Next Article in Journal
Mesenchymal Stem Cell Therapy: Therapeutic Opportunities and Challenges for Diabetic Kidney Disease
Previous Article in Journal
Circulating Matrix Metalloproteinases for Prediction of Aortic Dilatation in Children with Bicuspid Aortic Valve: A Single-Center, Observational Study
Previous Article in Special Issue
Nanoparticles-Delivered Circular RNA Strategy as a Novel Antitumor Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Graphene Quantum Dots from Natural Carbon Sources for Drug and Gene Delivery in Cancer Treatment

by
Henrry M. Osorio
1,
Fabián Castillo-Solís
2,
Selena Y. Barragán
1,
Cristina Rodríguez-Pólit
2,3,4 and
Rebeca Gonzalez-Pastor
2,*
1
Departamento de Física, Escuela Politécnica Nacional, Av. Ladrón de Guevara E11-253, Quito 170525, Ecuador
2
Centro de Investigación Biomédica (CENBIO), Facultad de Ciencias de la Salud Eugenio Espejo, Universidad UTE, Quito 170527, Ecuador
3
Escuela de Salud Pública, Universidad San Francisco de Quito USFQ, Quito 170527, Ecuador
4
Centro de Referencia Nacional de Genómica, Secuenciación y Bioinformática, Instituto Nacional de Investigación en Salud Pública “Leopoldo Izquieta Pérez”, Quito 170403, Ecuador
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(19), 10539; https://doi.org/10.3390/ijms251910539
Submission received: 29 August 2024 / Accepted: 10 September 2024 / Published: 30 September 2024
(This article belongs to the Special Issue Nanoparticles in Cancer Therapy)

Abstract

:
Cancer therapy is constantly evolving, with a growing emphasis on targeted and efficient treatment options. In this context, graphene quantum dots (GQDs) have emerged as promising agents for precise drug and gene delivery due to their unique attributes, such as high surface area, photoluminescence, up-conversion photoluminescence, and biocompatibility. GQDs can damage cancer cells and exhibit intrinsic photothermal conversion and singlet oxygen generation efficiency under specific light irradiation, enhancing their effectiveness. They serve as direct therapeutic agents and versatile drug delivery platforms capable of being easily functionalized with various targeting molecules and therapeutic agents. However, challenges such as achieving uniform size and morphology, precise bandgap engineering, and scalability, along with minimizing cytotoxicity and the environmental impact of their production, must be addressed. Additionally, there is a need for a more comprehensive understanding of cellular mechanisms and drug release processes, as well as improved purification methods. Integrating GQDs into existing drug delivery systems enhances the efficacy of traditional treatments, offering more efficient and less invasive options for cancer patients. This review highlights the transformative potential of GQDs in cancer therapy while acknowledging the challenges that researchers must overcome for broader application.

1. Introduction

Cancer remains a primary global health concern and a leading cause of mortality worldwide, with 20 million new cases and 9.7 million deaths reported in 2022 [1]. This disease not only represents a significant challenge to public health but also imposes a substantial socioeconomic burden due to its extensive associated costs [2]. Conventional therapies, including surgery, radiotherapy, and chemotherapy, often fall short due to the complex and heterogeneous nature of tumors, resulting in significant collateral damage to healthy cells. Additional therapeutic strategies have been developed to address these limitations, including hormonal therapy, targeted therapy, immunotherapy, and gene and cell therapy [3]. Gene and cell therapies, in particular, have demonstrated potential in preclinical studies by improving the precision and selectivity of treatments directed at cancer cells [4]. However, the clinical application of these therapies is hindered by challenges such as the inefficient delivery and release of therapeutic agents within the tumor microenvironment [5,6]. Recent advancements in cancer treatment have transformed the landscape, driven by a deeper understanding of cancer’s multifaceted nature, including genomic aberrations, dysregulated cellular pathways, and the tumor microenvironment [7,8]. This insight has paved the way for the development of innovative treatments targeting specific genetic alterations and other molecular features of cancer cells [9]. Approaches such as oncolytic viruses, genetic modification of cancer and immune cells, and siRNA technology have led to numerous clinical trials, some of which are in late-stage development [10]. Nevertheless, the implementation of these advanced therapies poses significant financial challenges, emphasizing the need for cost-effective solutions and equitable access to care, in line with the objectives of SDG 3 to ensure healthy lives and promote well-being for all at all ages [11].
In recent decades, nanomedicine has experienced accelerated growth in the design and synthesis of nanoscale vectors applied to cancer treatment [12,13]. These nanovectors include liposomes, micelles, dendrimers, metal nanoparticles, and carbon-based nanoparticles [14]. Nanoparticles can typically range from 1 to 100 nm, where quantum nature dominates over bulk properties due to the nanoscale dimensions. This results in a high surface-to-volume ratio and quantum effect such as the quantum confinement [15,16], which alters optical properties, including photoemission in semiconductor and carbon-based materials. Thanks to their size, nanoparticles can cross physiological barriers and interact with biomolecules within cells and organs at multiple levels. Also, these nanoscale vectors can remain at the tumor site given their enhanced permeability and retention (EPR) effect [17,18]. Furthermore, their surface chemistry can be tailored to support the attachment of therapeutic agents, including genes, hormones, and drugs. This adaptability also allows for the integration of features like reduced uptake by the reticuloendothelial system and controlled drug delivery, which are essential for designing more specific and effective therapies [19]. Thus, nanoparticles have been utilized to develop new treatments based on their physical properties and to enhance conventional cancer therapies, also serving as multifunctional platforms, enabling theranostic [20] and combined therapies [21].
In this context, carbon is one of the most abundant elements and forms multiple structures [22], including graphite, fullerenes [23], and graphene [24]. Graphene in particular has attracted considerable interest in different areas due to its exceptional thermal, mechanical, optical, and electronic properties, such as high intrinsic mobility, a null effective mass, and a zero-bandgap energy [25]. Furthermore, as observed for other nanoparticles, graphene exhibits quantum confinement effects that alter its electronic properties, resulting in a non-zero-bandgap and tunable photoluminescence depending on its size [26]. These atomically flat but zero-dimensional materials based on graphene are known as graphene quantum dots (GQDs). Undoubtedly, this exclusive structure offers unique properties that have been extensively studied in various areas and novel applications such as energy conversion and storage, catalysis, flexible devices, sensing, and display [27,28], as well as in biomedical applications like diagnostics, treatment, bioimaging, and biosensing [20,29,30,31,32,33].
GQDs offer several advantages that make them highly suitable for cancer therapy. Their high surface-to-volume ratio and versatile structure allow for different types of surface functionalization, making GQDs essential platforms for the anchoring and transporting of chemotherapeutic drugs, genes for gene and cell therapy [34], photosensitizers for photodynamic therapy (PDT) [35], heating agents under NIR light for photothermal therapy (PTT) [36], and different cancer-targeting ligands [37]. Their inherent carbon structure and specialized surface functionalization ensure high biocompatibility, low toxicity, and biodegradability, further enhancing their suitability for biomedical applications [38,39]. The biodegradable nature of GQDs aligns with SDG 12, as it underscores the importance of responsible production and consumption, reducing the environmental impact of nanomaterial-based cancer treatments [11]. Furthermore, their photoluminescence enables precise and controlled monitoring of the transported therapeutic agents [40]. These properties collectively contribute to the development of combined therapies using GQDs, leveraging their synergistic effects to enhance therapeutic efficiency.
Despite their advantageous properties, achieving precise properties is challenging due to the sensitivity of these properties to synthesis conditions. Various methods exist for the synthesis of GQDs, each with its own advantages and limitations [41,42]. To unlock the full potential of GQDs for diverse applications, it is crucial to develop novel synthesis strategies that address existing challenges, such as achieving uniform size and morphology, precise bandgap engineering, scalability, minimizing contaminants, and mitigating the environmental impact involving the synthesis reagents and carbon sources [43,44,45]. Such approaches are essential to advancing sustainable nanotechnology solutions, in line with SDG 9 (Industry, Innovation, and Infrastructure) [11].
Prompted by the current landscape, this review explores the recent advances in developing nano-formulations based on GQDs for cancer treatment. We discuss novel methodologies for obtaining GQDs and the unique properties that make them suitable for cancer therapy. Special emphasis is placed on the use of GQDs for drug and gene delivery for cancer treatment, examining their functionalization, efficiency, biocompatibility, and toxicity. Figure 1 highlights the main properties related to the applications of GQDs in cancer therapy.

2. Production of GQDs from Natural Carbon Sources

2.1. Methods of GQDs Production from Natural Carbon Sources

The preparation of GQDs generally falls into two categories: top-down and bottom-up methodologies. Top-down approaches involve utilizing large carbon materials that have large sp2 carbon domains. These materials are then subjected to processes that reduce their size and produce GQDs by cutting down the structure and obtaining small pieces of material, while preserving the sp2 structure in the inner part. On the other hand, bottom-up methodologies utilize carbon-based small molecules to fuse them, constructing the sp2 domains.
Graphite represents the natural material to fabricate GQDs through top-down methodologies. Graphite can be cut down to obtain GQDs using electrochemical and liquid-phase exfoliation techniques. In the electrochemical approach, GQDs are obtained by electrolysis using a two-electrode configuration. A graphite electrode is used as a working electrode together with an appropriate counter electrode, generally platinum. Both electrodes are immersed in an electrolyte solution, which is often used to control the functionalization of GQDs. Applying a voltage between electrodes, the graphene electrode is exfoliated, obtaining pieces of different sizes. Then, the solution is either centrifuged or filtrated to select the smallest pieces, the GQDs. Finally, the solution is subjected to dialysis or distillation to purify the GQDs [46]. In the liquid-phase exfoliation technique, graphite powder is dispersed in a suitable solvent and is complemented by ultrasonic agitation to cut down the layered structure. During ultrasonication, shear forces and cavitation induce exfoliation of the layered structure into small particles. The success of the methodology depends on the solvent used. Solvents with a surface tension around 40 mJ·m−2 are the best solvents, such as N-methyl-2-pyrrolidone (NMP) and N,N-dimethylformamide (DMF). At the end of the process, the resulting dispersion has to be processed to separate the GQDs from larger particles by centrifugation, filtration, and/or dialysis [47,48].
While electrochemical and liquid-phase exfoliation techniques applied to graphite allow the production of GQDs, the largest amount of exfoliated material is graphene, monolayered or multilayered, and non-exfoliated graphite. This results from exfoliation techniques overcoming the van der Waals attractions between adjacent layers of graphene with graphite, but the rupture of C-C bonds cannot be achieved. The higher the sp2 content in the structure, the greater the crystallinity and the fewer the defects, resulting in larger exfoliated particles. To obtain GQDs through top-down methodologies, cutting occurs at the defect sites, and the more defects, the more cutting sites. Therefore, if defects are introduced into the graphite or graphene structure before the exfoliation process, it will be easy to cut down the structure to obtain GQDs. In this context, another technique used to obtain GQDs is the hydrothermal/solvothermal technique. This technique utilizes graphene oxide (GO), which possesses a similar sp2 structure to graphene but with a larger amount of defect sites. GO can be obtained from graphite using the Hummers and modified Hummers methods [49,50,51]. The hydrothermal/solvothermal method consists of subjecting GO to high temperatures and pressures using an autoclave, which cuts down the structure through the defect sites, forming GQDs [52,53,54]. The obtained GQDs particle size can be modified by adjusting the hydrothermal/solvothermal temperature. An increase in temperature reduces the particle size of GQDs [55]. Additionally, reactants can be included in the autoclave to control the functionalization of GQDs. This methodology has also been used to cut down other graphene-based nanomaterials, such as carbon nanotubes [56] and fullerenes [57], to obtain GQDs. Another routine to cut down GO to obtain GQDs is using chemical methods. Here, strong acids or oxidizing agents fragment GO through their functional groups. The sizes of the obtained GQDs can be modified by varying the reaction temperature [58].
For their part, bottom-up methods for the preparation of GQDs utilize carbon-based small molecules such as amino acids, carbohydrates, citric acid, sucrose, glucose, pyrene, and trisodium citrate, among others, as carbon sources [59]. Here, molecules are fused to generate GQDs. In the first instance, the temperature of these molecules can be elevated until the melting point, which leads to the fusion of the molecules forming the GQDs. The increase in the temperature can be achieved using carbonization pyrolysis, hydrothermal/solvothermal method, or by explosion [60,61,62,63]. Controlling the reaction parameters leads to controlling the size of the GQDs [64]. Also, carbon-based molecules are combined with reagents in order to functionalize the GQDs [65]. Solution chemistry [66,67] and microwave irradiation [68] also appear as important routes to fuse carbon-based molecules to produce GQDs. The versatility of bottom-up methods makes them suitable for controlling the size of GQDs, but requires the use of ligands and several purification steps, especially at the end of each reaction stage.

2.2. Novel Approaches and Innovations in GQDs Production

While standard methods described previously allow the preparation of GQDs at the laboratory level, there are several challenges that must be overcome before GQDs can be implemented in the market, especially for cancer treatment. Both top-down and bottom-up methodologies use toxic reagents, strong acids, or potent oxidizing agents, which must be treated after the production process. Also, it is important to consider the reagents used in the pre-fabrication processes, such as the production of GO. Top-down methodologies align more closely with mass production for market implementation, but often result in GQDs with poor control over size and morphology, resulting in a wide variation in their properties. In contrast, bottom-up methodologies offer better control over the size and morphology of GQDs. However, these methods produce smaller quantities of GQDs, which limits their market applicability. We present the improvements developed in the last years to overcome the challenges associated with GQDs’ production.
Related to the electrochemical exfoliation to cut down graphene-based materials into GQDs, this methodology requires a solid graphite working electrode in principle. However, materials with large sp2 carbon domains that are easier to cut down, such as graphene or GO flakes, could be used if pressed or drop-casted onto a conductive material to form an electrode [69,70]. In addition, pre-treatment processes applied to the starting material can be implemented to create defects in the morphology, increasing the cutting sites and, consequently, the efficiency of the GQDs production. For example, pre-heating of graphite rods at 1050 °C before electrochemical exfoliation produces defects on the surface of graphite, which facilitates the cut-down and production of GQDs [71]. In the same way, graphite can also be treated before applying liquid-phase exfoliation techniques to create defects that facilitate the cutting down of the structure. Thus, an innovative pre-treatment that consists of immersing the precursor material into liquid nitrogen up to 10 h before the liquid-phase exfoliation has been proposed [72]. Also, defects can be introduced using chemical reagents. For example, acid treatment applied to GO can be used to facilitate the production of GQDs in hydrothermal/solvothermal methods [73]. In parallel, there has been growing attention to promoting better control of the properties of the obtained GQDs using top-down methods. Thus, it has been demonstrated that surface passivation can be controlled by incorporating glucose in the electrolytic solution and applying microplasmas during the electrochemical exfoliation process. This control allows the formation of excitation-independent GQDs using high surface passivation by oxygen [74].
Researchers have also proposed using innovative reagents that promote an efficient and/or eco-friendly cut-down of graphite structures. In this way, ionic liquid 1-butyl-3-methylimidazolium hexafluorophosphate (BMIMPF6) has been used satisfactorily as electrolyte to cut down self-supported foam-like 3D graphene to obtain GQDs using electrochemical exfoliation [75]. Likewise, supercritical CO2 has proved to be an efficient solvent for obtaining GQDs using liquid-phase graphite exfoliation [76,77]. In the same way, it has been demonstrated that a mixture of ethanol/water can also be used as an eco-friendly solvent to obtain GQDs in the liquid-phase exfoliation of graphene [78]. In the case of using chemical methods to cut down graphene structure, the use of green oxidants to replace strong acids has been extensively studied. For example, H2O2 has been employed as an oxidant to produce GQDs from graphene, obtaining only H2O and GQDs as products [79]. In addition, an oxidant combined with the sonochemical strategy and microwave radiation has been used satisfactorily to increase the efficiency of producing GQDs [80].
On the other hand, bottom-up methodologies can be improved by implementing processes that increase the quantity of produced GQDs, involve lower reaction temperatures, and perform an efficient route to achieve atomic-level accuracy. It has been probed that the use of microwave irradiation heating induces the decomposition of molecules, facilitating their assembly in the pyrolysis and hydrothermal/solvothermal methods [68,81,82]. In addition, novel solution chemistry processes, e.g., catalytic solution chemistry, can be implemented in order to induce chemical changes in molecules to accomplish their fusion to obtain GQDs [59]. Also, the synthesis and use of templates further complement the bottom-up methodologies, especially in order to obtain GQDs with high accuracy in their morphology. Thus, polycyclic aromatic hydrocarbons (PAHs), hexaborylated [6]cyclo-m-phenylene, and polyethylene glycol have been satisfactorily used to obtain efficient templates in bottom-up synthesis of GQDs [83,84,85].
It is also worth mentioning that bottom-up methods have also been exploited to produce graphene or graphitic structures from molecules such as PAHs or cysteine [86,87,88,89]. These artificial graphene and graphite have also been demonstrated to be suitable for cutting down using exfoliation or hydrothermal/solvothermal methods to obtain GQDs with well-defined morphology, diameter, and thickness [89,90]. Figure 2 illustrates the standard methods of GQDs production and their innovations.

2.3. Advantages and Limitations of Natural Sources

Many carbon materials have been utilized in the production of GQDs. Since the characteristics of the resulting product, and consequently, its suitability for future applications, can vary depending on the process and precursors, the selection of the source is equally as important as the selection of the fabrication method [91]. As discussed above, top-down methods are well-suited for the mass production of GQDs, with carbon materials containing large sp² carbon domains being ideal sources for GQD preparation. However, this strategy requires particular instruments and equipment, such as high-pressure and high-temperature equipment, resulting in low yields, limited production scalability, and increased production cost. Additionally, while using graphene, GO, carbon nanotubes, or fullerenes as starting materials increases the performance in the production of GQDs, producing these materials before the GQDs preparation also represents an increase in the production cost. To reduce production costs, carbon-based abundant natural resources have been proposed as sources for GQDs production, applying especially but not limited to top-down methodologies.
Coal and coke consist of small graphite-like clusters arranged in disordered structures with varying degrees of graphitization. Coke, produced through high-temperature thermal treatment of anthracite coal, has a higher degree of graphitization compared to coal. Due to their disorder and defects, both coal and coke are easier to cut down into GQDs than graphite [92]. Chemical methods are commonly employed to achieve this transformation. As the degree of graphitization increases, as seen with anthracite, the size of the produced GQDs shows a wider dispersion, ranging from 1 to 7 nm. Conversely, a lower degree of graphitization, such as in bituminous coal, results in a narrower GQD size distribution, from 1 to 3 nm [93]. However, using coke with a higher degree of graphitization leads to a significant increase in the size dispersion of the produced GQDs, ranging from 40 to 100 nm [94]. These findings suggest that controlling the graphitization degree in the source material is crucial for achieving better control over the size of the GQDs produced. Because of the compromise between the graphitization degree and the dispersion in the size of the GQDs, it is important to use a natural source with a convenient degree of graphitization that is not too high or too low. Considering that chemical exfoliation uses strong acids or oxidizing agents, other top-down procedures should be implemented to obtain GQDs from coal and coke. In this context, electrochemical exfoliation has also been used to satisfactorily obtain GQDs from coke. Remarkably, the right electrolyte selection conduces to the production of GQDs with low size dispersion [95].
Another source to produce GQDs is biomass-carbonaceous waste. Biomass-based GQDs and other carbon-based quantum dots have gained attention due to their biocompatibility and the abundance of biomass as a renewable and sustainable resource [96]. Biomass sources such as plants and animal derivatives, as well as municipal waste, can be used to prepare GQDs [46,97]. In this case, GQDs can be obtained using both top-down and bottom-up methodologies. To use a top-down strategy, biomass is carbonized to promote the formation of large sp2 carbon domains that are subsequently cut down [98]. On the other hand, to use a bottom-up strategy, biomass is treated to separate it into small molecules that can be fussed using pyrolysis, chemical, or hydrothermal/solvothermal routes to obtain GQDs [99,100]. Lastly, amino acids, the building blocks of proteins, have been employed as natural sources for GQD preparation via pyrolysis, a bottom-up approach [101,102]. The use of microwaves for uniform heating during this process enhances size control of the GQDs and provides the added benefits of good biocompatibility and low toxicity [103]. It has been demonstrated that byproducts of the food supply chain can be reused to produce GQDs. These byproducts contain different natural carbon-based molecules which make them a diverse source for GQDs production. Thus, natural carbon sources as wastes derived from vegetables, fruits, livestock, and the food industry contribute to the development of a sustainable GQDs production without contending with food suppliers and without use of any potential toxic chemical reagents [104].
Table 1 presents a compilation of carbon sources, fabrication methods, functionalization, size/thickness, and applications of the GQDs obtained in previous studies.

3. Unique Properties of GQDs for Cancer Therapy

3.1. Morphological, Optical and Electronic Properties

GQDs are composed of sp2-hybridized carbon atoms arranged in a hexagonal lattice, like graphene. This configuration allows for the delocalization of π electrons along the carbon–carbon bonds, forming a conjugated π system [24,25]. In graphene, this electron delocalization provides unique electronic properties, such as high electron mobility and optical transparency [116,117]. The delocalization of electrons within the graphene core of GQDs plays a key role in their unique properties, with quantum confinement effects becoming prominent due to their small size [62,118,119,120]. While quantum confinement leads to discrete energy levels for electrons, the bandgap depends on the final structure of GQDs, which varies based on the synthesis method [121,122] and conditions [27]. This variability leads to differences in size [123,124], shape [125], defects [126], edge structures [127,128,129], functional groups [130,131,132,133], doping [134,135,136,137,138,139,140,141], and conjugation degree [142,143,144]. Theoretical calculations can study GQDs with specific structures [145,146] precise production of identical GQDs with identical morphology, and in consequence, with identical properties, remains a big challenge that needs to be addressed [31,76,147,148].
GQDs exhibit strong absorption in the ultraviolet region, showing the formation of a broad absorption peak, which is attributed to π-π* transitions within the sp2-hybridized carbon structure [52,149]. As the size of the GQDs decreases, the absorption peak tends to blue-shift due to quantum confinement effects [58,150]. Functionalization and doping can also affect the absorption properties of GQDs. In this context, GQDs exhibit a shoulder peak extended to visible and NIR regions in the absorption spectra attributable to n–π* transitions in C=O bonds promoted by introducing oxygen groups [151,152]. In addition, GQDs functionalized with electron-donating groups may exhibit red-shifted absorption peaks, while those functionalized with electron-withdrawing groups may show blue-shifted peaks [57,153,154,155,156,157]. As mentioned before and similar to the absorption changes, quantum confinement causes size-dependent photoluminescence, where smaller GQDs exhibit a larger bandgap, and thus blue-shifted emission [58,150,158,159]. Additionally, their photoluminescence is often excitation wavelength-dependent, with emission characteristics varying according to the excitation light wavelength [160,161,162] and the efficiency of light emission (quantum yield) is typically high [163,164,165,166]. Excitation dependent emission from GQDs is one of the key properties that represents a substantial body of research. The main proposal to explain this behavior is based on the idea of selective excitation of different sizes corresponding to different dominant bandgaps [52,90,143]. However, recent studies indicate that excitation dependent emission depends on surface states, which involve energy transfer or charge transfer mechanisms [167,168]. Simultaneously, other studies have observed excitation independent photoluminescence emission from GQDs [169,170,171,172]. The photoluminescent properties of GQDs are of significant interest due to their potential applications in various fields, including bioimaging, sensing, and optoelectronics [26,27,28,29,30,31,173].
While GQDs are widely recognized for their photoluminescence [121], they also exhibit other crucial properties for cancer therapy. Conjugated π system and discrete energy levels formed through quantum confinement allow the use of GQDs for PTT and PDT. In these therapies, GQDs are delivered to cancer cells and then irradiated. In the case of PTT, cancer cells are irradiated with NIR light. GQDs absorb NIR radiation, which is efficiently converted into heat, producing localized cancer cell death while minimizing damage to healthy tissue [36,174]. Their high photothermal conversion is attributed to substantial optical absorption in the NIR directly related to the degree of π conjugation and low-energy vibrational modes. When π conjugation is disrupted, for example through oxidation of the structure, the optical absorption decreases in the visible and infrared [175]. Thermodynamic vibrations, nonradiative relaxation, and plasmonic localized heating are the three different mechanisms that constitute the photothermal conversion process at GQDs [176]. Thus, as long as a high π conjugated degree is maintained, top-down and bottom-up-synthesized GQDs can be successfully used as PTT agents [174]. In the case of PDT, cancer cells are irradiated with visible light. GQDs absorb visible radiation, and this energy is used to produce reactive oxygen species (ROS), specifically singlet oxygen (1O2), which induce cytotoxicity in target cells [177,178,179]. GQDs present higher 1O2 generation yield compared to conventional photosensitizers as protoporphyrin IX due to the multi-state sensitization mechanism (MSS) promoted by GQDs. In conventional PDT, when PDT agents absorb energy, these go to their respective excited singlet state (S1). Subsequently, an intersystem crossing transition is produced and agents go to their excited triplet state (T1). Finally, agents return to their respective ground states (S0) by transferring energy to molecular oxygen in triplet ground state (3O2). As a result, 1O2 is generated. In MSS, GQDs absorb energy and present two times 1O2 generation yield. First, during the intersystem crossing transition in GQDs (S1 → T1); and second, when GQDs return to their respective ground states (T1 → S0). This MMS is promoted because GQDs present an energy gap between S1 and T1 larger than the formation energy of 1O2, contrary to conventional photosensitizers [121]. This property has been observed specially in GQDs with tunable oxygen functional groups prepared by electrochemical exfoliation or hydrothermal method varying the alkali hydroxide concentration in the used solution [177,179].
While it has been probed that GQDs irradiated with visible light generate ROS and kill cells by causing oxidative stress, when they are not irradiated they can present high antioxidant activity. Both the total oxygen content and the type of oxygen groups contribute to the antioxidant activity of GQDs. The principal mechanism for free radical scavenging is attributed to the hydrogen donor from the surface produced specially by hydroxy and carbonyl moieties introduced when GQDs are prepared under NaOH or H2O2 electrolysis environment [180,181,182].
The conjugated π system promoted by the hexagonal structure of GQDs facilitates the self-assembly of different compounds onto the surface of the GQDs through non-covalent π-π interactions. Additionally, GQDs possess an extremely high surface-to-volume ratio due to their small size and planar structure [183,184]. This combination of characteristics is highly advantageous for applications requiring extensive interaction sites and increased reactivity, such as in catalysis, sensing, and energy storage [185,186,187,188,189,190,191]. In cancer treatment, GQDs serve as effective carriers for drug and gene delivery [34,37,192]. In addition, the heat generated by GQDs under light irradiation can trigger the release of the drug at the target site, enhancing delivery efficiency [193,194,195]. This versatility positions GQDs as a powerful platform for developing combinatorial cancer therapies that may achieve additive or synergistic therapeutic effects, improving treatment efficacy towards malignant tissue while reducing toxicity over healthy cells [36,196,197]. Additionally, GQDs’ photoluminescence enables their use as imaging agents in cancer diagnosis [198], in theranostics that combine diagnosis with treatment [20,192,199], and for real-time monitoring of therapeutic agent delivery, verifying targeting accuracy and assessing treatment effectiveness [177,200,201,202,203].
A relatively new observed property of GQDs promoted by the quantum confinement is the up-conversion photoluminescence. This is referred to as a process where low-energy photons (e.g., infrared light) are absorbed and converted into higher-energy photons (e.g., visible or ultraviolet light). This process involves multiple photon absorption events followed by energy transfer processes within the GQDs [204,205,206]. Applications of up-conversion photoluminescence in GQDs include bioimaging, where infrared light penetrates deeper into tissues and is absorbed by the GQDs, which then convert it into visible or ultraviolet light for capturing high-resolution images [207,208]. These properties make GQDs with up-conversion capabilities attractive for use in various imaging and sensing applications [209,210,211].

3.2. Drug and Gene Loading and Release Properties

In cancer therapy, GQDs have been used to transport drugs, genes, and other molecules with therapeutic activity [34]. Their extensive surface area, variable number of graphene layers, the presence of π electrons, and the ability to be doped with heteroatoms facilitate the formation of plentiful polar groups by providing abundant reactive sites, influencing their overall chemical reactivity and interaction with other molecules or materials [34,40]. These polar groups facilitate the binding of GQDs to aromatic molecules through various interactions, mainly π-π interactions, hydrogen bridges, and electrostatic interactions [212,213,214]. These characteristics enable the incorporation of hydrophobic chemotherapeutics on their surface, the production of more stable complexes, and increased solubility in various solvents [215]. The sp2 nature of the surface [216] and its double face allow GQDs to incorporate molecules forming a sandwich-like structure through π-π interactions [217]. Under physiological pH conditions (7.4), the hydroxyl groups on the GQDs ionize to form oxide ions (-O), which improves the loading capacity of molecules such as doxorubicin (DOX), methotrexate (MTX), and mitoxantrone (MTN) [113,121,217,218]. Furthermore, these interactions can be stabilized by increasing the presence of hydroxyl groups and incorporating other functionalities on the surface of GQDs, such as carboxyl groups. These groups enable electrostatic interactions with positively charged molecules, such as DOX, which contains amino groups [217], and cisplatin (CDDP), which in aqueous solutions forms positively charged dechlorinated species [219]. Additionally, doping generally increases the adsorption energy of drugs onto GQDs, making the formation of complexes thermodynamically favorable [149,220]. On the other hand, the presence of carboxyl groups on the surface of the GQDs facilitated the covalent bonding of CDDP [221] and cytarabine (Cyt) [222], while in other cases, the drugs were covalently linked with polyethylene glycol (PEG) or cysteamine [115,223,224]. These covalent bonds significantly improve control over drug loading, reduce potential leakage, and mitigate unwanted side effects [203].
The incorporation of nucleic acids into GQDs occurs mainly due to the GQDs’ unsaturated polycyclic structure consisting of hexatomic rings of carbon atoms and the nitrogenous bases of the nucleic acids, which facilitates π-π stacking without influencing the stability of DNA [225,226,227]. Unlike drugs, which require a large number of carboxyl or hydroxyl groups in GQDs for efficient loading, reducing the oxidation state and functionalizing GQDs with cationic molecules like polyethylenimine (PEI) promotes their electrostatic interaction with nucleic acids, which are negatively charged [31,214].
In terms of release, complexes can act in response to the acidic tumor microenvironment, both intracellular and extracellular [218,221,228,229]. Under these pH conditions, the H⁺ ions present in the medium tend to strengthen hydrogen bond interactions with the functional groups -COOH, -OH, and -NH2 present on GQDs, as well as with the -OH and -NH2 groups found in drugs like (DOX) [230]. This mechanism has been observed to disrupt the stabilization of loaded drugs and nucleic acids within GQDs, weakening the interactions between the components and facilitating its release. Similarly, low pH conditions promote the release of CDDP by affecting electrostatic interactions through the protonation of functional groups on GQDs, neutralizing their charge [219]. Moreover, for drugs covalently bonded to GQDs, such as Cyt or CDDP, the acidic environment triggers hydrolysis, leading to the ionization of -COOH groups to -COO, and subsequently to cumulative release into the intracellular environment [222].

3.3. Biocompatibility and Toxicity Considerations

Comprehensive research into the biocompatibility and toxicity of nanomaterials is essential for understanding the potential risks and ensuring their safe application in biomedicine [231,232]. Biocompatibility evaluates how well a material interacts with biological systems without eliciting adverse effects; toxicity, on the other hand, refers to the overall harmful effects a substance may exert on living organisms, while cytotoxicity specifically pertains to the detrimental effects on cells [233]. Unlike common semiconductor quantum dots, which are primarily composed of materials like cadmium telluride, cadmium sulfide, cadmium selenide, and indium arsenide [195], the carbon-centric composition of GQDs renders them highly biocompatible and relatively low in toxic potential, GQDs are predominantly based on carbon, which is essential for their safe use. However, their interaction with biological systems, including potential adverse effects, is determined by their size, shape, and chemistry [39]. These interactions can influence cellular functions and processes, such as cell adhesion, proliferation, differentiation, and gene expression [234,235]. Additionally, dosage, exposure time, and aggregation state further determine the toxicity and biocompatibility of GQDs [39]. GQDs acquired through bottom-up methods or sourced from natural carbon demonstrate remarkable biocompatibility and minimal toxicity and cytotoxicity, surpassing those from top-down approaches [236,237]. Together, these elements are critical in ensuring the safety, efficacy, and bioavailability of the therapeutic agents delivered by GQDs [238].
Over the past decade, extensive research has been conducted on the potential toxicity of graphene-based materials, including GQDs [239,240]. However, different studies show varying levels of cytotoxicity [113,241,242]; this is due to slight variations in the properties of GQDs, which can significantly influence their interaction with the cell membrane, proteins, and DNA, resulting in differences among some studies and complicating the understanding of their biological activity. The size of GQDs is crucial in determining their cytotoxic effects. Both small and large GQDs can enter cells through endocytosis and phagocytosis, potentially causing general cytoskeletal injury and mitochondrial dysfunction, but they exhibit cytotoxicity through different mechanisms [240,243]. Smaller GQDs (<10 nm) can penetrate cell membranes efficiently, potentially causing significant intracellular damage. In contrast, larger GQDs, although less likely to penetrate cells deeply, can disrupt cell membranes by forming pores or extracting lipids, leading to increased cytotoxicity over smaller sizes [244,245].
Surface chemistry is another critical factor governing toxicity. Different functional groups, their position, and their abundance on the surface of GQDs lead to different cytotoxicity traits [39,246,247]. In this sense, studies have reported that some GQDs can induce oxidative DNA damage, specifically causing point mutations, and different levels of autophagy induction that suggest different cellular damage mechanisms [248,249,250]. For instance, amine and amide modifications show no cytotoxicity up to 200 μg/mL [251,252]. Additionally, GQDs with oxygen groups are predictably less toxic than more pristine GQDs, as hydrophilic materials tend to exhibit lower toxicity. In vitro and in vivo experiments demonstrated that GQDs have low toxicity due to their high oxygen content, enhancing solubility and stability, facilitating interactions with biological molecules, and preventing aggregation, which are crucial for their safe use [38]. In this sense, while larger GO sheets show significant toxicity, GQDs are comparatively less toxic [240,253,254]. However, the exact relationship between GQDs’ biocompatibility and oxygen content depends on the specific oxygen-containing groups present. In particular, hydroxylated GQDs were found to be the most cytotoxic, causing significant cell death at concentrations under 100 μg/mL. This effect is thought to be due to increased production of ROS, although the cytotoxicity of functionalized GQDs cannot be solely associated with their ability to produce ROS [245,255,256]. In contrast, carboxylic-modified GQDs showed no significant cytotoxicity, maintaining cell viability above 80% even at concentrations of 500 μg/mL in normal cells [257]. On the other hand, heteroatom-doped GQDs are generally more biocompatible than undoped GQDs, often maintaining high cell viability at the same or higher concentrations [39]. Specifically, nitrogen-doped GQDs are widely studied as a potential cancer treatment due to their high biocompatibility [258,259]. Additionally, GQDs co-doped with nitrogen and boron, sulfur, phosphorus, or neodymium also show positive interaction with biological systems [260,261,262,263,264,265]. Lastly, similar to other nanoparticles, strategies like PEGylation aim to modify the surface of GQDs and enhance their biocompatibility by reducing the generation of ROS, improving biodistribution, and enhancing clearance, even at high concentrations [38,254,266].
GQDs generally show minimal organ accumulation and are rapidly cleared through the kidneys, demonstrating low toxicity even with multiple doses in mouse models [38,249]. Studies have shown that GQDs can be broken down by human enzymes myeloperoxidase and eosinophil peroxidase within a few hours, with noticeable degradation observed after several days of treatment [267]. However, consistent with the in vitro analysis, the impact of surface modifications on biological safety must be considered. For instance, carboxylated GQDs showed very low toxicity and organ damage, as confirmed by blood counts and histological analysis, even after several days; in contrast, aminated GQDs were found to be more toxic [257,268,269]. Additionally, it is necessary to analyze the biosafety of purportedly low-toxic GQDs to ensure an impartial assessment, as some GQDs have exhibited developmental toxicity and genotoxicity in zebrafish models, as well as other forms of cell death related to neurotoxic damage [270,271].
Given the critical role of the immune system, it is of great significance to investigate the response of immune-related cells to the exposure of GQDs. Fasbender et al. discovered that in a leukocyte population that retained 90% viability after exposure to 500 μg/mL, the uptake rate via a diffusion-driven mechanism was notably higher in granulocytes and monocytes than in lymphocytes. Within the lymphoid subgroups, a significant retention was observed in CD34+ cells, NK cells, and B cells compared to cytotoxic T cells and helper T cells [272]. These variations in uptake rate could be attributed to differences in cell size and membrane permeability. In a related study, GQDs induced apoptosis and inflammatory responses in human macrophages, accompanied by expression of characteristic response factors such as interleukin-8 or tumor necrosis factor, depending on the concentration [273].
Future research directions include exploring the long-term effects of GQDs, understanding their interaction mechanisms at the molecular level, and developing safer GQD-based materials for clinical applications. Extensive research is required to optimize these factors and enhance the biocompatibility of GQDs, thereby improving their potential for drug and gene delivery in cancer treatment. Despite these challenges, ongoing research explores innovative approaches to overcome the hurdles and facilitate the clinical application of GQDs [274,275].

4. Applications and Novel Strategies in Cancer Treatment

4.1. Drug Delivery Using GQDs

GQDs have shown promise as standalone chemotherapy agents due to their ability to enter the cell nucleus and interfere with DNA activity alongside other advantageous properties stemming from their small size and aromatic ring structure, disrupting cancer cell growth [276,277]. Recent advancements include modifying GQDs with nucleus-targeting ligands to enhance efficacy directly disrupting the cancer cell nucleus [242]. Furthermore, GQDs have been investigated for their role in altering membrane permeability through interactions with lipid rafts, allowing fast translocation of drugs across lipid membranes depending on phospholipid density [278,279,280].
GQDs not only show promise as therapeutic agents but also excel as versatile platforms for drug delivery. Their high surface area and tunable surface chemistry enable precise control over the concentration of oxygen-containing groups and other functionalities [28,40,281,282], enhancing drug-loading capacity of a wide range of organic and inorganic compounds [230,283], and enabling controlled drug release [22,284]. Additionally, their planar structure allows them to pack drugs via both faces and edges and intercalate small molecules between the graphene layers, further maximizing drug-loading capacity [113]. Poorly soluble drugs and complex macromolecules have been dispersed in GQD nanosystems, enhancing drug stability and solubility, ultimately improving drug bioavailability, minimizing side effects and improving the therapeutic index of drugs [34,285,286] and overall efficacy of chemotherapy [223,287].
Incorporating functional groups and linker molecules on the surface of GQDs allows the binding of the complex components through electrostatic interactions and covalent bonds. This prevents the nonspecific release of the drug, improves cellular uptake, promotes intracellular accumulation, and reduces therapeutic doses by up to 10 times compared to the drug-free versions [203,217,241]. Additionally, doping modulates GQDs’ properties, increasing drug absorption and fluorescence [28,220]. Other promising modifications include incorporating proteins, such as serum albumins, which increase the bioavailability and prolong the half-life of the loaded drug [193,288]. Integration of polymers, like PEI and PEG, improves solubility, dispersion, and biocompatibility of GQDs in aqueous media, in addition to increasing the drug-loading capacity and reducing the interaction with macrophages or other non-specific agents in the blood [224,228,289,290].
While some GQDs have the ability to induce oxidative stress [291], others exhibit antioxidant properties [215]. This duality arises from their unique and variable properties attributed to their intrinsic nature [243], doping [292], or functionalization with photosensitizing molecules [179]. For example, researchers have enhanced the oxidative damage capabilities of GQDs by co-doping them with nitrogen and phosphorus, enabling them to efficiently convert hydrogen peroxide into hydroxyl radicals within cancer cells, a process driven by a synergistic electron effect [261]. Additionally, the radical scavenging abilities of GQDs, as previously discussed in their capacity to neutralize free radicals through hydrogen donation, are crucial for mitigating oxidative stress. This stress is closely associated with tumor growth, metastasis, and drug resistance [273]. By effectively reducing oxidative stress, these antioxidant properties of GQDs could potentially enhance the effectiveness of drug treatments in cancer therapy [274,275,276]. Moreover, GQDs exhibit the ability to damage chemotherapy-resistant cells, such as MCF-7/ADR [277]. GQDs interact with cytosine-rich regions of promoters of drug transporter genes like P-glycoprotein, multidrug resistance protein 1 (MRP1), and breast cancer resistance protein (ABCG2), thereby negatively regulating the expression of multiple multidrug-resistant genes, allowing drugs to remain within the intracellular medium and exert their therapeutic activity [278].
The functionalization of GQDs with ligands targeting overexpressed receptors on cancer cells has been a strategy for targeted drug delivery [293]. For instance, gemcitabine has been delivered through GQDs functionalized with hyaluronic acid targeting the CD-44 receptor in pancreatic cancer cells [288]. Similarly, DOX has been administered to lung cancer cells through functionalization with biotin [294]. This has significantly increased the cytotoxicity of the drug compared to the non-functionalized complex and the drug in its free state. For example, one study demonstrated that CDDP administered to breast cancer cells via GQDs functionalized with an antibody fragment (scFvB10) targeting the epidermal growth factor increased treatment efficacy [219]. Another study showed that DOX delivered through GQDs functionalized with an RGD peptide had favorable treatment results [283,295]. Moreover, research involving folic acid targeting the folate receptor has shown that targeted delivery to breast cancer cells reduces the side effects of tamoxifen in normal cells [224] and can reduce tumor size by up to 99.68% in mouse models when mitoxantrone is administered [218]. Furthermore, the targeting capability of GQDs, combined with their intrinsic fluorescence or additional fluorescent molecules, has enabled their application as theranostic agents for diagnosing and monitoring cancer treatment [290,296,297].
The combination of anticancer drugs prevents drug resistance in oncological therapy, increases efficacy at lower doses, and reduces side effects [298]. This was demonstrated in a study where the large surface area of GQDs was utilized to simultaneously load DOX and CDDP, achieving a reduction of up to 10 times in tumor weight compared to the control group without causing alterations in blood chemistry [216]. Similarly, GQDs also combine their drug administration ability with their capabilities to act as photosensitizers when exposed to light of specific wavelengths [265]. This exposure produces oxidative stress in PDT and hyperthermia in PTT, demonstrating a synergistic therapeutic effect [179,299]. For instance, a study improved the therapeutic effect in pancreatic cancer by combining chemotherapy with DOX, gene therapy, and PTT, increasing the number of cells in the apoptosis state to 74% [300]. Another study demonstrated that silver nanoparticles decorated with GQDs loaded with DOX reduced cell viability from 75% to 68% after irradiation due to the enhanced effect of PDT [289]. Similarly, another research effort used a complex of nitrogen-doped GQDs (N-GQDs) as a sensitizing agent for PDT, DOX as chemotherapy, and 3-Aminopropyltriethoxysilane (APTES) as a charge reversal agent to facilitate drug delivery to the nucleus, significantly enhancing cancer therapy [301]. Alternatively, PTT using GQDs loaded with IR-780, a theranostic agent capable of increasing local temperature and inducing hyperthermia, contributed to the effectiveness of the treatment and nearly eliminated the tumor in a mouse model [302]. A theranostic system of GQDs released its aptamer to hybridize with a cancer biomarker (miRNA-155), allowing the differentiation from healthy cells through fluorescence. This system applied different wavelengths to increase the temperature in the PTT and activate the loaded porphyrins to produce singlet oxygen from molecular oxygen in alveolar carcinoma and breast cancer cell lines, demonstrating that the combination of both therapies is capable of reducing cell viability up to 14% [297].
In recent studies, GQDs act as a supporting element, combining their properties with those of more complex drug delivery systems (DDS) in cancer therapy. Firstly, combining GQDs with delivery systems constituted by polymers, such as chitosan, improves the biocompatibility and intracellular accumulation of the drug [303]. Carboxymethylcellulose provides greater mechanical strength [304], and poly(lactic-co-glycolic) acid (PLGA) helps maintain the complex’s stability by protecting it from oxygen and water [193]. Additionally, GQDs have endowed other nanoparticles with pH-responsive drug release, light sensitivity, monitoring abilities, and fluorescence imaging, fostering the evolution of more intricate strategies [305,306,307,308]. For example, to increase biocompatibility, wrapping the complexes with a red blood cell membrane prevents their interaction with macrophages and extends their circulation time [309]. Similarly, applying GQDs in a dual DDS carrying two distinct therapeutic agents improves treatment effectiveness by enhancing drug loading and release, combating drug resistance, and decreasing toxicity toward healthy cells [197]. The therapeutic effect was improved by subjecting magnetic hollow nanoparticles decorated with DOX-loaded GQDs to a magnetic field, providing mechanical stress and chemotherapeutic damage to synergistically eliminate cancer cells [310].
GQDs used in drug delivery have been predominantly synthesized from precursors such as GO [283,288,311], graphite, [216,283,312], acidic compounds like L-glutamic acid and citric acid, and their respective salts [113,224,307], polymers like polythiophene [313], sugars like glucosamine [241], other nanomaterials such as multi-walled carbon nanotubes [294], and materials like CX-72 carbon black [219].
Although GQDs are a promising alternative for drug delivery, more detailed and comprehensive investigations on their drug-loading and -release properties are necessary [35,284]. Additionally, exploring doping and functionalization mechanisms with previously untested elements and molecules to modify their properties based on their interaction with drugs, cancer cells, and the organism is crucial [31,104,314]. Combining GQDs with other drug delivery systems to develop more effective platforms in cancer therapy could lead to the first clinical trials and products in this field [315]. Moreover, the interaction between biomaterials like GQDs and immune cells plays a vital role in designing innovative and effective drug delivery systems capable of eliciting immune responses [316]. Zhang et al. synthesized a stable hybrid GQD-PEG photosensitizer with photodynamic efficiency and immunostimulatory properties, inducing tumor cell death through ROS generation upon irradiation, followed by immune response activation via inflammatory cytokine secretion, ultimately enhancing therapeutic efficacy [317].
Table 2 summarizes the functionalization strategies and pharmacological interactions of GQDs, highlighting data from both in vivo and in vitro experiments.

4.2. Gene Delivery and Gene Therapy with GQDs

While most research focuses on drug delivery, GQDs display encouraging opportunities in transporting various substances beyond pharmaceuticals, including DNA and peptides [34]. Cancer gene therapy, in particular, significantly relies on efficiently delivering nucleic acids (DNA or RNA) to an individual’s tumor to manipulate the genetic material or modulate gene expression [10]. However, the therapeutic potential of naked nucleic acids is limited by several barriers, such as size and charge constraints, degradation by extracellular nucleases, and entrapment within endosomal compartments. Hence, developing efficient delivery systems is crucial for gene therapy [319]. As such, utilizing the unique properties of GQDs to design specifically tailored platforms becomes particularly advantageous.
GQDs have demonstrated their versatility in facilitating gene expression by effectively delivering plasmid DNA (pDNA) and messenger RNA (mRNA). Ahn et al. demonstrated that positively charged N-doped GQDs coated with PEI could effectively deliver mRNA or pDNA into HeLa cells with low cytotoxicity [214]. The transfection efficiency with pDNA was similar to that of the positive control, while for mRNA it reached 27%, compared to the 18% achieved with Lipofectamine. Conversely, utilizing a higher molecular weight PEI for the functionalization of N-doped GQDs resulted in GQDs/mRNA complexes achieving up to 25% transfection efficiency at low doses in Huh-7 hepatocarcinoma cells, also capable of retaining considerable activity in a shear tolerance test [320]. However, these complexes were less effective and more toxic than the reference lipid nanoparticles. Notably, while the cationic charge intended to facilitate cellular uptake can enhance interaction with cell membranes, it can also lead to cytotoxic effects [321,322].
Other strategies have enhanced transfection efficiency [323]. As previously mentioned, modifying COOH-functionalized GQDs with PEI and an antibody targeted to EGFR enhanced the performance of a multifunctional system carrying a pDNA with a reporter gene and DOX in colon cancer cells [290]. When pDNA encoding the firefly luciferase gene was complexed with a positively-charged chimeric peptide (MPG-2H1 peptide) specifically designed to disrupt the endosomal membrane and containing a nuclear localization signal (NLS), subsequently attached to GQDs by electrostatic interactions, the complexes were capable of overcoming cellular trafficking barriers and showed a five-fold increase in transfection efficiency in HEK293t cells compared to the MPG-2H1 peptide/pDNA complex alone [324]. This was the first report on gene delivery using GQDs and could apply to various types of gene therapy, including potentially for cancer. Interestingly, they obtained two different sizes of GQDs with two different green and red emissions from the synthesis, but they did not separate the two populations before forming the complexes.
In addition, GQDs are a promising solution for enabling gene inhibition by efficiently delivering small interfering RNA (siRNA) and microRNA (miRNA). Even though the EphA2-siRNA loaded onto Eudragit-coated N-GQDs targeting the EphA2 receptor overexpressed in many solid tumors did not significantly reduce the viability of A549 lung tumor cells, a higher cell death rate was observed, likely due to the combined impact of the DNA effects from the GQDs and the effects of the siRNA [325]. A more efficient approach was designed by Yang et al., a novel nanocarrier consisting of N-doped GQDs functionalized with a biodegradable cationic polyester vector (BCPVs) and co-loaded with both DOX and a K-ras siRNA [300]. The resulting nanocomplexes exhibited high stability and demonstrated effective downregulation of K-ras expression, leading to significant inhibition of proliferation in MiaPaCa-2 pancreatic cells. The authors showed that BCPV improved the stability, biocompatibility, and cellular uptake of the loaded siRNA, aiding its escape from endosomes into the cytosol [326]. Additionally, the downregulation of mutant K-ras expression enhanced cell apoptosis, inhibited cell metastasis and colony formation, and is related to reducing drug resistance [327]. NIR-triggered photothermal effects further enhanced the synergistic anticancer activity of DOX and K-ras siRNA by improving cellular availability. In a separate study, nitrogen (N) and neodymium (Nd) were employed as doping ions to produce positively charged GQDs [264]. The NGQD/siRNA and Nd-NGQD/siRNA complexes effectively inhibited the expression of EGFR and K-ras proteins in HeLa cells by approximately 31-51%, demonstrating comparable efficacy to the Lipofectamine positive control. Furthermore, the addition of Nd enhanced the complexes’ NIR imaging capabilities.
Dong et al. designed a more intricate nanosystem by grafting PEG and PLA onto N-doped GQDs [328]. This nanosystem combined inhibitory probes (IP) and antisense oligodeoxynucleotide (ASODN) targeting miRNA-21 and survivin to enhance gene therapeutic efficacy. As a result, cell viability decreased to 68%, contrasting with GQDs loaded solely with IP (82% viability) or ASODN (76% viability). Moreover, cells transfected with the combination exhibited a significantly higher apoptotic ratio than those transfected with ASODN alone. This and other systems are also useful for detecting the presence of intracellular cancer-related miRNA [297].
Finally, drawing from the referenced studies, GQDs used for gene delivery in cancer have been synthesized from various sources, including natural compounds such as citric acid and glucosamine hydrochloride and other materials like graphite powder. Future perspectives include exploring green synthesis methods using natural carbon sources aiming for more sustainable and environmentally friendly production processes.
Although GQDs are not as extensively advanced as other carbon agents like GO [36,329], they hold promise as next-generation therapeutics for gene delivery in cancer therapy. Overcoming specific barriers, such as optimizing surface functionalization, endosomal escape, nuclear targeting, gene binding/release, biodegradability, immunogenicity, and in vivo tracking, is essential for their full development as gene therapy vectors. Similarly, for drug delivery, no clinical trials or approved products are currently available. Existing imaging strategies should be implemented towards modulating cancer-related genes, while other complex combinatorial therapies, such as those demonstrated in recent studies involving monocytes decorated with N-doped GQDs for miR223 delivery to macrophages within atherosclerotic plaques, show promise for targeted gene expression regulation and inflammation reduction, with potential implications for cancer therapy [330].
Figure 3 summarizes the different drug and gene delivery strategies based on GQDs systems used to eliminate cancer cells.

5. Conclusions and Future Perspectives

GQDs represent a promising frontier in cancer therapy due to their unique properties and versatile applications. These nanomaterials possess several advantageous characteristics, such as biocompatibility, non-toxicity, and high solubility. Their large surface area and flexibility in functional group manipulation allow for easy attachment of drugs and ligands, enhancing the stability and solubility of anticancer agents and making them effective for drug and gene-targeted delivery. GQDs hold significant promise in addressing research gaps in cancer therapy by offering synergistic treatment modalities and enhanced therapeutic efficacy, reducing toxicity compared to traditional approaches. Their ability to induce DNA damage in cancer cells, ROS generation, and enhancement of PDT and PTT contributes to tumor suppression. Optimizing the size and shape of GQDs can influence cellular uptake and gene delivery efficiency, while stimuli-responsive GQDs offer controlled gene release at specific sites. Hybrid systems combining GQDs with other materials further optimize delivery mechanisms, enhancing precision. Moreover, the interaction between GQDs and immune cells demonstrates promising potential for advancing cancer immunotherapy, enhancing therapeutic efficacy through tumor cell death and immune response activation.
However, despite these promising advancements, realizing the full biomedical utility of GQDs requires addressing several challenges and limitations in their application for cancer therapy. Ensuring biocompatibility and minimizing toxicity are primary concerns, necessitating precise synthesis techniques and surface functionalization methods. The synthesis of GQDs from natural carbon sources like biomass and organic waste highlights their potential for sustainable and scalable manufacturing. Top-down methods, such as electrochemical and liquid-phase exfoliation, are advantageous due to their ability to produce GQDs in large quantities and with relatively straightforward procedures. These methods, however, can yield mixtures of GQDs and larger graphene particles, posing a challenge in achieving uniform size and quality. Bottom-up methods, like pyrolysis and catalytic solution chemistry, offer higher precision in GQD morphology but can be more complex and resource-intensive. Refining synthesis protocols to ensure consistent, high-quality GQDs for clinical use remains a significant challenge. The methods used impact the physical, chemical, and biological properties of GQDs, influencing their effectiveness as therapeutic agents. Therapeutic efficacy also depends on the efficiency of drug-loading and controlled release systems, as well as strategies for ensuring stable dispersion and precise targeting. Achieving precise control over GQDs’ size, surface chemistry, and crystalline structure is difficult but crucial for their application in drug and gene delivery.
These findings underscore the importance of leveraging GQDs in future clinical trials to develop effective cancer therapies. While research on the in vitro and in vivo toxicity of GQDs is expanding, comprehensive studies are still needed to fully understand their safety profile. Despite the promising features and versatile applications of GQDs, the adoption of this technology in clinical settings has been slow. According to the National Institute of Health (NIH) database, as of today, only one clinical trial involves using GQDs in conjunction with Si nanowires as an early diagnostic tool for acute myocardial infarction (NCT04390490). This underscores the need for further research before these materials can be applied to biomedical applications. Lastly, to advance cancer treatment strategies using GQDs, further exploration, interdisciplinary research, and collaboration are encouraged. Researchers have highlighted the importance of GQDs in drug delivery, bioimaging, and targeted therapy, emphasizing the need for continued investigation into their future applications. Collaborative efforts between nanotechnology, medicine, and materials science experts can lead to innovative approaches in utilizing GQDs for cancer therapy, ultimately improving patient treatment outcomes.

Author Contributions

Conceptualization, H.M.O. and R.G.-P.; writing—original draft and figure preparation, H.M.O., F.C.-S., S.Y.B., C.R.-P. and R.G.-P.; writing—review and editing, H.M.O. and R.G.-P.; funding acquisition, H.M.O. and R.G.-P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Corporación Ecuatoriana para la Investigación y la Academia (CEDIA), grant number I+D+I-XVII-2022-48.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Bray, F.; Laversanne, M.; Sung, H.; Ferlay, J.; Siegel, R.L.; Soerjomataram, I.; Jemal, A. Global cancer statistics 2022: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2024, 74, 229–263. [Google Scholar] [CrossRef] [PubMed]
  2. Chen, S.; Cao, Z.; Prettner, K.; Kuhn, M.; Yang, J.; Jiao, L.; Wang, Z.; Li, W.; Geldsetzer, P.; Bärnighausen, T.; et al. Estimates and projections of the global economic cost of 29 cancers in 204 countries and territories from 2020 to 2050. JAMA Oncol. 2023, 9, 465–472. [Google Scholar] [CrossRef] [PubMed]
  3. Debela, D.T.; Muzazu, S.G.; Heraro, K.D.; Ndalama, M.T.; Mesele, B.W.; Haile, D.C.; Kitui, S.K.; Manyazewal, T. New approaches and procedures for cancer treatment: Current perspectives. SAGE Open Med. 2021, 9, 20503121211034370. [Google Scholar] [CrossRef] [PubMed]
  4. Pucci, C.; Martinelli, C.; Ciofani, G. Innovative approaches for cancer treatment: Current perspectives and new challenges. Ecancermedicalscience 2019, 13, 961. [Google Scholar] [CrossRef] [PubMed]
  5. Zhou, Y.; Chen, X.; Cao, J.; Gao, H. Overcoming the biological barriers in the tumor microenvironment for improving drug delivery and efficacy. J. Mater. Chem. B Mater. Biol. Med. 2020, 8, 6765–6781. [Google Scholar] [CrossRef]
  6. Shaha, S.; Rodrigues, D.; Mitragotri, S. Locoregional drug delivery for cancer therapy: Preclinical progress and clinical translation. J. Control. Release 2024, 367, 737–767. [Google Scholar] [CrossRef]
  7. Jin, J.; Wu, X.; Yin, J.; Li, M.; Shen, J.; Li, J.; Zhao, Y.; Zhao, Q.; Wu, J.; Wen, Q.; et al. Identification of genetic mutations in cancer: Challenge and opportunity in the new era of targeted therapy. Front. Oncol. 2019, 9, 263. [Google Scholar] [CrossRef]
  8. Asadi, M.; Zarredar, H.; Zafari, V.; Soleimani, Z.; Saeedi, H.; Caner, A.; Shanehbandi, D. Immune features of tumor microenvironment: A genetic spotlight. Cell Biochem. Biophys. 2024, 82, 107–118. [Google Scholar] [CrossRef]
  9. Watanabe, K.; Seki, N. Biology and Development of DNA-Targeted Drugs, Focusing on Synthetic Lethality, DNA Repair, and Epigenetic Modifications for Cancer: A Review. Int. J. Mol. Sci. 2024, 25, 752. [Google Scholar] [CrossRef]
  10. Belete, T.M. The current status of gene therapy for the treatment of cancer. Biologics 2021, 15, 67–77. [Google Scholar] [CrossRef]
  11. Transforming Our World: The 2030 Agenda for Sustainable Development | Department of Economic and Social Affairs. Available online: https://sdgs.un.org/2030agenda (accessed on 26 August 2024).
  12. Markman, J.L.; Rekechenetskiy, A.; Holler, E.; Ljubimova, J.Y. Nanomedicine therapeutic approaches to overcome cancer drug resistance. Adv. Drug Deliv. Rev. 2013, 65, 1866–1879. [Google Scholar] [CrossRef] [PubMed]
  13. Hejmady, S.; Pradhan, R.; Alexander, A.; Agrawal, M.; Singhvi, G.; Gorain, B.; Tiwari, S.; Kesharwani, P.; Dubey, S.K. Recent advances in targeted nanomedicine as promising antitumor therapeutics. Drug Discov. Today 2020, 25, 2227–2244. [Google Scholar] [CrossRef] [PubMed]
  14. Jian, W.; Hui, D.; Lau, D. Nanoengineering in biomedicine: Current development and future perspectives. Nanotechnol. Rev. 2020, 9, 700–715. [Google Scholar] [CrossRef]
  15. Zi, J.; Yu, X.; Li, Y.; Hu, X.; Xu, C.; Wang, X.; Liu, X.; Fu, R. Coloration strategies in peacock feathers. Proc. Natl. Acad. Sci. USA 2003, 100, 12576–12578. [Google Scholar] [CrossRef]
  16. Parker, A.R.; Townley, H.E. Biomimetics of photonic nanostructures. Nat. Nanotechnol. 2007, 2, 347–353. [Google Scholar] [CrossRef]
  17. Patra, J.K.; Das, G.; Fraceto, L.F.; Campos, E.V.R.; Rodriguez-Torres, M.D.P.; Acosta-Torres, L.S.; Diaz-Torres, L.A.; Grillo, R.; Swamy, M.K.; Sharma, S.; et al. Nano based drug delivery systems: Recent developments and future prospects. J. Nanobiotechnol. 2018, 16, 71. [Google Scholar] [CrossRef]
  18. Subhan, M.A.; Parveen, F.; Filipczak, N.; Yalamarty, S.S.K.; Torchilin, V.P. Approaches to Improve EPR-Based Drug Delivery for Cancer Therapy and Diagnosis. J. Pers. Med. 2023, 13, 389. [Google Scholar] [CrossRef]
  19. Blanco, E.; Shen, H.; Ferrari, M. Principles of nanoparticle design for overcoming biological barriers to drug delivery. Nat. Biotechnol. 2015, 33, 941–951. [Google Scholar] [CrossRef]
  20. Tade, R.S.; Patil, P.O. Theranostic prospects of graphene quantum dots in breast cancer. ACS Biomater. Sci. Eng. 2020, 6, 5987–6008. [Google Scholar] [CrossRef]
  21. Bajpai, S.; Tiwary, S.K.; Sonker, M.; Joshi, A.; Gupta, V.; Kumar, Y.; Shreyash, N.; Biswas, S. Recent Advances in Nanoparticle-Based Cancer Treatment: A Review. ACS Appl. Nano Mater. 2021, 4, 6441–6470. [Google Scholar] [CrossRef]
  22. Jampilek, J.; Kralova, K. Advances in Drug Delivery Nanosystems Using Graphene-Based Materials and Carbon Nanotubes. Materials 2021, 14, 1059. [Google Scholar] [CrossRef] [PubMed]
  23. Cataldo, F. Fullerenes. In New Frontiers in Nanochemistry: Concepts, Theories, and Trends; Putz, M.V., Ed.; Includes Bibliographical References and Indexes. | Contents: Volume 1. Structural Nanochemistry—Volume 2. Topological Nanochemistry—Volume 3. Sustainable Nanochemistry; Apple Academic Press: Palm Bay, FL, USA, 2020; pp. 219–225. ISBN 9780429022944. [Google Scholar]
  24. Tiwari, S.K.; Sahoo, S.; Wang, N.; Huczko, A. Graphene research and their outputs: Status and prospect. J. Sci. Adv. Mater. Devices 2020, 5, 10–29. [Google Scholar] [CrossRef]
  25. Plachá, D.; Jampilek, J. Graphenic materials for biomedical applications. Nanomaterials 2019, 9, 1758. [Google Scholar] [CrossRef]
  26. Yan, Y.; Gong, J.; Chen, J.; Zeng, Z.; Huang, W.; Pu, K.; Liu, J.; Chen, P. Recent advances on graphene quantum dots: From chemistry and physics to applications. Adv. Mater. 2019, 31, e1808283. [Google Scholar] [CrossRef]
  27. Tian, P.; Tang, L.; Teng, K.-S.; Lau, S.-P. Graphene quantum dots: Preparations, properties, functionalizations and applications. Mater. Futures 2024, 3, 022301. [Google Scholar] [CrossRef]
  28. Cui, Y.; Liu, L.; Shi, M.; Wang, Y.; Meng, X.; Chen, Y.; Huang, Q.; Liu, C. A review of advances in graphene quantum dots: From preparation and modification methods to application. C 2024, 10, 7. [Google Scholar] [CrossRef]
  29. Zhang, R.; Ding, Z. Recent advances in graphene quantum dots as bioimaging probes. J. Anal. Test. 2018, 2, 45–60. [Google Scholar] [CrossRef]
  30. Krasley, A.T.; Li, E.; Galeana, J.M.; Bulumulla, C.; Beyene, A.G.; Demirer, G.S. Carbon nanomaterial fluorescent probes and their biological applications. Chem. Rev. 2024, 124, 3085–3185. [Google Scholar] [CrossRef]
  31. Kalluri, A.; Dharmadhikari, B.; Debnath, D.; Patra, P.; Kumar, C.V. Advances in structural modifications and properties of graphene quantum dots for biomedical applications. ACS Omega 2023, 8, 21358–21376. [Google Scholar] [CrossRef]
  32. CN103432590B—Graphene Quantum Dot Nuclear Targeting Medicine Carrying System as Well as Preparation Method and Application Thereof—Google Patents. Available online: https://patents.google.com/patent/CN103432590B/en?oq=CN103432590B (accessed on 26 August 2024).
  33. CN104984349B—Epirubicin Loaded Graphene Quantum Dot Drug Carrying System and Preparation Method Thereof—Google Patents. Available online: https://patents.google.com/patent/CN104984349B/en?oq=CN104984349B (accessed on 26 August 2024).
  34. Zhao, C.; Song, X.; Liu, Y.; Fu, Y.; Ye, L.; Wang, N.; Wang, F.; Li, L.; Mohammadniaei, M.; Zhang, M.; et al. Synthesis of graphene quantum dots and their applications in drug delivery. J. Nanobiotechnol. 2020, 18, 142. [Google Scholar] [CrossRef]
  35. Zarepour, A.; Khosravi, A.; Yücel Ayten, N.; Çakır Hatır, P.; Iravani, S.; Zarrabi, A. Innovative approaches for cancer treatment: Graphene quantum dots for photodynamic and photothermal therapies. J. Mater. Chem. B Mater. Biol. Med. 2024, 12, 4307–4334. [Google Scholar] [CrossRef] [PubMed]
  36. Dar, M.S.; Tabish, T.A.; Thorat, N.D.; Swati, G.; Sahu, N.K. Photothermal therapy using graphene quantum dots. APL Bioeng. 2023, 7, 031502. [Google Scholar] [CrossRef] [PubMed]
  37. Iannazzo, D.; Celesti, C.; Espro, C. Recent advances on graphene quantum dots as multifunctional nanoplatforms for cancer treatment. Biotechnol. J. 2021, 16, e1900422. [Google Scholar] [CrossRef] [PubMed]
  38. Chong, Y.; Ma, Y.; Shen, H.; Tu, X.; Zhou, X.; Xu, J.; Dai, J.; Fan, S.; Zhang, Z. The in vitro and in vivo toxicity of graphene quantum dots. Biomaterials 2014, 35, 5041–5048. [Google Scholar] [CrossRef] [PubMed]
  39. Wang, S.; Cole, I.S.; Li, Q. The toxicity of graphene quantum dots. RSC Adv. 2016, 6, 89867–89878. [Google Scholar] [CrossRef]
  40. Chung, S.; Revia, R.A.; Zhang, M. Graphene quantum dots and their applications in bioimaging, biosensing, and therapy. Adv. Mater. 2021, 33, e1904362. [Google Scholar] [CrossRef]
  41. Ghaffarkhah, A.; Hosseini, E.; Kamkar, M.; Sehat, A.A.; Dordanihaghighi, S.; Allahbakhsh, A.; van der Kuur, C.; Arjmand, M. Synthesis, applications, and prospects of graphene quantum dots: A comprehensive review. Small 2022, 18, e2102683. [Google Scholar] [CrossRef]
  42. Kundu, S.; Pillai, V.K. Synthesis and characterization of graphene quantum dots. Phys. Sci. Rev. 2019, 5, 20190013. [Google Scholar] [CrossRef]
  43. Ban, G.; Hou, Y.; Shen, Z.; Jia, J.; Chai, L.; Ma, C. Potential biomedical limitations of graphene nanomaterials. Int. J. Nanomed. 2023, 18, 1695–1708. [Google Scholar] [CrossRef]
  44. Lyu, B.; Li, H.-J.; Xue, F.; Sai, L.; Gui, B.; Qian, D.; Wang, X.; Yang, J. Facile, gram-scale and eco-friendly synthesis of multi-color graphene quantum dots by thermal-driven advanced oxidation process. Chem. Eng. J. 2020, 388, 124285. [Google Scholar] [CrossRef]
  45. Kumar, P.; Dhand, C.; Dwivedi, N.; Singh, S.; Khan, R.; Verma, S.; Singh, A.; Gupta, M.K.; Kumar, S.; Kumar, R.; et al. Graphene quantum dots: A contemporary perspective on scope, opportunities, and sustainability. Renew. Sustain. Energy Rev. 2022, 157, 111993. [Google Scholar] [CrossRef]
  46. Gozali Balkanloo, P.; Mohammad Sharifi, K.; Poursattar Marjani, A. Graphene quantum dots: Synthesis, characterization, and application in wastewater treatment: A review. Mater. Adv. 2023, 4, 4272–4293. [Google Scholar] [CrossRef]
  47. Wilson, A.; Nambiar, A.; Sobha, A. Synthesis of graphene quantum dots by liquid phase exfoliation method and its applications. Mater. Today Proc. 2023, in press. [Google Scholar] [CrossRef]
  48. Zdrazil, L.; Zahradnicek, R.; Mohan, R.; Sedlacek, P.; Nejdl, L.; Schmiedova, V.; Pospisil, J.; Horak, M.; Weiter, M.; Zmeskal, O.; et al. Preparation of graphene quantum dots through liquid phase exfoliation method. J. Lumin. 2018, 204, 203–208. [Google Scholar] [CrossRef]
  49. Hummers, W.S.; Offeman, R.E. Preparation of Graphitic Oxide. J. Am. Chem. Soc. 1958, 80, 1339. [Google Scholar] [CrossRef]
  50. Yadav, N.; Lochab, B. A comparative study of graphene oxide: Hummers, intermediate and improved method. FlatChem 2019, 13, 40–49. [Google Scholar] [CrossRef]
  51. Zaaba, N.I.; Foo, K.L.; Hashim, U.; Tan, S.J.; Liu, W.-W.; Voon, C.H. Synthesis of Graphene Oxide using Modified Hummers Method: Solvent Influence. Procedia Eng. 2017, 184, 469–477. [Google Scholar] [CrossRef]
  52. Pan, D.; Zhang, J.; Li, Z.; Wu, M. Hydrothermal route for cutting graphene sheets into blue-luminescent graphene quantum dots. Adv. Mater. 2010, 22, 734–738. [Google Scholar] [CrossRef]
  53. Pan, D.; Guo, L.; Zhang, J.; Xi, C.; Xue, Q.; Huang, H.; Li, J.; Zhang, Z.; Yu, W.; Chen, Z.; et al. Cutting sp2 clusters in graphene sheets into colloidal graphene quantum dots with strong green fluorescence. J. Mater. Chem. 2012, 22, 3314. [Google Scholar] [CrossRef]
  54. Chen, W.; Lv, G.; Hu, W.; Li, D.; Chen, S.; Dai, Z. Synthesis and applications of graphene quantum dots: A review. Nanotechnol. Rev. 2018, 7, 157–185. [Google Scholar] [CrossRef]
  55. Kumar, Y.R.; Deshmukh, K.; Sadasivuni, K.K.; Pasha, S.K.K. Graphene quantum dot based materials for sensing, bio-imaging and energy storage applications: A review. RSC Adv. 2020, 10, 23861–23898. [Google Scholar] [CrossRef] [PubMed]
  56. Xie, J.-D.; Lai, G.-W.; Huq, M.M. Hydrothermal route to graphene quantum dots: Effects of precursor and temperature. Diam. Relat. Mater. 2017, 79, 112–118. [Google Scholar] [CrossRef]
  57. Chen, Y.-X.; Lu, D.; Wang, G.-G.; Huangfu, J.; Wu, Q.-B.; Wang, X.-F.; Liu, L.-F.; Ye, D.-M.; Yan, B.; Han, J. Highly Efficient Orange Emissive Graphene Quantum Dots Prepared by Acid-Free Method for White LEDs. ACS Sustain. Chem. Eng. 2020, 8, 6657–6666. [Google Scholar] [CrossRef]
  58. Peng, J.; Gao, W.; Gupta, B.K.; Liu, Z.; Romero-Aburto, R.; Ge, L.; Song, L.; Alemany, L.B.; Zhan, X.; Gao, G.; et al. Graphene quantum dots derived from carbon fibers. Nano Lett. 2012, 12, 844–849. [Google Scholar] [CrossRef]
  59. Lee, S.H.; Kim, D.Y.; Lee, J.; Lee, S.B.; Han, H.; Kim, Y.Y.; Mun, S.C.; Im, S.H.; Kim, T.-H.; Park, O.O. Synthesis of Single-Crystalline Hexagonal Graphene Quantum Dots from Solution Chemistry. Nano Lett. 2019, 19, 5437–5442. [Google Scholar] [CrossRef]
  60. Ruammaitree, A.; Praphanwong, K.; Taiphol, A. Facile one-step hydrothermal synthesis of monolayer and turbostratic bilayer n-doped graphene quantum dots using sucrose as a carbon source. RSC Adv. 2023, 13, 23700–23707. [Google Scholar] [CrossRef]
  61. Bayat, A.; Saievar-Iranizad, E. Synthesis of green-photoluminescent single layer graphene quantum dots: Determination of HOMO and LUMO energy states. J. Lumin. 2017, 192, 180–183. [Google Scholar] [CrossRef]
  62. Yan, Y.; Chen, J.; Li, N.; Tian, J.; Li, K.; Jiang, J.; Liu, J.; Tian, Q.; Chen, P. Systematic bandgap engineering of graphene quantum dots and applications for photocatalytic water splitting and CO2 reduction. ACS Nano 2018, 12, 3523–3532. [Google Scholar] [CrossRef]
  63. Mao, Y.; Zhao, C.; Ge, S.; Luo, T.; Chen, J.; Liu, J.; Xi, F.; Liu, J. Gram-scale synthesis of nitrogen doped graphene quantum dots for sensitive detection of mercury ions and l-cysteine. RSC Adv. 2019, 9, 32977–32983. [Google Scholar] [CrossRef]
  64. Hong, G.-L.; Zhao, H.-L.; Deng, H.-H.; Yang, H.-J.; Peng, H.-P.; Liu, Y.-H.; Chen, W. Fabrication of ultra-small monolayer graphene quantum dots by pyrolysis of trisodium citrate for fluorescent cell imaging. Int. J. Nanomed. 2018, 13, 4807–4815. [Google Scholar] [CrossRef]
  65. Zhang, R.; Adsetts, J.R.; Nie, Y.; Sun, X.; Ding, Z. Electrochemiluminescence of nitrogen- and sulfur-doped graphene quantum dots. Carbon 2018, 129, 45–53. [Google Scholar] [CrossRef]
  66. Yan, X.; Cui, X.; Li, L.-S. Synthesis of large, stable colloidal graphene quantum dots with tunable size. J. Am. Chem. Soc. 2010, 132, 5944–5945. [Google Scholar] [CrossRef] [PubMed]
  67. Shi, L.; Wang, B.; Lu, S. Efficient bottom-up synthesis of graphene quantum dots at an atomically precise level. Matter 2023, 6, 728–760. [Google Scholar] [CrossRef]
  68. Lee, Y.-J.; Yeh, T.-W.; Zou, C.; Yang, Z.-P.; Chen, J.-W.; Hsu, P.-L.; Shen, J.-L.; Chang, C.-C.; Sheu, J.-K.; Lin, L.Y. Graphene Quantum Dot Vertical Cavity Surface-Emitting Lasers. ACS Photonics 2019, 6, 2894–2901. [Google Scholar] [CrossRef]
  69. Huang, H.; Yang, S.; Li, Q.; Yang, Y.; Wang, G.; You, X.; Mao, B.; Wang, H.; Ma, Y.; He, P.; et al. Electrochemical cutting in weak aqueous electrolytes: The strategy for efficient and controllable preparation of graphene quantum dots. Langmuir 2018, 34, 250–258. [Google Scholar] [CrossRef]
  70. Kalita, H.; Palaparthy, V.S.; Baghini, M.S.; Aslam, M. Electrochemical synthesis of graphene quantum dots from graphene oxide at room temperature and its soil moisture sensing properties. Carbon 2020, 165, 9–17. [Google Scholar] [CrossRef]
  71. Ahirwar, S.; Mallick, S.; Bahadur, D. Electrochemical method to prepare graphene quantum dots and graphene oxide quantum dots. ACS Omega 2017, 2, 8343–8353. [Google Scholar] [CrossRef]
  72. Wang, Y.; Liu, Y.; Zhang, J.; Wu, J.; Xu, H.; Wen, X.; Zhang, X.; Tiwary, C.S.; Yang, W.; Vajtai, R.; et al. Cryo-mediated exfoliation and fracturing of layered materials into 2D quantum dots. Sci. Adv. 2017, 3, e1701500. [Google Scholar] [CrossRef]
  73. Das, R.C.; Gogoi, K.K.; Das, N.S.; Chowdhury, A. Optimization of quantum yield of highly luminescent graphene oxide quantum dots and their application in resistive memory devices. Semicond. Sci. Technol. 2019, 34, 125016. [Google Scholar] [CrossRef]
  74. Joffrion, J.B.; Clower, W.; Wilson, C.G. Tunable excitation-independent emissions from graphene quantum dots through microplasma-assisted electrochemical synthesis. Nano-Struct. Nano-Objects 2019, 19, 100341. [Google Scholar] [CrossRef]
  75. Ananthanarayanan, A.; Wang, X.; Routh, P.; Sana, B.; Lim, S.; Kim, D.; Lim, K.; Li, J.; Chen, P. Facile Synthesis of Graphene Quantum Dots from 3D Graphene and their Application for Fe3+ Sensing. Adv. Funct. Mater. 2014, 24, 3021–3026. [Google Scholar] [CrossRef]
  76. Alidad, F.; Navik, R.; Gai, Y.; Zhao, Y. Production of pristine graphene quantum dots from graphite by a shear-mixer in supercritical CO2. Chem. Phys. Lett. 2018, 710, 64–69. [Google Scholar] [CrossRef]
  77. Gao, H.; Xue, C.; Hu, G.; Zhu, K. Production of graphene quantum dots by ultrasound-assisted exfoliation in supercritical CO2/H2O medium. Ultrason. Sonochem. 2017, 37, 120–127. [Google Scholar] [CrossRef]
  78. Haque, E.; Kim, J.; Malgras, V.; Reddy, K.R.; Ward, A.C.; You, J.; Bando, Y.; Hossain, M.D.S.A.; Yamauchi, Y. Recent advances in graphene quantum dots: Synthesis, properties, and applications. Small Methods 2018, 2, 1800050. [Google Scholar] [CrossRef]
  79. Lu, Q.; Wu, C.; Liu, D.; Wang, H.; Su, W.; Li, H.; Zhang, Y.; Yao, S. A facile and simple method for synthesis of graphene oxide quantum dots from black carbon. Green Chem. 2017, 19, 900–904. [Google Scholar] [CrossRef]
  80. Nair, R.V.; Thomas, R.T.; Sankar, V.; Muhammad, H.; Dong, M.; Pillai, S. Rapid, Acid-Free Synthesis of High-Quality Graphene Quantum Dots for Aggregation Induced Sensing of Metal Ions and Bioimaging. ACS Omega 2017, 2, 8051–8061. [Google Scholar] [CrossRef]
  81. Gu, S.; Hsieh, C.-T.; Ashraf Gandomi, Y.; Chang, J.-K.; Li, J.; Li, J.; Zhang, H.; Guo, Q.; Lau, K.C.; Pandey, R. Microwave growth and tunable photoluminescence of nitrogen-doped graphene and carbon nitride quantum dots. J. Mater. Chem. C 2019, 7, 5468–5476. [Google Scholar] [CrossRef]
  82. Gu, D.; Shang, S.; Yu, Q.; Shen, J. Green synthesis of nitrogen-doped carbon dots from lotus root for Hg(II) ions detection and cell imaging. Appl. Surf. Sci. 2016, 390, 38–42. [Google Scholar] [CrossRef]
  83. Wu, J.; Pisula, W.; Müllen, K. Graphenes as potential material for electronics. Chem. Rev. 2007, 107, 718–747. [Google Scholar] [CrossRef]
  84. Zhao, X.-J.; Hou, H.; Ding, P.-P.; Deng, Z.-Y.; Ju, Y.-Y.; Liu, S.-H.; Liu, Y.-M.; Tang, C.; Feng, L.-B.; Tan, Y.-Z. Molecular defect-containing bilayer graphene exhibiting brightened luminescence. Sci. Adv. 2020, 6, eaay8541. [Google Scholar] [CrossRef]
  85. Tang, L.; Ji, R.; Li, X.; Teng, K.S.; Lau, S.P. Size-dependent structural and optical characteristics of glucose-derived graphene quantum dots. Part. Part. Syst. Charact. 2013, 30, 523–531. [Google Scholar] [CrossRef]
  86. Zhi, L.; Müllen, K. A bottom-up approach from molecular nanographenes to unconventional carbon materials. J. Mater. Chem. 2008, 18, 1472. [Google Scholar] [CrossRef]
  87. Wang, X.; Zhi, L.; Tsao, N.; Tomović, Z.; Li, J.; Müllen, K. Transparent carbon films as electrodes in organic solar cells. Angew. Chem. Int. Ed. 2008, 47, 2990–2992. [Google Scholar] [CrossRef]
  88. Liu, R.; Wu, D.; Feng, X.; Müllen, K. Nitrogen-doped ordered mesoporous graphitic arrays with high electrocatalytic activity for oxygen reduction. Angew. Chem. Int. Ed. 2010, 49, 2565–2569. [Google Scholar] [CrossRef]
  89. Li, F.; Li, Y.; Yang, X.; Han, X.; Jiao, Y.; Wei, T.; Yang, D.; Xu, H.; Nie, G. Highly Fluorescent Chiral N-S-Doped Carbon Dots from Cysteine: Affecting Cellular Energy Metabolism. Angew. Chem. Int. Ed. 2018, 57, 2377–2382. [Google Scholar] [CrossRef]
  90. Liu, R.; Wu, D.; Feng, X.; Müllen, K. Bottom-up fabrication of photoluminescent graphene quantum dots with uniform morphology. J. Am. Chem. Soc. 2011, 133, 15221–15223. [Google Scholar] [CrossRef]
  91. Bressi, V.; Ferlazzo, A.; Iannazzo, D.; Espro, C. Graphene Quantum Dots by Eco-Friendly Green Synthesis for Electrochemical Sensing: Recent Advances and Future Perspectives. Nanomaterials 2021, 11, 1120. [Google Scholar] [CrossRef]
  92. Ye, R.; Xiang, C.; Lin, J.; Peng, Z.; Huang, K.; Yan, Z.; Cook, N.P.; Samuel, E.L.G.; Hwang, C.-C.; Ruan, G.; et al. Coal as an abundant source of graphene quantum dots. Nat. Commun. 2013, 4, 2943. [Google Scholar] [CrossRef]
  93. Xu, Y.; Wang, S.; Hou, X.; Sun, Z.; Jiang, Y.; Dong, Z.; Tao, Q.; Man, J.; Cao, Y. Coal-derived nitrogen, phosphorus and sulfur co-doped graphene quantum dots: A promising ion fluorescent probe. Appl. Surf. Sci. 2018, 445, 519–526. [Google Scholar] [CrossRef]
  94. Yew, Y.T.; Loo, A.H.; Sofer, Z.; Klímová, K.; Pumera, M. Coke-derived graphene quantum dots as fluorescence nanoquencher in DNA detection. Appl. Mater. Today 2017, 7, 138–143. [Google Scholar] [CrossRef]
  95. He, M.; Guo, X.; Huang, J.; Shen, H.; Zeng, Q.; Wang, L. Mass production of tunable multicolor graphene quantum dots from an energy resource of coke by a one-step electrochemical exfoliation. Carbon 2018, 140, 508–520. [Google Scholar] [CrossRef]
  96. Wareing, T.C.; Gentile, P.; Phan, A.N. Biomass-Based Carbon Dots: Current Development and Future Perspectives. ACS Nano 2021, 15, 15471–15501. [Google Scholar] [CrossRef] [PubMed]
  97. Lou, Y.; Hao, X.; Liao, L.; Zhang, K.; Chen, S.; Li, Z.; Ou, J.; Qin, A.; Li, Z. Recent advances of biomass carbon dots on syntheses, characterization, luminescence mechanism, and sensing applications. Nano Sel. 2021, 2, 1117–1145. [Google Scholar] [CrossRef]
  98. Kim, D.J.; Yoo, J.M.; Suh, Y.; Kim, D.; Kang, I.; Moon, J.; Park, M.; Kim, J.; Kang, K.-S.; Hong, B.H. Graphene Quantum Dots from Carbonized Coffee Bean Wastes for Biomedical Applications. Nanomaterials 2021, 11, 1423. [Google Scholar] [CrossRef]
  99. 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]
  100. Chai, X.; He, H.; Fan, H.; Kang, X.; Song, X. A hydrothermal-carbonization process for simultaneously production of sugars, graphene quantum dots, and porous carbon from sugarcane bagasse. Bioresour. Technol. 2019, 282, 142–147. [Google Scholar] [CrossRef]
  101. Nafiujjaman, M.; Kim, J.; Park, H.-K.; Lee, Y.-K. Preparation of blue-color-emitting graphene quantum dots and their in vitro and in vivo toxicity evaluation. J. Ind. Eng. Chem. 2017, 57, 171–180. [Google Scholar] [CrossRef]
  102. Wu, X.; Tian, F.; Wang, W.; Chen, J.; Wu, M.; Zhao, J.X. Fabrication of highly fluorescent graphene quantum dots using L-glutamic acid for in vitro/in vivo imaging and sensing. J. Mater. Chem. C Mater. Opt. Electron. Devices 2013, 1, 4676–4684. [Google Scholar] [CrossRef]
  103. Zhang, C.; Cui, Y.; Song, L.; Liu, X.; Hu, Z. Microwave assisted one-pot synthesis of graphene quantum dots as highly sensitive fluorescent probes for detection of iron ions and pH value. Talanta 2016, 150, 54–60. [Google Scholar] [CrossRef]
  104. Kadyan, P.; Malik, R.; Bhatia, S.; Al Harrasi, A.; Mohan, S.; Yadav, M.; Dalal, S.; Ramniwas, S.; Kumar Kataria, S.; Arasu, T. Comprehensive review on synthesis, applications, and challenges of graphene quantum dots (gqds). J. Nanomater. 2023, 2023, 2832964. [Google Scholar] [CrossRef]
  105. Kalita, H.; Mohapatra, J.; Pradhan, L.; Mitra, A.; Bahadur, D.; Aslam, M. Efficient synthesis of rice based graphene quantum dots and their fluorescent properties. RSC Adv. 2016, 6, 23518–23524. [Google Scholar] [CrossRef]
  106. Wang, X.; Sun, X.; Lao, J.; He, H.; Cheng, T.; Wang, M.; Wang, S.; Huang, F. Multifunctional graphene quantum dots for simultaneous targeted cellular imaging and drug delivery. Colloids Surf. B Biointerfaces 2014, 122, 638–644. [Google Scholar] [CrossRef] [PubMed]
  107. Kumawat, M.K.; Thakur, M.; Gurung, R.B.; Srivastava, R. Graphene Quantum Dots from Mangifera indica: Application in Near-Infrared Bioimaging and Intracellular Nanothermometry. ACS Sustain. Chem. Eng. 2017, 5, 1382–1391. [Google Scholar] [CrossRef]
  108. Su, J.; Zhang, X.; Tong, X.; Wang, X.; Yang, P.; Yao, F.; Guo, R.; Yuan, C. Preparation of graphene quantum dots with high quantum yield by a facile one-step method and applications for cell imaging. Mater. Lett. 2020, 271, 127806. [Google Scholar] [CrossRef]
  109. Chen, W.; Shen, J.; Lv, G.; Li, D.; Hu, Y.; Zhou, C.; Liu, X.; Dai, Z. Green Synthesis of Graphene Quantum Dots from Cotton Cellulose. Chem. Sel. 2019, 4, 2898–2902. [Google Scholar] [CrossRef]
  110. Wang, Z.; Yu, J.; Zhang, X.; Li, N.; Liu, B.; Li, Y.; Wang, Y.; Wang, W.; Li, Y.; Zhang, L.; et al. Large-Scale and Controllable Synthesis of Graphene Quantum Dots from Rice Husk Biomass: A Comprehensive Utilization Strategy. ACS Appl. Mater. Interfaces 2016, 8, 1434–1439. [Google Scholar] [CrossRef]
  111. Geng, H.; Qiu, J.; Zhu, H.; Liu, X. Achieving stem cell imaging and osteogenic differentiation by using nitrogen doped graphene quantum dots. J. Mater. Sci. Mater. Med. 2018, 29, 85. [Google Scholar] [CrossRef]
  112. Wang, X.; Sun, X.; He, H.; Yang, H.; Lao, J.; Song, Y.; Xia, Y.; Xu, H.; Zhang, X.; Huang, F. A two-component active targeting theranostic agent based on graphene quantum dots. J. Mater. Chem. B 2015, 3, 3583–3590. [Google Scholar] [CrossRef]
  113. Khodadadei, F.; Safarian, S.; Ghanbari, N. Methotrexate-loaded nitrogen-doped graphene quantum dots nanocarriers as an efficient anticancer drug delivery system. Mater. Sci. Eng. C Mater. Biol. Appl. 2017, 79, 280–285. [Google Scholar] [CrossRef]
  114. Kumawat, M.K.; Thakur, M.; Bahadur, R.; Kaku, T.; Prabhuraj, R.S.; Suchitta, A.; Srivastava, R. Preparation of graphene oxide-graphene quantum dots hybrid and its application in cancer theranostics. Mater. Sci. Eng. C Mater. Biol. Appl. 2019, 103, 109774. [Google Scholar] [CrossRef]
  115. Thakur, M.; Mewada, A.; Pandey, S.; Bhori, M.; Singh, K.; Sharon, M.; Sharon, M. Milk-derived multi-fluorescent graphene quantum dot-based cancer theranostic system. Mater. Sci. Eng. C Mater. Biol. Appl. 2016, 67, 468–477. [Google Scholar] [CrossRef] [PubMed]
  116. Castro Neto, A.H.; Guinea, F.; Peres, N.M.R.; Novoselov, K.S.; Geim, A.K. The electronic properties of graphene. Rev. Mod. Phys. 2009, 81, 109–162. [Google Scholar] [CrossRef]
  117. Novoselov, K.S.; Morozov, S.V.; Mohinddin, T.M.G.; Ponomarenko, L.A.; Elias, D.C.; Yang, R.; Barbolina, I.I.; Blake, P.; Booth, T.J.; Jiang, D.; et al. Electronic properties of graphene. Phys. Status Solidi B 2007, 244, 4106–4111. [Google Scholar] [CrossRef]
  118. Gokus, T.; Nair, R.R.; Bonetti, A.; Böhmler, M.; Lombardo, A.; Novoselov, K.S.; Geim, A.K.; Ferrari, A.C.; Hartschuh, A. Making graphene luminescent by oxygen plasma treatment. ACS Nano 2009, 3, 3963–3968. [Google Scholar] [CrossRef]
  119. Sk, M.A.; Ananthanarayanan, A.; Huang, L.; Lim, K.H.; Chen, P. Revealing the tunable photoluminescence properties of graphene quantum dots. J. Mater. Chem. C 2014, 2, 6954–6960. [Google Scholar] [CrossRef]
  120. Campuzano, S.; Yáñez-Sedeño, P.; Pingarrón, J.M. Carbon dots and graphene quantum dots in electrochemical biosensing. Nanomaterials 2019, 9, 634. [Google Scholar] [CrossRef]
  121. Cho, H.; Bae, G.; Hong, B.H. Engineering functionalization and properties of graphene quantum dots (GQDs) with controllable synthesis for energy and display applications. Nanoscale 2024, 16, 3347–3378. [Google Scholar] [CrossRef]
  122. Facure, M.H.M.; Schneider, R.; Mercante, L.A.; Correa, D.S. A review on graphene quantum dots and their nanocomposites: From laboratory synthesis towards agricultural and environmental applications. Environ. Sci. Nano 2020, 7, 3710–3734. [Google Scholar] [CrossRef]
  123. Yeh, T.-F.; Huang, W.-L.; Chung, C.-J.; Chiang, I.-T.; Chen, L.-C.; Chang, H.-Y.; Su, W.-C.; Cheng, C.; Chen, S.-J.; Teng, H. Elucidating Quantum Confinement in Graphene Oxide Dots Based on Excitation-Wavelength-Independent Photoluminescence. J. Phys. Chem. Lett. 2016, 7, 2087–2092. [Google Scholar] [CrossRef]
  124. Yang, J.-S.; Chang, Y.-C.; Huang, Q.-H.; Lai, Y.-Y.; Chiang, W.-H. Microplasma-enabled nanocarbon assembly for the diameter-selective synthesis of colloidal graphene quantum dots. Chem. Commun. 2020, 56, 10365–10368. [Google Scholar] [CrossRef]
  125. Yuan, F.; Yuan, T.; Sui, L.; Wang, Z.; Xi, Z.; Li, Y.; Li, X.; Fan, L.; Tan, Z.; Chen, A.; et al. Engineering triangular carbon quantum dots with unprecedented narrow bandwidth emission for multicolored LEDs. Nat. Commun. 2018, 9, 2249. [Google Scholar] [CrossRef] [PubMed]
  126. Rabeya, R.; Mahalingam, S.; Manap, A.; Satgunam, M.; Akhtaruzzaman, M.D.; Chia, C.H. Structural defects in graphene quantum dots: A review. Int. J. Quantum Chem. 2022, 122, e26900. [Google Scholar] [CrossRef]
  127. Zhu, S.; Song, Y.; Wang, J.; Wan, H.; Zhang, Y.; Ning, Y.; Yang, B. Photoluminescence mechanism in graphene quantum dots: Quantum confinement effect and surface/edge state. Nano Today 2017, 13, 10–14. [Google Scholar] [CrossRef]
  128. Ritter, K.A.; Lyding, J.W. The influence of edge structure on the electronic properties of graphene quantum dots and nanoribbons. Nat. Mater. 2009, 8, 235–242. [Google Scholar] [CrossRef] [PubMed]
  129. Lingam, K.; Podila, R.; Qian, H.; Serkiz, S.; Rao, A.M. Evidence for edge-state photoluminescence in graphene quantum dots. Adv. Funct. Mater. 2013, 23, 5062–5065. [Google Scholar] [CrossRef]
  130. Sun, H.; Zhao, A.; Gao, N.; Li, K.; Ren, J.; Qu, X. Deciphering a nanocarbon-based artificial peroxidase: Chemical identification of the catalytically active and substrate-binding sites on graphene quantum dots. Angew. Chem. Int. Ed. 2015, 54, 7176–7180. [Google Scholar] [CrossRef]
  131. Li, K.; Chen, J.; Yan, Y.; Min, Y.; Li, H.; Xi, F.; Liu, J.; Chen, P. Quasi-homogeneous carbocatalysis for one-pot selective conversion of carbohydrates to 5-hydroxymethylfurfural using sulfonated graphene quantum dots. Carbon 2018, 136, 224–233. [Google Scholar] [CrossRef]
  132. Ravi, P.V.; Subramaniyam, V.; Saravanakumar, N.; Pattabiraman, A.; Pichumani, M. What works and what doesn’t when graphene quantum dots are functionalized for contemporary applications? Coord. Chem. Rev. 2023, 493, 215270. [Google Scholar] [CrossRef]
  133. Du, T.; She, J.; Yang, X.; Zhao, Y.; Zhou, S.; Zhao, J. Role of functionalization in the fluorescence quantum yield of graphene quantum dots. Appl. Phys. Lett. 2023, 122, 142107. [Google Scholar] [CrossRef]
  134. Yeh, T.-F.; Teng, C.-Y.; Chen, S.-J.; Teng, H. Nitrogen-doped graphene oxide quantum dots as photocatalysts for overall water-splitting under visible light illumination. Adv. Mater. 2014, 26, 3297–3303. [Google Scholar] [CrossRef]
  135. Tetsuka, H. Nitrogen-Functionalized Graphene Quantum Dots: A Versatile Platform for Integrated Optoelectronic Devices. Chem. Rec. 2020, 20, 429–439. [Google Scholar] [CrossRef] [PubMed]
  136. Kadian, S.; Manik, G.; Ashish, K.; Singh, M.; Chauhan, R.P. Effect of sulfur doping on fluorescence and quantum yield of graphene quantum dots: An experimental and theoretical investigation. Nanotechnology 2019, 30, 435704. [Google Scholar] [CrossRef]
  137. Wang, H.; Revia, R.; Mu, Q.; Lin, G.; Yen, C.; Zhang, M. Single-layer boron-doped graphene quantum dots for contrast-enhanced in vivo T1-weighted MRI. Nanoscale Horiz. 2020, 5, 573–579. [Google Scholar] [CrossRef] [PubMed]
  138. Du, F.; Zeng, Q.; Lai, Z.; Cheng, Z.; Ruan, G. Silicon doped graphene quantum dots combined with ruthenium (iii) ions as a fluorescent probe for turn-on detection of triclosan. New J. Chem. 2019, 43, 12907–12915. [Google Scholar] [CrossRef]
  139. Boonta, W.; Talodthaisong, C.; Sattayaporn, S.; Chaicham, C.; Chaicham, A.; Sahasithiwat, S.; Kangkaew, L.; Kulchat, S. The synthesis of nitrogen and sulfur co-doped graphene quantum dots for fluorescence detection of cobalt (ii) ions in water. Mater. Chem. Front. 2020, 4, 507–516. [Google Scholar] [CrossRef]
  140. Xu, L.; Dun, X.; Zou, J.; Li, Y.; Jia, M.; Cui, L.; Gao, J.; Jin, X. Graphene Hydrogel Decorated with N, O Co-Doped Carbon Dots for Flexible Supercapacitor Electrodes. J. Electrochem. Soc. 2018, 165, A2217–A2224. [Google Scholar] [CrossRef]
  141. Zhang, Y.; Zhao, J.; Sun, H.; Zhu, Z.; Zhang, J.; Liu, Q. B, N, S, Cl doped graphene quantum dots and their effects on gas-sensing properties of Ag-LaFeO3. Sens. Actuators B Chem. 2018, 266, 364–374. [Google Scholar] [CrossRef]
  142. Eda, G.; Lin, Y.-Y.; Mattevi, C.; Yamaguchi, H.; Chen, H.-A.; Chen, I.-S.; Chen, C.-W.; Chhowalla, M. Blue photoluminescence from chemically derived graphene oxide. Adv. Mater. 2010, 22, 505–509. [Google Scholar] [CrossRef]
  143. Loh, K.P.; Bao, Q.; Eda, G.; Chhowalla, M. Graphene oxide as a chemically tunable platform for optical applications. Nat. Chem. 2010, 2, 1015–1024. [Google Scholar] [CrossRef]
  144. Wang, B.; Yu, J.; Sui, L.; Zhu, S.; Tang, Z.; Yang, B.; Lu, S. Rational Design of Multi-Color-Emissive Carbon Dots in a Single Reaction System by Hydrothermal. Adv. Sci. 2020, 8, 2001453. [Google Scholar] [CrossRef]
  145. Rajhi, A.A.; Alaziz, K.M.A.; Oviedo, B.S.R.; Yadav, A.; Hernández, E.; Gallegos, C.; Alamri, S.; Duhduh, A.A. Enhancing the performance of quantum dot solar cells through halogen adatoms on carboxyl edge-functionalized graphene quantum dots. J. Photochem. Photobiol. A 2024, 447, 115240. [Google Scholar] [CrossRef]
  146. Saleem, Y.; Najera Baldo, L.; Delgado, A.; Szulakowska, L.; Hawrylak, P. Oscillations of the bandgap with size in armchair and zigzag graphene quantum dots. J. Phys. Condens. Matter 2019, 31, 305503. [Google Scholar] [CrossRef] [PubMed]
  147. Ye, R.; Peng, Z.; Metzger, A.; Lin, J.; Mann, J.A.; Huang, K.; Xiang, C.; Fan, X.; Samuel, E.L.G.; Alemany, L.B.; et al. Bandgap engineering of coal-derived graphene quantum dots. ACS Appl. Mater. Interfaces 2015, 7, 7041–7048. [Google Scholar] [CrossRef] [PubMed]
  148. Fan, T.; Zeng, W.; Tang, W.; Yuan, C.; Tong, S.; Cai, K.; Liu, Y.; Huang, W.; Min, Y.; Epstein, A.J. Controllable size-selective method to prepare graphene quantum dots from graphene oxide. Nanoscale Res. Lett. 2015, 10, 55. [Google Scholar] [CrossRef] [PubMed]
  149. Medina-Berríos, N.; Pantoja-Romero, W.; Lavín Flores, A.; Díaz Vélez, S.C.; Martínez Guadalupe, A.C.; Torres Mulero, M.T.; Kisslinger, K.; Martínez-Ferrer, M.; Morell, G.; Weiner, B.R. Synthesis and Characterization of Carbon-Based Quantum Dots and Doped Derivatives for Improved Andrographolide’s Hydrophilicity in Drug Delivery Platforms. ACS Omega 2024, 9, 12575–12584. [Google Scholar] [CrossRef]
  150. Kim, S.; Hwang, S.W.; Kim, M.-K.; Shin, D.Y.; Shin, D.H.; Kim, C.O.; Yang, S.B.; Park, J.H.; Hwang, E.; Choi, S.-H.; et al. Anomalous behaviors of visible luminescence from graphene quantum dots: Interplay between size and shape. ACS Nano 2012, 6, 8203–8208. [Google Scholar] [CrossRef]
  151. Kang, H.; Kim, D.Y.; Cho, J. Top-Down Fabrication of Luminescent Graphene Quantum Dots Using Self-Assembled Au Nanoparticles. ACS Omega 2023, 8, 5885–5892. [Google Scholar] [CrossRef]
  152. Lin, L.; Zhang, S. Creating high yield water soluble luminescent graphene quantum dots via exfoliating and disintegrating carbon nanotubes and graphite flakes. Chem. Commun. 2012, 48, 10177–10179. [Google Scholar] [CrossRef]
  153. Yan, X.; Li, B.; Cui, X.; Wei, Q.; Tajima, K.; Li, L.-S. Independent tuning of the band gap and redox potential of graphene quantum dots. J. Phys. Chem. Lett. 2011, 2, 1119–1124. [Google Scholar] [CrossRef]
  154. Mocci, F.; de Villiers Engelbrecht, L.; Olla, C.; Cappai, A.; Casula, M.F.; Melis, C.; Stagi, L.; Laaksonen, A.; Carbonaro, C.M. Carbon Nanodots from an In Silico Perspective. Chem. Rev. 2022, 122, 13709–13799. [Google Scholar] [CrossRef]
  155. Qiang, R.; Sun, W.; Hou, K.; Li, Z.; Zhang, J.; Ding, Y.; Wang, J.; Yang, S. Electrochemical trimming of graphene oxide affords graphene quantum dots for fe3+ detection. ACS Appl. Nano Mater. 2021, 4, 5220–5229. [Google Scholar] [CrossRef]
  156. Zhang, M.; Bai, L.; Shang, W.; Xie, W.; Ma, H.; Fu, Y.; Fang, D.; Sun, H.; Fan, L.; Han, M.; et al. Facile synthesis of water-soluble, highly fluorescent graphene quantum dots as a robust biological label for stem cells. J. Mater. Chem. 2012, 22, 7461. [Google Scholar] [CrossRef]
  157. Tetsuka, H.; Asahi, R.; Nagoya, A.; Okamoto, K.; Tajima, I.; Ohta, R.; Okamoto, A. Optically tunable amino-functionalized graphene quantum dots. Adv. Mater. 2012, 24, 5333–5338. [Google Scholar] [CrossRef]
  158. Das, S.K.; Luk, C.M.; Martin, W.E.; Tang, L.; Kim, D.Y.; Lau, S.P.; Richards, C.I. Size and dopant dependent single particle fluorescence properties of graphene quantum dots. J. Phys. Chem. C 2015, 119, 17988–17994. [Google Scholar] [CrossRef]
  159. Liu, Z.; Li, F.; Luo, Y.; Li, M.; Hu, G.; Pu, X.; Tang, T.; Wen, J.; Li, X.; Li, W. Size effect of graphene quantum dots on photoluminescence. Molecules 2021, 26, 3922. [Google Scholar] [CrossRef]
  160. Choppadandi, M.; Guduru, A.T.; Gondaliya, P.; Arya, N.; Kalia, K.; Kumar, H.; Kapusetti, G. Structural features regulated photoluminescence intensity and cell internalization of carbon and graphene quantum dots for bioimaging. Mater. Sci. Eng. C Mater. Biol. Appl. 2021, 129, 112366. [Google Scholar] [CrossRef]
  161. Zeng, Z.; Li, W.; Li, Y.; Zhao, Z.; Lin, K.; Fan, X.; Zhu, M. Lubrication behavior of fluorescent graphene quantum dots hybrid polyethylene glycol lubricant. Appl. Surf. Sci. 2023, 612, 155933. [Google Scholar] [CrossRef]
  162. Saleem, H.; Saud, A.; Zaidi, S.J. Sustainable Preparation of Graphene Quantum Dots from Leaves of Date Palm Tree. ACS Omega 2023, 8, 28098–28108. [Google Scholar] [CrossRef]
  163. Jiang, D.; Chen, Y.; Li, N.; Li, W.; Wang, Z.; Zhu, J.; Zhang, H.; Liu, B.; Xu, S. Synthesis of Luminescent Graphene Quantum Dots with High Quantum Yield and Their Toxicity Study. PLoS ONE 2015, 10, e0144906. [Google Scholar] [CrossRef]
  164. Sun, J.; Yang, S.; Wang, Z.; Shen, H.; Xu, T.; Sun, L.; Li, H.; Chen, W.; Jiang, X.; Ding, G.; et al. Ultra-High Quantum Yield of Graphene Quantum Dots: Aromatic-Nitrogen Doping and Photoluminescence Mechanism. Part. Part. Syst. Charact. 2015, 32, 434–440. [Google Scholar] [CrossRef]
  165. Shen, J.; Zhu, Y.; Yang, X.; Zong, J.; Zhang, J.; Li, C. One-pot hydrothermal synthesis of graphenequantum dots surface-passivated by polyethylene glycol and their photoelectric conversion under near-infrared light. New J. Chem 2012, 36, 97–101. [Google Scholar] [CrossRef]
  166. Das, R.; Parveen, S.; Bora, A.; Giri, P.K. Origin of high photoluminescence yield and high SERS sensitivity of nitrogen-doped graphene quantum dots. Carbon 2020, 160, 273–286. [Google Scholar] [CrossRef]
  167. Ding, H.; Yu, S.-B.; Wei, J.-S.; Xiong, H.-M. Full-Color Light-Emitting Carbon Dots with a Surface-State-Controlled Luminescence Mechanism. ACS Nano 2016, 10, 484–491. [Google Scholar] [CrossRef]
  168. Siddique, A.B.; Hossain, S.M.; Pramanick, A.K.; Ray, M. Excitation dependence and independence of photoluminescence in carbon dots and graphene quantum dots: Insights into the mechanism of emission. Nanoscale 2021, 13, 16662–16671. [Google Scholar] [CrossRef]
  169. Teng, C.-Y.; Nguyen, B.-S.; Yeh, T.-F.; Lee, Y.-L.; Chen, S.-J.; Teng, H. Roles of nitrogen functionalities in enhancing the excitation-independent green-color photoluminescence of graphene oxide dots. Nanoscale 2017, 9, 8256–8265. [Google Scholar] [CrossRef]
  170. Xu, Y.; Liu, L.; Xu, M.; Feng, W.; Wang, L.; Shan, X. Single-phase excitation-wavelength-independent nitrogen doped graphene quantum dots for the fabrication of white light-emitting diodes (WLEDs). New J. Chem. 2023, 47, 17339–17345. [Google Scholar] [CrossRef]
  171. Lee, J.-W.; Kwak, J.H.; Kim, J.; Jang, Y.-K.; Han, J.T.; Kim, T.-J.; Hong, K.-S.; Jeong, H.J.; Yang, I.H.-S. Highly emissive blue graphene quantum dots with excitation-independent emission via ultrafast liquid-phase photoreduction. RSC Adv. 2024, 14, 11524–11532. [Google Scholar] [CrossRef]
  172. Ruiyi, L.; Xie, Q.; Wang, Y.; Li, Z.; Liu, X. Synthesis of folic acid, histidine, and serine-functionalized and boron and phosphorus-doped graphene quantum dots with excellent yellow luminescence behavior in aqueous/solid states and their use for the fluorescence detection of Fe2+ in urine. New J. Chem. 2024, 48, 14984–14994. [Google Scholar] [CrossRef]
  173. Wang, B.; Lu, S. The light of carbon dots: From mechanism to applications. Matter 2022, 5, 110–149. [Google Scholar] [CrossRef]
  174. Lee, B.; Stokes, G.A.; Valimukhametova, A.; Nguyen, S.; Gonzalez-Rodriguez, R.; Bhaloo, A.; Coffer, J.; Naumov, A.V. Automated Approach to In Vitro Image-Guided Photothermal Therapy with Top-Down and Bottom-Up-Synthesized Graphene Quantum Dots. Nanomaterials 2023, 13, 805. [Google Scholar] [CrossRef]
  175. Robinson, J.T.; Tabakman, S.M.; Liang, Y.; Wang, H.; Casalongue, H.S.; Vinh, D.; Dai, H. Ultrasmall reduced graphene oxide with high near-infrared absorbance for photothermal therapy. J. Am. Chem. Soc. 2011, 133, 6825–6831. [Google Scholar] [CrossRef] [PubMed]
  176. Song, C.; Wang, Z.; Yin, Z.; Xiao, D.; Ma, D. Principles and applications of photothermal catalysis. Chem. Catal. 2022, 2, 52–83. [Google Scholar] [CrossRef]
  177. Ge, J.; Lan, M.; Zhou, B.; Liu, W.; Guo, L.; Wang, H.; Jia, Q.; Niu, G.; Huang, X.; Zhou, H.; et al. A graphene quantum dot photodynamic therapy agent with high singlet oxygen generation. Nat. Commun. 2014, 5, 4596. [Google Scholar] [CrossRef] [PubMed]
  178. Markovic, Z.M.; Ristic, B.Z.; Arsikin, K.M.; Klisic, D.G.; Harhaji-Trajkovic, L.M.; Todorovic-Markovic, B.M.; Kepic, D.P.; Kravic-Stevovic, T.K.; Jovanovic, S.P.; Milenkovic, M.M.; et al. Graphene quantum dots as autophagy-inducing photodynamic agents. Biomaterials 2012, 33, 7084–7092. [Google Scholar] [CrossRef] [PubMed]
  179. Ahirwar, S.; Mallick, S.; Bahadur, D. Photodynamic therapy using graphene quantum dot derivatives. J. Solid State Chem. 2020, 282, 121107. [Google Scholar] [CrossRef]
  180. Wang, Y.; Kong, W.; Wang, L.; Zhang, J.Z.; Li, Y.; Liu, X.; Li, Y. Optimizing oxygen functional groups in graphene quantum dots for improved antioxidant mechanism. Phys. Chem. Chem. Phys. 2019, 21, 1336–1343. [Google Scholar] [CrossRef]
  181. Hu, X.; Ma, X.-Y.; Tian, J.; Huang, Z. Rapid and facile synthesis of graphene quantum dots with high antioxidant activity. Inorg. Chem. Commun. 2020, 122, 108288. [Google Scholar] [CrossRef]
  182. Zhao, L.; Wang, Y.; Li, Y. Antioxidant activity of graphene quantum dots prepared in different electrolyte environments. Nanomaterials 2019, 9, 1708. [Google Scholar] [CrossRef]
  183. Cho, H.-H.; Yang, H.; Kang, D.J.; Kim, B.J. Surface engineering of graphene quantum dots and their applications as efficient surfactants. ACS Appl. Mater. Interfaces 2015, 7, 8615–8621. [Google Scholar] [CrossRef]
  184. Mehrdad-Vahdati, B.; Pourhashem, S.; Sedghi, M.; Vaezi, Z.; Shojaedin-Givi, B.; Rashidi, A.; Naderi-Manesh, H. A novel aspect of functionalized graphene quantum dots in cytotoxicity studies. Toxicol. Vitr. 2019, 61, 104649. [Google Scholar] [CrossRef]
  185. Zhang, S.; Sui, L.; Dong, H.; He, W.; Dong, L.; Yu, L. High-Performance Supercapacitor of Graphene Quantum Dots with Uniform Sizes. ACS Appl. Mater. Interfaces 2018, 10, 12983–12991. [Google Scholar] [CrossRef] [PubMed]
  186. Liu, Q.; Sun, J.; Gao, K.; Chen, N.; Sun, X.; Ti, D.; Bai, C.; Cui, R.; Qu, L. Graphene quantum dots for energy storage and conversion: From fabrication to applications. Mater. Chem. Front. 2020, 4, 421–436. [Google Scholar] [CrossRef]
  187. Liu, W.; Li, M.; Jiang, G.; Li, G.; Zhu, J.; Xiao, M.; Zhu, Y.; Gao, R.; Yu, A.; Feng, M.; et al. Graphene quantum dots-based advanced electrode materials: Design, synthesis and their applications in electrochemical energy storage and electrocatalysis. Adv. Energy Mater. 2020, 10, 2001275. [Google Scholar] [CrossRef]
  188. Payami, E.; Teimuri-Mofrad, R. Development of ternary nanocomposite based on ferrocenyl-modified graphene quantum dots for high-performance energy storage applications. J. Energy Storage 2023, 72, 108346. [Google Scholar] [CrossRef]
  189. Shaker, M.; Ng, S.; Sadeghi Ghazvini, A.A.; Javanmardi, S.; Gaho, M.A.; Jin, Z.; Ge, Q. Carbon/graphene quantum dots as electrolyte additives for batteries and supercapacitors: A review. J. Energy Storage 2024, 85, 111040. [Google Scholar] [CrossRef]
  190. Ochi, Y.; Otani, A.; Katakami, R.; Ogura, A.; Takao, K.; Iso, Y.; Isobe, T. Open system massive synthesis of narrow-band blue and green fluorescent graphene quantum dots and their application to water sensing. J. Mater. Chem. C 2024, 12, 6548–6558. [Google Scholar] [CrossRef]
  191. Li, X.; Lin, G.; Zhou, L.; Prosser, O.; Malakooti, M.H.; Zhang, M. Green synthesis of iron-doped graphene quantum dots: An efficient nanozyme for glucose sensing. Nanoscale Horiz. 2024, 9, 976–989. [Google Scholar] [CrossRef]
  192. Iannazzo, D.; Celesti, C.; Giofrè, S.V.; Ettari, R.; Bitto, A. Theranostic Applications of 2D Graphene-Based Materials for Solid Tumors Treatment. Nanomaterials 2023, 13, 2380. [Google Scholar] [CrossRef]
  193. Liang, J.; Liu, J.; Jin, X.; Yao, S.; Chen, B.; Huang, Q.; Hu, J.; Wan, J.; Hu, Z.; Wang, B. Versatile Nanoplatform Loaded with Doxorubicin and Graphene Quantum Dots/Methylene Blue for Drug Delivery and Chemophotothermal/Photodynamic Synergetic Cancer Therapy. ACS Appl. Bio Mater. 2020, 3, 7122–7132. [Google Scholar] [CrossRef]
  194. Wang, C.; Chen, Y.; Xu, Z.; Chen, B.; Zhang, Y.; Yi, X.; Li, J. Fabrication and characterization of novel cRGD modified graphene quantum dots for chemo-photothermal combination therapy. Sens. Actuators B Chem. 2020, 309, 127732. [Google Scholar] [CrossRef]
  195. Xu, Y.; Hu, X.; Guan, P.; Du, C.; Tian, Y.; Ding, S.; Li, Z.; Yan, C. A novel controllable molecularly imprinted drug delivery system based on the photothermal effect of graphene oxide quantum dots. J. Mater. Sci. 2019, 54, 9124–9139. [Google Scholar] [CrossRef]
  196. Nene, L.C.; Nyokong, T. Phthalocyanines and graphene quantum dots nano-systems as dual anti-cancer sensitizers for photo-sonodynamic combinatorial therapy. Diam. Relat. Mater. 2023, 131, 109549. [Google Scholar] [CrossRef]
  197. Lin, J.; Lin, J.-H.; Yeh, T.-Y.; Zheng, J.-H.; Cho, E.-C.; Lee, K.-C. Fabrication of hyaluronic acid with graphene quantum dot as a dual drug delivery system for cancer therapy. FlatChem 2024, 44, 100607. [Google Scholar] [CrossRef]
  198. Kansara, V.; Shukla, R.; Flora, S.J.S.; Bahadur, P.; Tiwari, S. Graphene quantum dots: Synthesis, optical properties and navigational applications against cancer. Mater. Today Commun. 2022, 31, 103359. [Google Scholar] [CrossRef]
  199. Khan, M.S.; Sheikh, A.; Abourehab, M.A.S.; Gupta, N.; Kesharwani, P. Understanding the theranostic potential of quantum dots in cancer management. Mater. Today Commun. 2023, 36, 106424. [Google Scholar] [CrossRef]
  200. Roy, P.; Periasamy, A.P.; Lin, C.-Y.; Her, G.-M.; Chiu, W.-J.; Li, C.-L.; Shu, C.-L.; Huang, C.-C.; Liang, C.-T.; Chang, H.-T. Photoluminescent graphene quantum dots for in vivo imaging of apoptotic cells. Nanoscale 2015, 7, 2504–2510. [Google Scholar] [CrossRef]
  201. Kim, J.; Song, S.H.; Jin, Y.; Park, H.-J.; Yoon, H.; Jeon, S.; Cho, S.-W. Multiphoton luminescent graphene quantum dots for in vivo tracking of human adipose-derived stem cells. Nanoscale 2016, 8, 8512–8519. [Google Scholar] [CrossRef]
  202. Narasimhan, A.K.; Lakshmi, B.S.; Santra, T.S.; Rao, M.S.R.; Krishnamurthi, G. Oxygenated graphene quantum dots (GQDs) synthesized using laser ablation for long-term real-time tracking and imaging. RSC Adv. 2017, 7, 53822–53829. [Google Scholar] [CrossRef]
  203. Gómez, I.J.; Ovejero-Paredes, K.; Méndez-Arriaga, J.M.; Pizúrová, N.; Filice, M.; Zajíčková, L.; Prashar, S.; Gómez-Ruiz, S. Organotin(IV)-Decorated Graphene Quantum Dots as Dual Platform for Molecular Imaging and Treatment of Triple Negative Breast Cancer. Chem. Eur. J 2023, 29, e202301845. [Google Scholar] [CrossRef]
  204. Shen, J.; Zhu, Y.; Chen, C.; Yang, X.; Li, C. Facile preparation and upconversion luminescence of graphene quantum dots. Chem. Commun. 2011, 47, 2580–2582. [Google Scholar] [CrossRef]
  205. Zhu, S.; Zhang, J.; Liu, X.; Li, B.; Wang, X.; Tang, S.; Meng, Q.; Li, Y.; Shi, C.; Hu, R.; et al. Graphene quantum dots with controllable surface oxidation, tunable fluorescence and up-conversion emission. RSC Adv. 2012, 2, 2717. [Google Scholar] [CrossRef]
  206. Sarkar, S.; Gandla, D.; Venkatesh, Y.; Bangal, P.R.; Ghosh, S.; Yang, Y.; Misra, S. Graphene quantum dots from graphite by liquid exfoliation showing excitation-independent emission, fluorescence upconversion and delayed fluorescence. Phys. Chem. Chem. Phys. 2016, 18, 21278–21287. [Google Scholar] [CrossRef] [PubMed]
  207. Wen, X.; Yu, P.; Toh, Y.-R.; Ma, X.; Tang, J. On the upconversion fluorescence in carbon nanodots and graphene quantum dots. Chem. Commun. 2014, 50, 4703–4706. [Google Scholar] [CrossRef]
  208. Kuo, W.-S.; Shen, X.-C.; Chang, C.-Y.; Kao, H.-F.; Lin, S.-H.; Wang, J.-Y.; Wu, P.-C. Multiplexed Graphene Quantum Dots with Excitation-Wavelength-Independent Photoluminescence, as Two-Photon Probes, and in Ultraviolet-Near Infrared Bioimaging. ACS Nano 2020, 14, 11502–11509. [Google Scholar] [CrossRef]
  209. Saeidi, S.; Rezaei, B.; Irannejad, N.; Ensafi, A.A. Efficiency improvement of luminescent solar concentrators using upconversion nitrogen-doped graphene quantum dots. J. Power Sources 2020, 476, 228647. [Google Scholar] [CrossRef]
  210. Zhu, S.; Zhang, J.; Tang, S.; Qiao, C.; Wang, L.; Wang, H.; Liu, X.; Li, B.; Li, Y.; Yu, W.; et al. Surface Chemistry Routes to Modulate the Photoluminescence of Graphene Quantum Dots: From Fluorescence Mechanism to Up-Conversion Bioimaging Applications. Adv. Funct. Mater 2012, 22, 4732–4740. [Google Scholar] [CrossRef]
  211. Wang, Z.; Zeng, H.; Sun, L. Graphene quantum dots: Versatile photoluminescence for energy, biomedical, and environmental applications. J. Mater. Chem. C 2015, 3, 1157–1165. [Google Scholar] [CrossRef]
  212. Vatanparast, M.; Shariatinia, Z. AlN and AlP doped graphene quantum dots as novel drug delivery systems for 5-fluorouracil drug: Theoretical studies. J. Fluor. Chem. 2018, 211, 81–93. [Google Scholar] [CrossRef]
  213. Mohammed-Ahmed, H.K.; Nakipoglu, M.; Tezcaner, A.; Keskin, D.; Evis, Z. Functionalization of graphene oxide quantum dots for anticancer drug delivery. J. Drug Deliv. Sci. Technol. 2023, 80, 104199. [Google Scholar] [CrossRef]
  214. Ahn, M.; Song, J.; Hong, B.H. Facile Synthesis of N-Doped Graphene Quantum Dots as Novel Transfection Agents for mRNA and pDNA. Nanomaterials 2021, 11, 2816. [Google Scholar] [CrossRef]
  215. Karami, M.H.; Abdouss, M.; Rahdar, A.; Pandey, S. Graphene quantum dots: Background, synthesis methods, and applications as nanocarrier in drug delivery and cancer treatment: An updated review. Inorg. Chem. Commun. 2024, 161, 112032. [Google Scholar] [CrossRef]
  216. Yu, C.; Long, Z.; Qiu, Q.; Liu, F.; Xu, Y.; Zhang, T.; Guo, R.; Zhong, W.; Huang, S.; Chen, S. Graphene quantum dots-based targeted nanoprobes detecting drug delivery, imaging, and enhanced chemotherapy of nasopharyngeal carcinoma. Bioeng. Transl. Med. 2022, 7, e10270. [Google Scholar] [CrossRef] [PubMed]
  217. Sawy, A.M.; Barhoum, A.; Abdel Gaber, S.A.; El-Hallouty, S.M.; Shousha, W.G.; Maarouf, A.A.; Khalil, A.S.G. Insights of doxorubicin loaded graphene quantum dots: Synthesis, DFT drug interactions, and cytotoxicity. Mater. Sci. Eng. C Mater. Biol. Appl. 2021, 122, 111921. [Google Scholar] [CrossRef]
  218. Li, Z.; Fan, J.; Tong, C.; Zhou, H.; Wang, W.; Li, B.; Liu, B.; Wang, W. A smart drug-delivery nanosystem based on carboxylated graphene quantum dots for tumor-targeted chemotherapy. Nanomedicine 2019, 14, 2011–2025. [Google Scholar] [CrossRef]
  219. Nasrollahi, F.; Koh, Y.R.; Chen, P.; Varshosaz, J.; Khodadadi, A.A.; Lim, S. Targeting graphene quantum dots to epidermal growth factor receptor for delivery of cisplatin and cellular imaging. Mater. Sci. Eng. C Mater. Biol. Appl. 2019, 94, 247–257. [Google Scholar] [CrossRef]
  220. Vatanparast, M.; Shariatinia, Z. Computational studies on the doped graphene quantum dots as potential carriers in drug delivery systems for isoniazid drug. Struct. Chem. 2018, 29, 1427–1448. [Google Scholar] [CrossRef]
  221. Wei, Z.; Yin, X.; Cai, Y.; Xu, W.; Song, C.; Wang, Y.; Zhang, J.; Kang, A.; Wang, Z.; Han, W. Antitumor effect of a Pt-loaded nanocomposite based on graphene quantum dots combats hypoxia-induced chemoresistance of oral squamous cell carcinoma. Int. J. Nanomed. 2018, 13, 1505–1524. [Google Scholar] [CrossRef]
  222. Sheng, Y.; Dai, W.; Gao, J.; Li, H.; Tan, W.; Wang, J.; Deng, L.; Kong, Y. pH-sensitive drug delivery based on chitosan wrapped graphene quantum dots with enhanced fluorescent stability. Mater. Sci. Eng. C Mater. Biol. Appl. 2020, 112, 110888. [Google Scholar] [CrossRef]
  223. Iannazzo, D.; Pistone, A.; Celesti, C.; Triolo, C.; Patané, S.; Giofré, S.V.; Romeo, R.; Ziccarelli, I.; Mancuso, R.; Gabriele, B.; et al. A smart nanovector for cancer targeted drug delivery based on graphene quantum dots. Nanomaterials 2019, 9, 282. [Google Scholar] [CrossRef]
  224. Amraee, N.; Majd, A.; Torbati, M.B.; Shaabanzadeh, M. Folic Acid-decorated and PEGylated Graphene Quantum Dots as efficient Tamoxifen delivery system against breast cancer cells: In vitro studies. Res. Sq. 2023. [Google Scholar] [CrossRef]
  225. Jeong, S.; Pinals, R.L.; Dharmadhikari, B.; Song, H.; Kalluri, A.; Debnath, D.; Wu, Q.; Ham, M.-H.; Patra, P.; Landry, M.P. Graphene quantum dot oxidation governs noncovalent biopolymer adsorption. Sci. Rep. 2020, 10, 7074. [Google Scholar] [CrossRef] [PubMed]
  226. Lu, L.; Guo, L.; Wang, X.; Kang, T.; Cheng, S. Complexation and intercalation modes: A novel interaction of DNA and graphene quantum dots. RSC Adv. 2016, 6, 33072–33075. [Google Scholar] [CrossRef]
  227. Ratre, P.; Jain, B.; Kumari, R.; Thareja, S.; Tiwari, R.; Srivastava, R.K.; Goryacheva, I.Y.; Mishra, P.K. Bioanalytical Applications of Graphene Quantum Dots for Circulating Cell-Free Nucleic Acids: A Review. ACS Omega 2022, 7, 39586–39602. [Google Scholar] [CrossRef]
  228. Lee, G.-Y.; Lo, P.-Y.; Cho, E.-C.; Zheng, J.-H.; Li, M.; Huang, J.-H.; Lee, K.-C. Integration of PEG and PEI with graphene quantum dots to fabricate pH-responsive nanostars for colon cancer suppression in vitro and in vivo. FlatChem 2022, 31, 100320. [Google Scholar] [CrossRef]
  229. Bogdanov, A.; Bogdanov, A.; Chubenko, V.; Volkov, N.; Moiseenko, F.; Moiseyenko, V. Tumor acidity: From hallmark of cancer to target of treatment. Front. Oncol. 2022, 12, 979154. [Google Scholar] [CrossRef]
  230. Qiu, J.; Zhang, R.; Li, J.; Sang, Y.; Tang, W.; Rivera Gil, P.; Liu, H. Fluorescent graphene quantum dots as traceable, pH-sensitive drug delivery systems. Int. J. Nanomed. 2015, 10, 6709–6724. [Google Scholar] [CrossRef]
  231. Yang, Y.; Qin, Z.; Zeng, W.; Yang, T.; Cao, Y.; Mei, C.; Kuang, Y. Toxicity assessment of nanoparticles in various systems and organs. Nanotechnol. Rev. 2017, 6, 279–289. [Google Scholar] [CrossRef]
  232. Shvedova, A.; Pietroiusti, A.; Kagan, V. Nanotoxicology ten years later: Lights and shadows. Toxicol. Appl. Pharmacol. 2016, 299, 1–2. [Google Scholar] [CrossRef]
  233. González-Muñoz, M.; Díez, P.; González-González, M.; Dégano, R.M.; Ibarrola, N.; Orfao, A.; Fuentes, M. Evaluation strategies of nanomaterials toxicity. In Nanomaterials—Toxicity and Risk Assessment; Larramendy, M.L., Soloneski, S., Eds.; InTech: Houston TX, USA, 2015; ISBN 978-953-51-2143-5. [Google Scholar]
  234. Arathi; Megha, K.B.; Joseph, X.; Mohanan, P.V. Biological safety and cellular interactions of nanoparticles. In Biomedical Applications and Toxicity of Nanomaterials; Mohanan, P.V., Kappalli, S., Eds.; Springer Nature Singapore: Singapore, 2023; pp. 559–587. ISBN 978-981-19-7833-3. [Google Scholar]
  235. Behzadi, S.; Serpooshan, V.; Tao, W.; Hamaly, M.A.; Alkawareek, M.Y.; Dreaden, E.C.; Brown, D.; Alkilany, A.M.; Farokhzad, O.C.; Mahmoudi, M. Cellular uptake of nanoparticles: Journey inside the cell. Chem. Soc. Rev. 2017, 46, 4218–4244. [Google Scholar] [CrossRef]
  236. Sangam, S.; Gupta, A.; Shakeel, A.; Bhattacharya, R.; Sharma, A.K.; Suhag, D.; Chakrabarti, S.; Garg, S.K.; Chattopadhyay, S.; Basu, B.; et al. Sustainable synthesis of single crystalline sulphur-doped graphene quantum dots for bioimaging and beyond. Green Chem. 2018, 20, 4245–4259. [Google Scholar] [CrossRef]
  237. Yan, C.; Hu, X.; Guan, P.; Hou, T.; Chen, P.; Wan, D.; Zhang, X.; Wang, J.; Wang, C. Highly biocompatible graphene quantum dots: Green synthesis, toxicity comparison and fluorescence imaging. J. Mater. Sci. 2020, 55, 1198–1215. [Google Scholar] [CrossRef]
  238. Mathew, A.A.; Varghese, M.; Balachandran, M. Biosafety and toxicity evaluation of carbon nanomaterials. In Carbon Nanostructures in Biomedical Applications; Hasnain, M.S., Nayak, A.K., Alkahtani, S., Eds.; Advances in Material Research and Technology; Springer International Publishing: Cham, Switzerland, 2023; pp. 363–398. ISBN 978-3-031-28262-1. [Google Scholar]
  239. Raja, I.S.; Molkenova, A.; Kang, M.S.; Lee, S.H.; Lee, J.E.; Kim, B.; Han, D.-W.; Atabaev, T.S. Differential toxicity of graphene family nanomaterials concerning morphology. Adv. Exp. Med. Biol. 2022, 1351, 23–39. [Google Scholar] [CrossRef] [PubMed]
  240. Xiaoli, F.; Qiyue, C.; Weihong, G.; Yaqing, Z.; Chen, H.; Junrong, W.; Longquan, S. Toxicology data of graphene-family nanomaterials: An update. Arch. Toxicol. 2020, 94, 1915–1939. [Google Scholar] [CrossRef] [PubMed]
  241. Frieler, M.; Pho, C.; Lee, B.H.; Dobrovolny, H.; Akkaraju, G.R.; Naumov, A.V. Effects of Doxorubicin Delivery by Nitrogen-Doped Graphene Quantum Dots on Cancer Cell Growth: Experimental Study and Mathematical Modeling. Nanomaterials 2021, 11, 140. [Google Scholar] [CrossRef] [PubMed]
  242. Qi, L.; Pan, T.; Ou, L.; Ye, Z.; Yu, C.; Bao, B.; Wu, Z.; Cao, D.; Dai, L. Biocompatible nucleus-targeted graphene quantum dots for selective killing of cancer cells via DNA damage. Commun. Biol. 2021, 4, 214. [Google Scholar] [CrossRef]
  243. Liang, L.; Peng, X.; Sun, F.; Kong, Z.; Shen, J.-W. A review on the cytotoxicity of graphene quantum dots: From experiment to simulation. Nanoscale Adv. 2021, 3, 904–917. [Google Scholar] [CrossRef]
  244. Wang, X.-Y.; Lei, R.; Huang, H.-D.; Wang, N.; Yuan, L.; Xiao, R.-Y.; Bai, L.-D.; Li, X.; Li, L.-M.; Yang, X. The permeability and transport mechanism of graphene quantum dots (GQDs) across the biological barrier. Nanoscale 2015, 7, 2034–2041. [Google Scholar] [CrossRef]
  245. Sheikh Mohd Ghazali, S.A.I.; Fatimah, I.; Zamil, Z.N.; Zulkifli, N.N.; Adam, N. Graphene quantum dots: A comprehensive overview. Open Chem. 2023, 21, 20220285. [Google Scholar] [CrossRef]
  246. Perini, G.; Palmieri, V.; Ciasca, G.; De Spirito, M.; Papi, M. Unravelling the potential of graphene quantum dots in biomedicine and neuroscience. Int. J. Mol. Sci. 2020, 21, 3712. [Google Scholar] [CrossRef]
  247. Ku, T.-H.; Shen, W.-T.; Hsieh, C.-T.; Chen, G.S.; Shia, W.-C. Specific forms of graphene quantum dots induce apoptosis and cell cycle arrest in breast cancer cells. Int. J. Mol. Sci. 2023, 24, 4046. [Google Scholar] [CrossRef]
  248. Wang, D.; Zhu, L.; Chen, J.-F.; Dai, L. Can graphene quantum dots cause DNA damage in cells? Nanoscale 2015, 7, 9894–9901. [Google Scholar] [CrossRef] [PubMed]
  249. de Menezes, F.D.; Dos Reis, S.R.R.; Pinto, S.R.; Portilho, F.L.; do Vale Chaves E Mello, F.; Helal-Neto, E.; da Silva de Barros, A.O.; Alencar, L.M.R.; de Menezes, A.S.; Dos Santos, C.C.; et al. Graphene quantum dots unraveling: Green synthesis, characterization, radiolabeling with 99mTc, in vivo behavior and mutagenicity. Mater. Sci. Eng. C Mater. Biol. Appl. 2019, 102, 405–414. [Google Scholar] [CrossRef] [PubMed]
  250. Xie, Y.; Wan, B.; Yang, Y.; Cui, X.; Xin, Y.; Guo, L.-H. Cytotoxicity and autophagy induction by graphene quantum dots with different functional groups. J. Environ. Sci. 2018, 77, 198–209. [Google Scholar] [CrossRef] [PubMed]
  251. Xu, L.; Dai, Y.; Wang, Z.; Zhao, J.; Li, F.; White, J.C.; Xing, B. Graphene quantum dots in alveolar macrophage: Uptake-exocytosis, accumulation in nuclei, nuclear responses and DNA cleavage. Part. Fibre Toxicol. 2018, 15, 45. [Google Scholar] [CrossRef]
  252. Yuan, X.; Liu, Z.; Guo, Z.; Ji, Y.; Jin, M.; Wang, X. Cellular distribution and cytotoxicity of graphene quantum dots with different functional groups. Nanoscale Res. Lett. 2014, 9, 108. [Google Scholar] [CrossRef]
  253. Liao, K.-H.; Lin, Y.-S.; Macosko, C.W.; Haynes, C.L. Cytotoxicity of graphene oxide and graphene in human erythrocytes and skin fibroblasts. ACS Appl. Mater. Interfaces 2011, 3, 2607–2615. [Google Scholar] [CrossRef]
  254. Schroeder, K.L.; Goreham, R.V.; Nann, T. Graphene quantum dots for theranostics and bioimaging. Pharm. Res. 2016, 33, 2337–2357. [Google Scholar] [CrossRef]
  255. Zhou, Y.; Sun, H.; Wang, F.; Ren, J.; Qu, X. How functional groups influence the ROS generation and cytotoxicity of graphene quantum dots. Chem. Commun. 2017, 53, 10588–10591. [Google Scholar] [CrossRef]
  256. Ren, C.; Hu, X.; Zhou, Q. Graphene Oxide Quantum Dots Reduce Oxidative Stress and Inhibit Neurotoxicity In Vitro and In Vivo through Catalase-like Activity and Metabolic Regulation. Adv. Sci. 2018, 5, 1700595. [Google Scholar] [CrossRef]
  257. Nurunnabi, M.; Khatun, Z.; Huh, K.M.; Park, S.Y.; Lee, D.Y.; Cho, K.J.; Lee, Y.-K. In vivo biodistribution and toxicology of carboxylated graphene quantum dots. ACS Nano 2013, 7, 6858–6867. [Google Scholar] [CrossRef]
  258. Wang, T.; Zhu, S.; Jiang, X. Toxicity mechanism of graphene oxide and nitrogen-doped graphene quantum dots in RBCs revealed by surface-enhanced infrared absorption spectroscopy. Toxicol. Res. 2015, 4, 885–894. [Google Scholar] [CrossRef]
  259. Fasbender, S.; Zimmermann, L.; Cadeddu, R.-P.; Luysberg, M.; Moll, B.; Janiak, C.; Heinzel, T.; Haas, R. The low toxicity of graphene quantum dots is reflected by marginal gene expression changes of primary human hematopoietic stem cells. Sci. Rep. 2019, 9, 12028. [Google Scholar] [CrossRef] [PubMed]
  260. Hai, X.; Mao, Q.-X.; Wang, W.-J.; Wang, X.-F.; Chen, X.-W.; Wang, J.-H. An acid-free microwave approach to prepare highly luminescent boron-doped graphene quantum dots for cell imaging. J. Mater. Chem. B 2015, 3, 9109–9114. [Google Scholar] [CrossRef]
  261. Li, B.; Xiao, X.; Hu, M.; Wang, Y.; Wang, Y.; Yan, X.; Huang, Z.; Servati, P.; Huang, L.; Tang, J. Mn, B, N co-doped graphene quantum dots for fluorescence sensing and biological imaging. Arab. J. Chem. 2022, 15, 103856. [Google Scholar] [CrossRef]
  262. Qu, C.; Zhang, D.; Yang, R.; Hu, J.; Qu, L. Nitrogen and sulfur co-doped graphene quantum dots for the highly sensitive and selective detection of mercury ion in living cells. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2019, 206, 588–596. [Google Scholar] [CrossRef]
  263. Ananthanarayanan, A.; Wang, Y.; Routh, P.; Sk, M.A.; Than, A.; Lin, M.; Zhang, J.; Chen, J.; Sun, H.; Chen, P. Nitrogen and phosphorus co-doped graphene quantum dots: Synthesis from adenosine triphosphate, optical properties, and cellular imaging. Nanoscale 2015, 7, 8159–8165. [Google Scholar] [CrossRef]
  264. Valimukhametova, A.R.; Lee, B.H.; Topkiran, U.C.; Gries, K.; Gonzalez-Rodriguez, R.; Coffer, J.L.; Akkaraju, G.; Naumov, A. Cancer Therapeutic siRNA Delivery and Imaging by Nitrogen- and Neodymium-Doped Graphene Quantum Dots. ACS Biomater. Sci. Eng. 2023, 9, 3425–3434. [Google Scholar] [CrossRef]
  265. Wang, H.; Mu, Q.; Wang, K.; Revia, R.A.; Yen, C.; Gu, X.; Tian, B.; Liu, J.; Zhang, M. Nitrogen and Boron Dual-Doped Graphene Quantum Dots for Near-Infrared Second Window Imaging and Photothermal Therapy. Appl. Mater. Today 2019, 14, 108–117. [Google Scholar] [CrossRef]
  266. Chandra, A.; Deshpande, S.; Shinde, D.B.; Pillai, V.K.; Singh, N. Mitigating the cytotoxicity of graphene quantum dots and enhancing their applications in bioimaging and drug delivery. ACS Macro Lett. 2014, 3, 1064–1068. [Google Scholar] [CrossRef]
  267. Martín, C.; Jun, G.; Schurhammer, R.; Reina, G.; Chen, P.; Bianco, A.; Ménard-Moyon, C. Enzymatic degradation of graphene quantum dots by human peroxidases. Small 2019, 15, e1905405. [Google Scholar] [CrossRef]
  268. Deng, S.; Zhang, E.; Wang, Y.; Zhao, Y.; Yang, Z.; Zheng, B.; Mu, X.; Deng, X.; Shen, H.; Rong, H.; et al. In vivo toxicity assessment of four types of graphene quantum dots (GQDs) using mRNA sequencing. Toxicol. Lett. 2022, 363, 55–66. [Google Scholar] [CrossRef] [PubMed]
  269. Huang, C.-L.; Huang, C.-C.; Mai, F.-D.; Yen, C.-L.; Tzing, S.-H.; Hsieh, H.-T.; Ling, Y.-C.; Chang, J.-Y. Application of paramagnetic graphene quantum dots as a platform for simultaneous dual-modality bioimaging and tumor-targeted drug delivery. J. Mater. Chem. B 2015, 3, 651–664. [Google Scholar] [CrossRef]
  270. Wang, Z.G.; Zhou, R.; Jiang, D.; Song, J.E.; Xu, Q.; Si, J.; Chen, Y.P.; Zhou, X.; Gan, L.; Li, J.Z.; et al. Toxicity of graphene quantum dots in zebrafish embryo. Biomed. Environ. Sci. 2015, 28, 341–351. [Google Scholar] [CrossRef]
  271. Zhang, J.-H.; Sun, T.; Niu, A.; Tang, Y.-M.; Deng, S.; Luo, W.; Xu, Q.; Wei, D.; Pei, D.-S. Perturbation effect of reduced graphene oxide quantum dots (rGOQDs) on aryl hydrocarbon receptor (AhR) pathway in zebrafish. Biomaterials 2017, 133, 49–59. [Google Scholar] [CrossRef]
  272. Fasbender, S.; Allani, S.; Wimmenauer, C.; Cadeddu, R.-P.; Raba, K.; Fischer, J.C.; Bulat, B.; Luysberg, M.; Seidel, C.A.M.; Heinzel, T.; et al. Uptake dynamics of graphene quantum dots into primary human blood cells following in vitro exposure. RSC Adv. 2017, 7, 12208–12216. [Google Scholar] [CrossRef]
  273. Qin, Y.; Zhou, Z.-W.; Pan, S.-T.; He, Z.-X.; Zhang, X.; Qiu, J.-X.; Duan, W.; Yang, T.; Zhou, S.-F. Graphene quantum dots induce apoptosis, autophagy, and inflammatory response via p38 mitogen-activated protein kinase and nuclear factor-κB mediated signaling pathways in activated THP-1 macrophages. Toxicology 2015, 327, 62–76. [Google Scholar] [CrossRef]
  274. Abdellatif, A.A.H.; Younis, M.A.; Alsharidah, M.; Al Rugaie, O.; Tawfeek, H.M. Biomedical applications of quantum dots: Overview, challenges, and clinical potential. Int. J. Nanomed. 2022, 17, 1951–1970. [Google Scholar] [CrossRef]
  275. Le, N.; Kim, K. Current advances in the biomedical applications of quantum dots: Promises and challenges. Int. J. Mol. Sci. 2023, 24, 12682. [Google Scholar] [CrossRef]
  276. Sui, X.; Luo, C.; Wang, C.; Zhang, F.; Zhang, J.; Guo, S. Graphene quantum dots enhance anticancer activity of cisplatin via increasing its cellular and nuclear uptake. Nanomedicine 2016, 12, 1997–2006. [Google Scholar] [CrossRef]
  277. Zheng, B.; Wang, C.; Xin, X.; Liu, F.; Zhou, X.; Zhang, J.; Guo, S. Electron Transfer from Graphene Quantum Dots to the Copper Complex Enhances Its Nuclease Activity. J. Phys. Chem. C 2014, 118, 7637–7642. [Google Scholar] [CrossRef]
  278. Perini, G.; Palmieri, V.; Ciasca, G.; D’Ascenzo, M.; Primiano, A.; Gervasoni, J.; De Maio, F.; De Spirito, M.; Papi, M. Enhanced chemotherapy for glioblastoma multiforme mediated by functionalized graphene quantum dots. Materials 2020, 13, 4139. [Google Scholar] [CrossRef] [PubMed]
  279. Sinha, R.; Chatterjee, A.; Purkayastha, P. Graphene Quantum Dot Assisted Translocation of Daunomycin through an Ordered Lipid Membrane: A Study by Fluorescence Lifetime Imaging Microscopy and Resonance Energy Transfer. J. Phys. Chem. B 2022, 126, 1232–1241. [Google Scholar] [CrossRef] [PubMed]
  280. Zhang, P.; Jiao, F.; Wu, L.; Kong, Z.; Hu, W.; Liang, L.; Zhang, Y. Molecular Dynamics Simulation of Transport Mechanism of Graphene Quantum Dots through Different Cell Membranes. Membranes 2022, 12, 753. [Google Scholar] [CrossRef]
  281. Biswas, M.C.; Islam, M.T.; Nandy, P.K.; Hossain, M.M. Graphene quantum dots (gqds) for bioimaging and drug delivery applications: A review. ACS Mater. Lett. 2021, 3, 889–911. [Google Scholar] [CrossRef]
  282. Barati, F.; Avatefi, M.; Moghadam, N.B.; Asghari, S.; Ekrami, E.; Mahmoudifard, M. A review of graphene quantum dots and their potential biomedical applications. J. Biomater. Appl. 2023, 37, 1137–1158. [Google Scholar] [CrossRef]
  283. Dong, J.; Wang, K.; Sun, L.; Sun, B.; Yang, M.; Chen, H.; Wang, Y.; Sun, J.; Dong, L. Application of graphene quantum dots for simultaneous fluorescence imaging and tumor-targeted drug delivery. Sens. Actuators B Chem. 2018, 256, 616–623. [Google Scholar] [CrossRef]
  284. George, D.; Suri, A.; Dutta, K.; Nayak, S. Targeted drug delivery using graphene quantum dots: Approaches, limitations and future perspectives. ECS Trans. 2022, 107, 16081–16098. [Google Scholar] [CrossRef]
  285. Kortel, M.; Mansuriya, B.D.; Vargas Santana, N.; Altintas, Z. Graphene Quantum Dots as Flourishing Nanomaterials for Bio-Imaging, Therapy Development, and Micro-Supercapacitors. Micromachines 2020, 11, 866. [Google Scholar] [CrossRef]
  286. Chen, Y.; Chu, M. Graphene quantum dots for drug delivery. In The World Scientific Encyclopedia of Nanomedicine and Bioengineering II: Bioimplants, Regenerative Medicine, and Nano-Cancer Diagnosis and PhototherapyVolume 1: Synthesis and Biomedical Applications of Graphene Quantum Dots; Frontiers in Nanobiomedical Research; World Scientific: Singapore, 2017; Volume 1, pp. 115–125. ISBN 978-981-320-255-9. [Google Scholar]
  287. Sharma, H.; Mondal, S. Functionalized graphene oxide for chemotherapeutic drug delivery and cancer treatment: A promising material in nanomedicine. Int. J. Mol. Sci. 2020, 21, 6280. [Google Scholar] [CrossRef]
  288. Nigam, P.; Waghmode, S.; Louis, M.; Wangnoo, S.; Chavan, P.; Sarkar, D. Graphene quantum dots conjugated albumin nanoparticles for targeted drug delivery and imaging of pancreatic cancer. J. Mater. Chem. B Mater. Biol. Med. 2014, 2, 3190–3195. [Google Scholar] [CrossRef]
  289. Habiba, K.; Encarnacion-Rosado, J.; Garcia-Pabon, K.; Villalobos-Santos, J.C.; Makarov, V.I.; Avalos, J.A.; Weiner, B.R.; Morell, G. Improving cytotoxicity against cancer cells by chemo-photodynamic combined modalities using silver-graphene quantum dots nanocomposites. Int. J. Nanomed. 2016, 11, 107–119. [Google Scholar] [CrossRef] [PubMed]
  290. Lo, P.-Y.; Lee, G.-Y.; Zheng, J.-H.; Huang, J.-H.; Cho, E.-C.; Lee, K.-C. GFP Plasmid and Chemoreagent Conjugated with Graphene Quantum Dots as a Novel Gene Delivery Platform for Colon Cancer Inhibition In Vitro and In Vivo. ACS Appl. Bio Mater. 2020, 3, 5948–5956. [Google Scholar] [CrossRef]
  291. Chen, J.; Wu, W.; Zhang, F.; Zhang, J.; Liu, H.; Zheng, J.; Guo, S.; Zhang, J. Graphene quantum dots in photodynamic therapy. Nanoscale Adv. 2020, 2, 4961–4967. [Google Scholar] [CrossRef]
  292. Bhaloo, A.; Nguyen, S.; Lee, B.H.; Valimukhametova, A.; Gonzalez-Rodriguez, R.; Sottile, O.; Dorsky, A.; Naumov, A.V. Doped graphene quantum dots as biocompatible radical scavenging agents. Antioxidants 2023, 12, 1536. [Google Scholar] [CrossRef]
  293. Song, S.; Qin, Y.; He, Y.; Huang, Q.; Fan, C.; Chen, H.-Y. Functional nanoprobes for ultrasensitive detection of biomolecules. Chem. Soc. Rev. 2010, 39, 4234–4243. [Google Scholar] [CrossRef]
  294. Iannazzo, D.; Pistone, A.; Salamò, M.; Galvagno, S.; Romeo, R.; Giofré, S.V.; Branca, C.; Visalli, G.; Di Pietro, A. Graphene quantum dots for cancer targeted drug delivery. Int. J. Pharm. 2017, 518, 185–192. [Google Scholar] [CrossRef]
  295. Davoodi, Z.; Shafiee, F. Internalizing RGD, a great motif for targeted peptide and protein delivery: A review article. Drug Deliv. Transl. Res. 2022, 12, 2261–2274. [Google Scholar] [CrossRef]
  296. Ding, H.; Zhang, F.; Zhao, C.; Lv, Y.; Ma, G.; Wei, W.; Tian, Z. Beyond a Carrier: Graphene Quantum Dots as a Probe for Programmatically Monitoring Anti-Cancer Drug Delivery, Release, and Response. ACS Appl. Mater. Interfaces 2017, 9, 27396–27401. [Google Scholar] [CrossRef]
  297. Cao, Y.; Dong, H.; Yang, Z.; Zhong, X.; Chen, Y.; Dai, W.; Zhang, X. Aptamer-Conjugated Graphene Quantum Dots/Porphyrin Derivative Theranostic Agent for Intracellular Cancer-Related MicroRNA Detection and Fluorescence-Guided Photothermal/Photodynamic Synergetic Therapy. ACS Appl. Mater. Interfaces 2017, 9, 159–166. [Google Scholar] [CrossRef]
  298. Correia, A.S.; Gärtner, F.; Vale, N. Drug combination and repurposing for cancer therapy: The example of breast cancer. Heliyon 2021, 7, e05948. [Google Scholar] [CrossRef]
  299. Wang, M.; Li, B.; Du, Y.; Bu, H.; Tang, Y.; Huang, Q. Fluorescence imaging-guided cancer photothermal therapy using polydopamine and graphene quantum dot-capped Prussian blue nanocubes. RSC Adv. 2021, 11, 8420–8429. [Google Scholar] [CrossRef] [PubMed]
  300. Yang, C.; Chan, K.K.; Xu, G.; Yin, M.; Lin, G.; Wang, X.; Lin, W.-J.; Birowosuto, M.D.; Zeng, S.; Ogi, T.; et al. Biodegradable Polymer-Coated Multifunctional Graphene Quantum Dots for Light-Triggered Synergetic Therapy of Pancreatic Cancer. ACS Appl. Mater. Interfaces 2019, 11, 2768–2781. [Google Scholar] [CrossRef] [PubMed]
  301. Ju, J.; Regmi, S.; Fu, A.; Lim, S.; Liu, Q. Graphene quantum dot based charge-reversal nanomaterial for nucleus-targeted drug delivery and efficiency controllable photodynamic therapy. J. Biophotonics 2019, 12, e201800367. [Google Scholar] [CrossRef] [PubMed]
  302. Li, S.; Zhou, S.; Li, Y.; Li, X.; Zhu, J.; Fan, L.; Yang, S. Exceptionally High Payload of the IR780 Iodide on Folic Acid-Functionalized Graphene Quantum Dots for Targeted Photothermal Therapy. ACS Appl. Mater. Interfaces 2017, 9, 22332–22341, Erratum in ACS Appl. Mater. Interfaces 2020, 12, 27819–27820. [Google Scholar] [CrossRef]
  303. Chen, L.; Hong, W.; Duan, S.; Li, Y.; Wang, J.; Zhu, J. Graphene quantum dots mediated magnetic chitosan drug delivery nanosystems for targeting synergistic photothermal-chemotherapy of hepatocellular carcinoma. Cancer Biol. Ther. 2022, 23, 281–293. [Google Scholar] [CrossRef]
  304. Rakhshaei, R.; Namazi, H.; Hamishehkar, H.; Rahimi, M. Graphene quantum dot cross-linked carboxymethyl cellulose nanocomposite hydrogel for pH-sensitive oral anticancer drug delivery with potential bioimaging properties. Int. J. Biol. Macromol. 2020, 150, 1121–1129. [Google Scholar] [CrossRef]
  305. Tian, Z.; Yao, X.; Ma, K.; Niu, X.; Grothe, J.; Xu, Q.; Liu, L.; Kaskel, S.; Zhu, Y. Metal-Organic Framework/Graphene Quantum Dot Nanoparticles Used for Synergistic Chemo- and Photothermal Therapy. ACS Omega 2017, 2, 1249–1258. [Google Scholar] [CrossRef]
  306. Zhou, L.; Zhou, L.; Ge, X.; Zhou, J.; Wei, S.; Shen, J. Multicolor imaging and the anticancer effect of a bifunctional silica nanosystem based on the complex of graphene quantum dots and hypocrellin A. Chem. Commun. 2015, 51, 421–424. [Google Scholar] [CrossRef]
  307. Flak, D.; Przysiecka, Ł.; Nowaczyk, G.; Scheibe, B.; Kościński, M.; Jesionowski, T.; Jurga, S. GQDs-MSNs nanocomposite nanoparticles for simultaneous intracellular drug delivery and fluorescent imaging. J. Nanopart. Res. 2018, 20, 306. [Google Scholar] [CrossRef]
  308. Teng, Y.; Yuan, S.; Shi, J.; Pong, P.W.T. A multifunctional nanoplatform based on graphene quantum dots-cobalt ferrite for monitoring of drug delivery and fluorescence/magnetic resonance bimodal cellular imaging. Adv. NanoBiomed Res. 2022, 2, 2200044. [Google Scholar] [CrossRef]
  309. Sung, S.-Y.; Su, Y.-L.; Cheng, W.; Hu, P.-F.; Chiang, C.-S.; Chen, W.-T.; Hu, S.-H. Graphene Quantum Dots-Mediated Theranostic Penetrative Delivery of Drug and Photolytics in Deep Tumors by Targeted Biomimetic Nanosponges. Nano Lett. 2019, 19, 69–81. [Google Scholar] [CrossRef] [PubMed]
  310. Wo, F.; Xu, R.; Shao, Y.; Zhang, Z.; Chu, M.; Shi, D.; Liu, S. A Multimodal System with Synergistic Effects of Magneto-Mechanical, Photothermal, Photodynamic and Chemo Therapies of Cancer in Graphene-Quantum Dot-Coated Hollow Magnetic Nanospheres. Theranostics 2016, 6, 485–500. [Google Scholar] [CrossRef] [PubMed]
  311. Wang, C.; Wu, C.; Zhou, X.; Han, T.; Xin, X.; Wu, J.; Zhang, J.; Guo, S. Enhancing cell nucleus accumulation and DNA cleavage activity of anti-cancer drug via graphene quantum dots. Sci. Rep. 2013, 3, 2852. [Google Scholar] [CrossRef] [PubMed]
  312. Prabhakar, A.K.; Ajith, M.P.; Ananthanarayanan, A.; Routh, P.; Mohan, B.C.; Thamizhchelvan, A.M. Ball-milled graphene quantum dots for enhanced anti-cancer drug delivery. OpenNano 2022, 8, 100072. [Google Scholar] [CrossRef]
  313. Dong, J.; Yao, X.; Sun, S.; Zhong, Y.; Qian, C.; Yang, D. In vivo targeting of breast cancer with a vasculature-specific GQDs/hMSN nanoplatform. RSC Adv. 2019, 9, 11576–11584. [Google Scholar] [CrossRef]
  314. Iravani, S.; Varma, R.S. Green synthesis, biomedical and biotechnological applications of carbon and graphene quantum dots. A review. Environ. Chem. Lett. 2020, 18, 703–727. [Google Scholar] [CrossRef]
  315. Dar, M.S.; Sahu, N.K. Graphene quantum dot-crafted nanocomposites: Shaping the future landscape of biomedical advances. Discov. Nano 2024, 19, 79. [Google Scholar] [CrossRef]
  316. Riley, R.S.; June, C.H.; Langer, R.; Mitchell, M.J. Delivery technologies for cancer immunotherapy. Nat. Rev. Drug Discov. 2019, 18, 175–196. [Google Scholar] [CrossRef]
  317. Zhang, X.; Li, H.; Yi, C.; Chen, G.; Li, Y.; Zhou, Y.; Chen, G.; Li, Y.; He, Y.; Yu, D. Host Immune Response Triggered by Graphene Quantum-Dot-Mediated Photodynamic Therapy for Oral Squamous Cell Carcinoma. Int. J. Nanomed. 2020, 15, 9627–9638. [Google Scholar] [CrossRef]
  318. Abdullah-Al-Nahain; Lee, J.-E.; In, I.; Lee, H.; Lee, K.D.; Jeong, J.H.; Park, S.Y. Target delivery and cell imaging using hyaluronic acid-functionalized graphene quantum dots. Mol. Pharm. 2013, 10, 3736–3744. [Google Scholar] [CrossRef]
  319. McErlean, E.M.; McCrudden, C.M.; McCarthy, H.O. Multifunctional delivery systems for cancer gene therapy. In Gene Therapy—Principles and Challenges; Hashad, D., Ed.; InTech: Houston, TX, USA, 2015; ISBN 978-953-51-2221-0. [Google Scholar]
  320. Liu, Y.; Zhao, C.; Sabirsh, A.; Ye, L.; Wu, X.; Lu, H.; Liu, J. A Novel Graphene Quantum Dot-Based mRNA Delivery Platform. Chem. Open 2021, 10, 666–671. [Google Scholar] [CrossRef] [PubMed]
  321. Fischer, D.; Li, Y.; Ahlemeyer, B.; Krieglstein, J.; Kissel, T. In vitro cytotoxicity testing of polycations: Influence of polymer structure on cell viability and hemolysis. Biomaterials 2003, 24, 1121–1131. [Google Scholar] [CrossRef] [PubMed]
  322. Nimesh, S. Polymeric nanoparticles for gene delivery. In Gene Therapy; Elsevier: Amsterdam, The Netherlands, 2013; pp. 137–145. ISBN 9781907568404. [Google Scholar]
  323. Su, Z.; Shen, H.; Wang, H.; Wang, J.; Li, J.; Nienhaus, G.U.; Shang, L.; Wei, G. Motif-Designed Peptide Nanofibers Decorated with Graphene Quantum Dots for Simultaneous Targeting and Imaging of Tumor Cells. Adv. Funct. Mater 2015, 25, 5472–5478. [Google Scholar] [CrossRef]
  324. Ghafary, S.M.; Nikkhah, M.; Hatamie, S.; Hosseinkhani, S. Simultaneous Gene Delivery and Tracking through Preparation of Photo-Luminescent Nanoparticles Based on Graphene Quantum Dots and Chimeric Peptides. Sci. Rep. 2017, 7, 9552. [Google Scholar] [CrossRef] [PubMed]
  325. Şenel, B.; Demir, N.; Büyükköroğlu, G.; Yıldız, M. Graphene quantum dots: Synthesis, characterization, cell viability, genotoxicity for biomedical applications. Saudi Pharm. J. 2019, 27, 846–858. [Google Scholar] [CrossRef]
  326. Chen, C.-K.; Huang, P.-K.; Law, W.-C.; Chu, C.-H.; Chen, N.-T.; Lo, L.-W. Biodegradable Polymers for Gene-Delivery Applications. Int. J. Nanomed. 2020, 15, 2131–2150. [Google Scholar] [CrossRef]
  327. Kim, H.J.; Lee, H.N.; Jeong, M.S.; Jang, S.B. Oncogenic KRAS: Signaling and drug resistance. Cancers 2021, 13, 5599. [Google Scholar] [CrossRef]
  328. Dong, H.; Dai, W.; Ju, H.; Lu, H.; Wang, S.; Xu, L.; Zhou, S.-F.; Zhang, Y.; Zhang, X. Multifunctional Poly(L-lactide)-Polyethylene Glycol-Grafted Graphene Quantum Dots for Intracellular MicroRNA Imaging and Combined Specific-Gene-Targeting Agents Delivery for Improved Therapeutics. ACS Appl. Mater. Interfaces 2015, 7, 11015–11023. [Google Scholar] [CrossRef]
  329. Zhao, H.; Ding, R.; Zhao, X.; Li, Y.; Qu, L.; Pei, H.; Yildirimer, L.; Wu, Z.; Zhang, W. Graphene-based nanomaterials for drug and/or gene delivery, bioimaging, and tissue engineering. Drug Discov. Today 2017, 22, 1302–1317. [Google Scholar] [CrossRef]
  330. Liu, F.; Ding, N.; Huo, D.; Yang, G.; Wei, K.; Guan, G.; Li, Y.; Yang, J.; Wang, T.; Wang, Y.; et al. Surface-Engineered Monocyte Inhibits Atherosclerotic Plaque Destabilization via Graphene Quantum Dot-Mediated MicroRNA Delivery. Adv. Healthc. Mater. 2019, 8, e1900386. [Google Scholar] [CrossRef]
Figure 1. Properties and applications of GQDs in cancer therapy. High surface area enables efficient drug loading and interaction with biological molecules; photoluminescence allows tracking and can be used to trigger drug release and to induce cytotoxic effects in cancer cells; biocompatibility ensures safety for use in biological environments; and drug and gene delivery allow targeted transport and controlled release of chemotherapeutics and expression of therapeutic genes in cancer cells. Figure was created using Biorender.com (accessed on 1 June 2024).
Figure 1. Properties and applications of GQDs in cancer therapy. High surface area enables efficient drug loading and interaction with biological molecules; photoluminescence allows tracking and can be used to trigger drug release and to induce cytotoxic effects in cancer cells; biocompatibility ensures safety for use in biological environments; and drug and gene delivery allow targeted transport and controlled release of chemotherapeutics and expression of therapeutic genes in cancer cells. Figure was created using Biorender.com (accessed on 1 June 2024).
Ijms 25 10539 g001
Figure 2. Illustration of standard methods of GQDs production and their innovations. (a) Standard methods, top-down and bottom-up approaches; (b) introducing defects at the graphene sp2 structure facilitates cutting the structure to obtain GQDs; (c) graphene-based nanomaterials, such as carbon nanotubes and fullerenes can be utilized to obtain GQDs by means of the top-down methods; (d) solution chemistry processes induce chemical changes in molecules to facilitate their fusion to produce GQDs; (e) use of templates to produce GQDs improves the control in their sizes and morphologies. Insets show characteristics of each method. For comparison, requirements to implement GQDs in market are: low size dispersion, low price, and high quantity of produced material. Figure was created using Biorender.com (accessed on 1 June 2024).
Figure 2. Illustration of standard methods of GQDs production and their innovations. (a) Standard methods, top-down and bottom-up approaches; (b) introducing defects at the graphene sp2 structure facilitates cutting the structure to obtain GQDs; (c) graphene-based nanomaterials, such as carbon nanotubes and fullerenes can be utilized to obtain GQDs by means of the top-down methods; (d) solution chemistry processes induce chemical changes in molecules to facilitate their fusion to produce GQDs; (e) use of templates to produce GQDs improves the control in their sizes and morphologies. Insets show characteristics of each method. For comparison, requirements to implement GQDs in market are: low size dispersion, low price, and high quantity of produced material. Figure was created using Biorender.com (accessed on 1 June 2024).
Ijms 25 10539 g002
Figure 3. Drug and gene delivery for cancer therapy. The figure illustrates the entry of a multifunctional GQD-based drug and gene delivery system through receptor-mediated and non-receptor-mediated endocytosis. This system inherently possesses photothermal properties and is modified with peptides and polymers that enhance its bioavailability, biocompatibility, half-life, and physicochemical properties. The figure also shows a targeting ligand molecule that guides the complex specifically toward cancer cells. Additionally, this system is loaded with a photosensitizer agent that induces oxidative stress as part of photodynamic therapy, a drug that exerts its biological activity by causing DNA damage, and nucleic acids that express therapeutic proteins or silence damage-causing genes. All these components share the common goal of inducing tumor cell death. Figure was created using Biorender.com (accessed on 1 June 2024).
Figure 3. Drug and gene delivery for cancer therapy. The figure illustrates the entry of a multifunctional GQD-based drug and gene delivery system through receptor-mediated and non-receptor-mediated endocytosis. This system inherently possesses photothermal properties and is modified with peptides and polymers that enhance its bioavailability, biocompatibility, half-life, and physicochemical properties. The figure also shows a targeting ligand molecule that guides the complex specifically toward cancer cells. Additionally, this system is loaded with a photosensitizer agent that induces oxidative stress as part of photodynamic therapy, a drug that exerts its biological activity by causing DNA damage, and nucleic acids that express therapeutic proteins or silence damage-causing genes. All these components share the common goal of inducing tumor cell death. Figure was created using Biorender.com (accessed on 1 June 2024).
Ijms 25 10539 g003
Table 1. Carbon sources and methods used for the synthesis of GQDs, their functionalization, average size and thickness, and applications.
Table 1. Carbon sources and methods used for the synthesis of GQDs, their functionalization, average size and thickness, and applications.
Carbon SourceMethodsFunctionalizationAverage Size/ThicknessApplicationReference
Rice grainsPyrolysis-OH, C-H,
C-O-C, C=O
2.0–6.5 nm/–Biomarkers for cell imaging [105]
Carbon black, nitric acidLiquid-phase exfoliationC-N, N-H,
-NH-CO-
1.0–4.0 nm
/<0.8 nm
Cellular imaging, drug delivery [106]
Mango leavesMicrowave-OH, C=C, C-C2.0–8.0 nm/–Bioimaging, cytoplasm labeling [107]
Graphene oxideHydrothermalC-N1.84 nm/–Biological imaging[108]
Cotton celluloseHydrothermalO-C=O, C-OH,
C=O, C=C/C-C
3.0–3.5 nm
/0.5–0.8 nm
In vitro imaging[109]
Rice huskHydrothermal3.0–6.0 nm
/0.8–1.6 nm
Cell imaging[110]
Graphene oxide, Ammonium hydroxideHydrothermalC-O, C-C, C=C,
C-N, N-H
4.0–6.0 nm/–Cells imaging, Osteogenic differentiation [111]
Carbon blackLiquid-phase exfoliation–/1.0 nmTumor cell labeling and photothermal therapy[112]
Citric acid, ureaHydrothermalO-H, N-H,
C=O, C-N
2.0–5.0 nm/–Methotrexate-loaded drug delivery system[113]
L-glutathione (Glycine, cysteine, and glutamic acid)HydrothermalC=C, C-O-C, C=O Cancer theranostics[114]
Cow milkMicrowaveC=C, C-O, C-N, C=N, O-C=O5.0 nm/–Cancer theranostic system[115]
Table 2. Functionalization and pharmacological interaction characteristics of graphene quantum dots in in vivo and in vitro assays.
Table 2. Functionalization and pharmacological interaction characteristics of graphene quantum dots in in vivo and in vitro assays.
Functionalization/DopingDrugSize of ComplexInteraction
(GQDs-Drug)
CharacteristicsAssaysReference
In VitroIn Vivo
Cell LineEvaluated ConcentrationsAnimal ModelEvaluated Concentrations
-DOXGQD-DOX: 99–107 nmπ-πRapid nuclear uptake
Biocompatibility
MCF-7
WI-38
3.125 up to 200 µg/mL of GQDs-DOXAlbino Wistar rats5 up to 20 mg/kg of GQDs [217]
NDOX-π-πIncreased biocompatibilityMCF-7
HeLa
HEK-293
0.00625 up to 1.6 µg/mL of DOX --[241]
N, S and Immobilized on MSNDOXMSN-GQD: 49.5 nmNon-covalent bondGreater internalization
Biocompatibility
HeLa
MSU-1.1
20 up to 400 µg/mL of MSN-GQDs--[307]
-COOH and Cu +2DOX-π-πIncreased cytotoxicityMCF-7/ADR50 up to 250 µg/mL of GQDs--[311]
NMTX-π-πIncreased biocompatibilityMCF-70 up to 4 µM of MTX-GQD--[113]
BTNDOX-π-πIncreased targeting Increased biocompatibilityA5493.75 up to 15 µg/mL of GQDs-BTN-DOX--[294]
scFvB10CDDP-Electrostatic interactionsIncreased targetingMDA-MB-2315 up to 75 µg/mL of CDDP--[219]
FA and mPEG2000TMXTMX-FPG: 294.7 nm.π-πIncreased targetingMCF-7
HDF
1 up to 200 µg/mL of TMX-FPG --[224]
HAHSA-encapsulated gemcitabine HSA-NPs: 150 nm(CO–NH)Increased targeting
Increased drug half-life
Panc-15 up to 400 µg/mL of HSA-NPs--[288]
RGDDOX-π-π
Hydrophobic interactions.
Increased targeting
Increased toxicity
U2510 up to 200 µg/mL of RGD-GQDs-DOX--[283]
PEG and FAMTNFA-PEG-cGQDs: 4–10 nmπ-π
Hydrogen bonds
Increased targeting
Increased biocompatibility
HeLa0.05 up to 1 μg/mL of MTNAthymic nude mice2.5 mg/kg of MTN.[218]
FAIR780-π-πIncreased targeting
Combination with photothermal therapy
HeLa0 up to 30 µg/mL of FA-GQDs- IR780Balb/c nude mice2 mg/kg of FA-GQDs- IR780[302]
GE11CDDP
DOX
GE11-GQD:14.23 nmπ-πIncreased targeting
Combination of two pharmacological therapies
CNE-20 up to 100 µg/mL of GE11-GQDBalb/c nude mice2 mg/kg of DOX and CDDP[216]
-B-Lap-π-πGreater antitumor effect compared to the free drug
High biocompatibility
Rapid release of the drug in aqueous medium
LO2
HeLa
PC-12
MCF-7
5 up to 10 uM of B-Lap--[312]
Cy5.5DOXDOX-GQDs-P-CY: 15.6 nmπ-πHigh therapeutic activity in vitro and in vivo
Theranostic and monitoring agent
4T10 up to 4 µg/mL of DOX-GQDs-P-CYOrthotopic mice model-[296]
BM and PEGPorphyrin--Detection of cancer-associated miRNAs
Efficient photothermal conversion
High production of singlet oxygen
A54910 up to 200 µg/mL of GQD-PEG-P--[297]
PEI and EGFRDOX GQDs-DOX:
40–80 nm
π-πTumor suppression capacity
Longer-sustained inhibition capacity
Less toxicity than DOX
HCT-116<5 μM of GQDs-DOXNude mice200 μL of GQDs-DOX[290]
hMSN, PEG and VEGFDOXGQDs-hMSN-PEG-VEGF-DOX and GQDs-hMSN-PEG-PEG-DOX: 100–150 nm-Biocompatibility
High DOX loading efficiency
Structural stability
Enhanced DOX delivery
High specificity
MCF-7
L929
0–10 μg/mL of GQDs-hMSN-PEG-VEGF-DOX and GQDs-hMSN-PEG-PEG-DOXBALB/c nude mice5 mg/kg of GQD-hMSN[313]
HADOXGQD-HA: 20 nm Increased targeting
Increased toxicity
A5490.5–5 μg/mL of GQD-HA-DOXBALB/c mice10 mg/kg of GQD-HA[318]
Note: DOX: Doxorubicin; MCF-7: Human breast cancer; WI-38: Human fetal lung fibroblasts; N: Nitrogen; HeLa: Human cervical cancer; HEK-293: Human embryonic kidney; S: Sulfur; MSN: Mesoporous silica nanoparticles; MSU-1.1: Human fibroblasts; Cu+2: Copper II ion; MCF-7/ADR: Human breast cancer with doxorubicin resistance; MTX: Methotrexate; BTN: Biotin; A549: Human alveolar carcinoma cells; scFvB10: Single chain variable fragment antibody; CDDP: Cisplatin; MDA-MB-231: Human breast adenocarcinoma; FA: Folic acid; mPEG2000: Methoxypolyethylene glycol; TMX: Tamoxifen; TMX-FPG: mPEG and FA-functionalized GQDs loaded with TMX; HDF: Human dermal fibroblasts; HA: Hyaluronic acid; HSA-NPs: Hyaluronic acid conjugated graphene quantum dots and human serum albumin nanoparticles; Panc-1: Human pancreatic epithelioid carcinoma; RGD: Arginine, glycine, aspartic acid; U251: Human glioma cells; Ag: Silver; PEG: Polyethylene glycol; DU145: Human prostate cancer; NH2-PEG-NH2: Aminated polyethylene glycol; MTN: Mitoxantrone; FA-PEG-cGQDs: IR780: Theranostic agent for fluorescence imaging and photothermal therapy; GE11: Peptide targeting epidermal growth factor; CNE-2: Human nasopharyngeal carcinoma; B-Lap: B-Lapachone; LO2: Human liver cells; PC-12: Mouse neural cells; Cy5.5: Cyanine-5.5; 4T1: Mouse T cells; BM: Aptamer for miRNA 155 detection; PEI: Polyethylenimine; EGFR: Epidermal growth factor receptor; HCT-116: Human colorectal cancer cells; hMSN: hollow mesoporous silica nanoparticles; L-929: Mouse fibroblasts.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Osorio, H.M.; Castillo-Solís, F.; Barragán, S.Y.; Rodríguez-Pólit, C.; Gonzalez-Pastor, R. Graphene Quantum Dots from Natural Carbon Sources for Drug and Gene Delivery in Cancer Treatment. Int. J. Mol. Sci. 2024, 25, 10539. https://doi.org/10.3390/ijms251910539

AMA Style

Osorio HM, Castillo-Solís F, Barragán SY, Rodríguez-Pólit C, Gonzalez-Pastor R. Graphene Quantum Dots from Natural Carbon Sources for Drug and Gene Delivery in Cancer Treatment. International Journal of Molecular Sciences. 2024; 25(19):10539. https://doi.org/10.3390/ijms251910539

Chicago/Turabian Style

Osorio, Henrry M., Fabián Castillo-Solís, Selena Y. Barragán, Cristina Rodríguez-Pólit, and Rebeca Gonzalez-Pastor. 2024. "Graphene Quantum Dots from Natural Carbon Sources for Drug and Gene Delivery in Cancer Treatment" International Journal of Molecular Sciences 25, no. 19: 10539. https://doi.org/10.3390/ijms251910539

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

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop