn/a[118]
Peanut shellPyrolysis262–402/413–50010.58%Cu2+sensingn/a[119]
Assam teaPyrolysis340/446n/aDopamine and ascorbic acid sensingn/a[120]
Peanut shellPyrolysis320–480/441–5249.91%Cell imaging0–1.2 mg/mL[121]
Roast duckPyrolysis300–400/40010.53%/38.05%C. elegans bioimaging15 mg/mL[122]
Artemisia argyi leafPyrolysis360–440/450–480n/aAntibacterial and cell imaging0–150 μg/mL[123]
Sugarcane bagassePyrolysis405/550n/aDrug deliveryn/a[124]
Silkworm cocoonPyrolysis378/4596.32%Anti-inflammatory1.4 mg/mL;[125]
Lychee exocarpPyrolysis365/423n/aDrug delivery and cell imaging0–15 μg/mL[126]
Bamboo leafPyrolysis300–400/425–475n/aCell imaging and anticancer drug delivery0–400 μg/mL[127]
Walnut shellPyrolysis360–540/500–560n/aCell imaging100 μg/mL[128]
Table 3. Modifications of CDs.
Table 3. Modifications of CDs.
CDs TypeModificationGoalRef.
Carbon quantum dotsEthylene diamineNucleoli selection[101]
Nitrogen-doped carbon dotsFolic acidCancer cell targeting[104]
Carbon dotsPolyethylene glycol diamine; chlorin e6; transferrinPhotosensitizing and cancer cell targeting
Carbon dots4-carboxy-benzyl boronic acidTumor cell targeting[127]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chan, M.-H.; Chen, B.-G.; Ngo, L.T.; Huang, W.-T.; Li, C.-H.; Liu, R.-S.; Hsiao, M. Natural Carbon Nanodots: Toxicity Assessment and Theranostic Biological Application. Pharmaceutics 2021, 13, 1874. https://doi.org/10.3390/pharmaceutics13111874

AMA Style

Chan M-H, Chen B-G, Ngo LT, Huang W-T, Li C-H, Liu R-S, Hsiao M. Natural Carbon Nanodots: Toxicity Assessment and Theranostic Biological Application. Pharmaceutics. 2021; 13(11):1874. https://doi.org/10.3390/pharmaceutics13111874

Chicago/Turabian Style

Chan, Ming-Hsien, Bo-Gu Chen, Loan Thi Ngo, Wen-Tse Huang, Chien-Hsiu Li, Ru-Shi Liu, and Michael Hsiao. 2021. "Natural Carbon Nanodots: Toxicity Assessment and Theranostic Biological Application" Pharmaceutics 13, no. 11: 1874. https://doi.org/10.3390/pharmaceutics13111874

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