Next Article in Journal
Evaluation of the Inhibitory Potential of Synthetic Peptides Homologous to CDR3 Regions of a Monoclonal Antibody against Bothropic Venom Serine Proteases
Previous Article in Journal
Syncytiotrophoblast Markers Are Downregulated in Placentas from Idiopathic Stillbirths
Previous Article in Special Issue
Exploring the Effect of V2O5 and Nb2O5 Content on the Structural, Thermal, and Electrical Characteristics of Sodium Phosphate Glasses and Glass–Ceramics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Perspective

Graphene and Natural Products: A Review of Antioxidant Properties in Graphene Oxide Reduction

by
Filipe Kayodè Felisberto dos Santos
1,*,
Antônio Augusto Martins Pereira Júnior
2,
Arquimedes Lopes Nunes Filho
3,
Clícia Joanna Neves Fonseca
3,
Daysianne Kessy Mendes Isidorio
4,
Filipe de Almeida Araújo
5,
Pablo Henrique Ataide Oliveira
6 and
Valdir Florêncio da Veiga Júnior
1,*
1
Chemical Engineering Section, Military Institute of Engineering, Rio de Janeiro 22.290-270, Brazil
2
Higher School of Engineering, Technology and Innovation, University of the Federal District Professor Maia Nunes, Brasília 70.635-815, Brazil
3
Postgraduate Program in Materials Science and Engineering, Military Institute of Engineering, Rio de Janeiro 22.290-270, Brazil
4
Department of Metallurgical and Materials Engineering, Federal University of Rio de Janeiro, Rio de Janeiro 21.941-901, Brazil
5
Postgraduate Program in Materials Science and Engineering, Federal University of São Carlos, São Carlos 13.565-905, Brazil
6
Higher Education Department of Education, Federal Institute of the North of Minas Gerais, Bom Jardim 39.480-000, Brazil
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(10), 5182; https://doi.org/10.3390/ijms25105182
Submission received: 22 March 2024 / Revised: 30 April 2024 / Accepted: 2 May 2024 / Published: 9 May 2024
(This article belongs to the Collection Feature Papers in Materials Science)

Abstract

:
This review article addresses the antioxidant properties of different natural products, including ascorbic acid, gallic acid, oxalic acid, L-glutathione (GSH), bacteriorhodopsin, green tea polyphenols, glucose, hydroxycinnamic acid, ethanoic acid, betanin, and L-glutathione, in the reduction of graphene oxide (rGO). rGO can cause damage to cells, including oxidative stress and inflammation, limiting its application in different sectors that use graphene, such as technologies used in medicine and dentistry. The natural substances reviewed have properties that help reduce this damage, neutralizing free radicals and maintaining cellular integrity. This survey demonstrates that the combination of these antioxidant compounds can be an effective strategy to minimize the harmful effects of rGO and promote cellular health.

Graphical Abstract

1. Introduction

The advancement of studies related to nanotechnology has developed products and by-products with a multitude of technological applications, among which, graphene stands out. Graphene consists of a two-dimensional (2D) network of carbon atoms (C) arranged in densely packed hexagonal structures, whose thickness is of the order of 0.15 nm (diameter of the carbon atom) and density of 0.3 mg/cm. Having said this, it is noted that graphene is a thin, light, flexible, and highly resistant nanomaterial [1,2].
Regarding its properties, excellent electrical, mechanical, and thermal properties can be highlighted, resulting, above all, from its nanostructural and electronic characteristics. For example, it has a high Young’s modulus (1.0 TPa), higher than that of many steels, indicating that high mechanical stress is necessary to cause deformation [3]. Furthermore, it presents an extraordinary mobility of electrons due to the rapid transport of charges between the layers and the large surface area, which results in high electrical conductivity at room temperature, making graphene very suitable for use in supercapacitors [4]. Furthermore, it also has excellent thermal conductivity and can be transparent, making it suitable for designing solar panels and touch screens. The small areas of each carbon hexagon are responsible for the high impermeability of graphene, which can be used as a small network capable of holding gases that easily leak from their containers, such as hydrogen gas. In addition to being extremely resistant, graphene is very light: its density is 0.77 g/m2, that is, around a thousand times lighter than a sheet of paper [4]. These well-known characteristics have encouraged studies and research for into graphene’s application in energy storage and conversion and in environmental applications, as well as in the assessment of toxicity in biological systems [5,6,7].
In electronics, graphene’s high electrical conductivity and transparency have led to advancements in flexible and transparent electronic devices. Moreover, in energy storage, graphene-based materials show potential for enhancing the performance of supercapacitors and batteries due to their large surface area and excellent conductivity [8,9]. In the realm of medical science, the exploration of nanomaterials (such as graphene and its derivatives) has emerged as a groundbreaking avenue in disease management, offering a promising application in healthcare. Taking advantage of graphene’s unique characteristics, researchers are developing advanced drug delivery systems capable of precisely targeting diseased cells while minimizing collateral damage to healthy tissues [10]. Furthermore, graphene-based biosensors exhibit exceptional sensitivity, enabling the early detection and monitoring of various diseases, from cancer to neurological disorders [11,12]. Additionally, graphene’s antimicrobial properties hold potential in combating antibiotic-resistant pathogens, presenting a promising solution to the global healthcare challenge of infectious diseases [13].
The introduction of reduced graphene oxide (rGO) into cells can induce oxidative stress. In the context of oxidative stress-related biological outcomes, two primary mechanisms dominate. The first mechanism centers on the production of reactive oxygen species (ROS) and reactive nitrogen species (RNS), leading to biomolecular oxidation, consequentially resulting in cellular dysfunction and mortality. Simultaneously, the second oxidative stress mechanism revolves around the perturbation of redox signaling pathways. Consequently, antioxidants play a crucial role, either by scavenging ROS and RNS [14,15] or by modulating redox signaling pathways [16].
On the other hand, the method of obtaining rGO through different synthesis can be worrying due to the risk of toxicity, such as reduction with hydrazine [17]. Another point to be highlighted is that GO has been shown to be a biocompatible material; however, when inserted into cell culture medium, there is a loss of cell functionality in 20% and 50% when 20 mg.mL−1 and 50 mg.mg−1, respectively, is added [18]. Furthermore, studies report that graphene-based particles present in the blood can travel to different organs in our body and can reach the nervous system, and when ingested intravenously or orally, they mainly reach the kidneys, lungs, and liver, leading to inflammation [19,20]. Another publication with in vivo studies shows that prolonged exposure to GO in doses of 25 micrograms L−1 caused opacity in the animals’ eyeballs, in contrast to RGO, which exhibited biosafety [21].
Concern about environmental issues, especially in terms of minimizing or eliminating harmful chemical substances or those that result in toxic waste, has encouraged many researchers to obtain reduced graphene oxide, and this method has been called green reduction of graphene oxide. Among the various approaches, one can include the use of natural reducing agents, such as plant extracts, and reduction methods using water as a solvent. However, the big challenge of this green route is related to efficiency when compared to the synthesis route with hydrazine, for example, and obtaining an rGO with good properties. Among all the green routes mentioned by [22] the one where ascorbic acid was used to obtain RGO showed the most promise. In addition to being a recent process, it has shown excellent results compared to hydrazine methods; however, more research is needed in order to model and find a better optimization process for this route with changing parameters and conditions in search of large-scale production [22,23,24].
Given the relevance of graphene to industry and its promising properties, synthesis methods for this material are the target of research, as the large surface area and oxygen-containing functional groups of graphene oxide (GO), such as phenol, hydroxyls, epoxy and carboxylic groups, make it ideal for providing high substance loading efficiency, good dispersion, and easy functionalization [25]. Among those already used are the micromechanical exfoliation of highly oriented pyrolytic graphite (HOPG), chemical vapor deposition (CVD), and epitaxy. These techniques produce different qualities of graphene with relative structural perfection and excellent properties. A highly relevant process for obtaining graphene consists in reducing its oxide, as it is the most reliable and commercially viable method among those currently known for this purpose [26]. Reduction can have different approaches, such as thermal, photochemical, biological, and chemical. In the latter, the substance interacts with the oxygenated groups present in GO, releasing water, which will later be used in a heat treatment to form the π bonds of the graphene structure [27].
In this context, different natural products emerge as potential agents for reducing GO, such as antioxidant compounds and their great capacity for removing and/or stabilizing oxygenated groups present in GO. Substances such as ascorbic acid, gallic acid, oxalic acid, and others have been presented as important natural sources for obtaining graphene [28,29,30]. This review article presents a detailed and comprehensive analysis on the potential applications of natural antioxidants in the graphene oxide reduction process. This review stands out for its depth and scientific rigor, offering a solid foundation of knowledge that not only expands the understanding of reduced graphene oxide, but also sheds light on the effective use of natural reducing agents. Furthermore, this study also investigates the environmental and biological implications of these interactions, suggesting promising paths for future research and technological applications that combine sustainability and innovation in materials engineering.

1.1. Chemical Reduction

The formation of GO involves the reaction of graphite with strong oxidants such as sulfuric acid, nitric acid, potassium chlorate, and potassium permanganate. The introduction of functional groups containing oxygen, through complete oxidation, results in an increase in the spacing between the graphite layers, subsequently making it possible to obtain graphene [30]. SiO2 surfaces at 150 °C are spray-coated and the graphite oxide can be chemically reduced using hydrazine. However, this reducing agent is toxic and dangerous, and its use must be carried out with extreme care and in minimum quantities. On the other hand, natural reducing agents can act in a safer and more environmentally friendly way with regard to graphene production, thus enabling the development and compatibility of more sustainable and less expensive synthesis routes (Figure 1) [31,32].

1.2. Graphene Oxide-Reducing Agents from Natural Products

Graphene is a two-dimensional (2D) material composed of a single allotropic carbon sheet with sp2 hybridization. Producing graphene in large quantities is one of the biggest challenges in the materials sector. Of the various methods that have been explored, GO reduction is considered the most reliable and commercially viable method [30]. GO can exhibit exceptional electronic, mechanical, and thermal properties similar to graphene, reducing treatments that transform GO into reduced graphene oxide (RGO) [33]. Various reduction methods have been developed, such as chemical, biological, photo-mediated, and thermal reduction [30]. Some substances commonly used for these reductions are hydrazine, hydroquinone, and sodium borohydride. Unfortunately, these are dangerous chemicals in terms of human health and environmental awareness. Thus, implementing sustainable alternatives to these reducing agents would be highly beneficial, in turn facilitating the large-scale production of GO graphene for multiple applications [31]. The use of ascorbic acid as a GO reducing agent has been explored for various purposes and using the most varied combinations of analysis techniques, as can be seen in Table 1.
Next, in Table 2, oxalic acid is presented, a common reducing agent due to its greater reducing power when compared to other acids such as acetic acid.
Gallic acid (Table 3) and GSH (Table 4) are reducing agents that have been recently exploited mainly as precursors for nanoparticle deposition, sensory applications, and electrodes.
Polyphenols (Table 5) and glucose (Table 6) are natural reducers whose use has grown in recent years. This is due to their characteristics and applications, such as drug delivery systems and treatments for specific conditions of the human body.

1.2.1. Ascorbic Acid

Ascorbic acid (AA), widely known as “vitamin C” (Figure 2), is a ketolactone found mainly in citrus fruits, such as oranges and lemons, but is also available in other natural sources, such as acerola, kale, tomatoes, broccoli, strawberries, and peppers [48]. When applied as an antioxidant, its main function is to remove oxygen, capturing it in the reaction medium, making it unavailable to propagate auto-oxidation [49]. The removal of oxygen functional groups by AA is relatively easy, as a small amount of AA could carry out the reduction in a short time, resulting in a product with a high carbon/oxygen ratio. It can also present another activity, in which the molecule donates electrons to radicals and which will stabilize through resonance, preventing such radicals from reacting with oxygen or other molecules to form more reactive radicals [50].
Its antioxidant properties as a scavenger of free radicals are related to its ability to form a stabilized radical, allowing the formed ascorbate to react with more reactive molecules. As such, the ascorbate both suppresses free radicals and stimulates the oxidant system on the cells while acting on antioxidant enzymes as a ROS eliminator [51].
Furthermore, its lower toxicity and greater biocompatibility compared to other common reductants such as hydrazine or sodium borohydride make ascorbic acid safer and more sustainable for applications in biotechnology and other areas. Its wide availability, stability, and antioxidant properties are also recommended for its effectiveness and practicality, protecting the formed graphene from oxidative manipulation. Therefore, ascorbic acid proves to be an efficient reducing agent not only due to its physicochemical characteristics, but also due to its environmental and safety advantages, making it an ideal choice for the sustainable production of graphene in scientific and industrial applications [52].

1.2.2. Gallic Acid

Gallic acid (3,4,5-trihydroxybenzoic acid) (GA) (Figure 3) is a polyphenol, naturally found in teas, grapes, raspberries, blackberries, blueberries, and other fruits, in addition to being present in honey and wine. Furthermore, it is present in some plants, such as oak (Quercus robur) and Portuguese chestnut, as this species is known in Brazil (Castanea sativa L.) [53]. The substance is characterized as being a primary antioxidant, that is, it promotes the removal or inactivation of free radicals present in the reaction medium through the donation of hydrogen [54]. When used to reduce GO, gallic acid proves to be a compound with promising results in the process. The phenolic hydroxyl in its structure reacts with the epoxy group present in GO, to form a hydroxyl. The carbon of this newly formed hydroxyl reacts with another hydroxyl, creating an intermediate product and water. This intermediate product is subsequently subjected to elimination reactions to form the sp2 hybridized structure [55].
Faced with the need to develop a low-cost, high-efficiency, and environmentally friendly method that is capable of balancing graphene quality and production efficiency, several methodologies were created that apply GA to reduce GO, such as the GOAG nanocomposite, an anticancer nano dose delivery system using GO as a nanocarrier for an active anticancer agent, AG. When GOAG was applied to liver cancer cells, it showed an inhibitory effect on the growth of cancer cells without affecting normal cell growth [56]. Another approach to chemical reduction of GO using GA is the production of hybrid papers, multicarboxylic carbon-walled nanotubes (AG-rGO/MWCNTs), a mechanically robust material with gravimetric capacitance and ultra-high volumetric capacitance, which has cyclic stability capacity for capacitive electrochemical energy storage. Therefore, high-performance electrodes can be constructed using AG-rGO combined with carbon nanotubes. The results of these research studies are highly encouraging to continue exploring the potential of GO reduction by AG as well as other environmentally friendly and applicable reducers [57].

1.2.3. Oxalic Acid

Oxalic acid (OA) or ethanedioic acid (Figure 4) is an organic acid with a simple structure. Members of the plant kingdom, including land plants and woody trees such as conifers, accumulate and retain oxalic acid in their leaves, needles, and twigs. OA stands out for being present in broccoli, cabbage, and cauliflower leaves when subjected to cooking, as well as in the outer layers of cereal grains, such as wheat bran [58]. OA stands out for preserving quality and postponing the physiological effects of fruit aging. Although its mechanism of action has not yet been elucidated, the substance is related to preventing the accumulation of reactive oxygen species in the reaction medium [59].
Furthermore, its action in reducing GO occurs by removing oxygenated groups from GO, releasing H2O. A hydroxyl group is protonated in a basic sodium oxalate medium, producing water and forming the carbocation. The oxalate ions then interact with the neighboring hydrogens to form the π bonds of the graphene structure [59]. Based on previous studies, rGO reduced using oxalic acid (rGO/AO) has a rough surface, folded edges, ripples, a multilayer structure, tangles of graphene layers, and the presence of ultra-thin and transparent graphene sheets as its characteristic morphological profile. Furthermore, in view of this, it is noteworthy that the folds in the edge regions confer high tenacity to the material, therefore leading to excellent mechanical properties [60].
rGO/OA has a structure based on flat layers with agglomerated graphene sheets, whose dimensions vary from ≅8 to 10 μm. However, there are reports of the synthesis of large-sized GO sheets reduced with OA in the range of 0.5 to 10 μm. rGO/AO, in general, has a porous nature with pore sizes of the order of ≅2–5 μm. The process of reducing GO with OA leads to a rapid removal of oxygen atoms from the graphene basal plane, thus allowing the modulation of the desired morphology (dendrites, nanosheets, and nanorods) [60]. In addition, it is worth highlighting that OA (an oxidized form of ascorbic acid) confers greater stability to rGO. This fact arises from the affinity of this compound to bind hydrogen atoms from residual oxygenated functional groups, thus interrupting π–π between the rGO sheets and the formation of agglomerations [61]. It is also noteworthy that graphene obtained via chemical exfoliation with OA can present a nanostructural configuration based on monolayers or even multiple overlapping carbon layers [60,61].

1.2.4. L-Glutathione (GSH)

L-glutathione (Figure 5), commonly known as GSH, is the main thiol-containing tripeptide of low molecular mass (307.32 g mol−1), composed of glutamate, cysteine, and glycine [62]. It is found in most living cells of bacteria, mammals, plants, and fungi. GSH is often used for its antioxidant and anti-inflammatory properties and is available in supplement form but can also be obtained from dietary sources. Foods rich in GSH include Brazil nuts, broccoli, Brussels sprouts, kohlrabi, cauliflower, cabbage, onions, and some mushrooms [63]. Among its many functions, it can also protect important cellular compounds from oxidative stress caused by free radicals, air pollution, heavy metals, and the harmful effects of radiation, by acting as a scavenger, neutralizing reactive oxygen species (ROS) [62]. Thus, by incorporating GSH into the preparation of GO, it simultaneously reduces and stabilizes the monolayers of GO nanosheets [64]. This occurs because the electrostatic repulsion between the terminal carboxylic acid of GSH and GO provides sufficient negative charge, promoting stable dispersion in aqueous solutions and polar aprotic solvents, thus avoiding agglomeration and precipitation [63].
L-glutathione (GSH) is a natural antioxidant that allows the reduction of multiple types of oxygenated compounds [63]. Thus, the oxygenated form of L-glutathione has external carboxylic groups with a negative charge, which generates electrostatic repulsion, promoting a stable dispersion of graphene, inhibiting the regrouping of its nanosheets [64]. In recent works [65], the chemical reduction of GO was carried out using GSH. By observing the FTIR spectrum, this reduction can be checked by analyzing the absorption peaks (Figure 6). The peaks corresponding to the oxygenated functional groups were reduced, with the functional groups disappearing, leaving only one at 1534 cm−1, which relates to the graphitic structure of graphene nanosheets [66].
Looking at Figure 7, the morphology of GO can be seen, characterized by thin sheets of graphene, whose appearance is characterized by wrinkles and folds. Furthermore, it appears that the material has a multilayer configuration, that is, densely packed areas of stacks and agglomerations.

1.2.5. Bacteriorhodopsin (bR)

Bacteriorhodopsin (bR) is a protein membrane found in salt marshes of the Archaea Halobacterium salinarum, and consists of photoactive fractions called retinal chromophores (C20H28O) in the middle of a channel formed by seven transmembrane alpha-helices (A–G) [67]. As it is a proton pump powered by light, when light of one wavelength is irradiated, it can convert the bound electrons into free radicals [68]. Thus, its use as a GO reduction mechanism involves the distribution of bR protons between the GO sheets. When yellow light (≅80 mW cm−2, 530 nm) is irradiated, the oxygen group interacts, forming water molecules and restoring the sp2 hybridized carbon structure, therefore resulting in rGO [69].
Despite the sustainable potential of bR as an effective graphene reducer, there are few studies in the literature that report its morphology. Only one study was found on the morphology of bR/rGO. Researchers evaluated bR molecules as light-driven proton pumps for the reduction of single layers of GO. In this sense, through AFM, the topography and distribution of the height profile of the graphene and bR sheets can be observed. Furthermore, there was an absence of direct contact between bR molecules and GO sheets [70].

1.2.6. Green Tea Polyphenols

Green tea polyphenols (GTPs) are a group of compounds that increase human metabolism, resulting in many health benefits. GTPs include several phytochemicals such as epicatechin (EC), epicatechin gallate (ECG), epigallocatechin (EGC), and epigallocatechin gallate (EGCG) (Figure 8). Among these, EGCG has the highest concentration of catechins, indicating its high antioxidant functionality [71]. In its chemical structure, tea polyphenols (TPs) are able to donate an electron or hydrogen atom, making it an effective reducing agent.
Thus, the reaction process between the phenolic hydroxyl groups of Tps and the epoxy group in GO forms a hydroxyl group by opening the nucleophilic ring. These hydrogen ions act as catalysts to promote the dehydration of OG, resulting in the formation of water as a byproduct and the sp2 hybridized carbon structure of rGO (Figure 9).
GTPs, up to 30% of dry weight, have the highest concentration of all catechins in GT (≅9–13% of dry weight) and are known to be the most powerful active antioxidant components of GT as a reductant during the synthesis of rGO [72]. Thus, it is observed that the mechanism for eliminating oxygenated functional groups from GO via these polyphenols is reflected in the internal folds and pores of the rGO structure. Furthermore, it appears that this process favors the production of nanosheets with a large structure, providing a specific surface, which is important for fixing metals [73].
The removal of electrons from the catechin rings makes the hydroxyl bonds more acidic, so the hydroxyl bonds are organized to split two protons and function as a nucleophile. The GO leaves mainly contain two types of reactive oxygen species: epoxide and hydroxyl. Both epoxide and hydroxyl bonds can be opened by means of nucleophilic attack by the oxygen anions of catechins [74]. The reduction must also be followed by the dehydration process, carried out by absorbing an electron from the environment. Next, the catechins are oxidized into benzoquinone-type products. Benzoquinone-type products can be absorbed in leaves reduced by π–π interactions between the aromatic rings of the products and leaves [75].
The GO suspension is reduced in a few minutes and has excellent dispersibility and chemical stability for a period of six months. This stability can be attributed to the fixation of oxidized GTPs on the surface of the reduced leaves [76]. Temperature directly affects the thermal stability of rGO using green tea polyphenols, as there is a weight loss of approximately 22% in the temperature range between 150 and 230 °C. The significant weight loss at this temperature was associated with the decomposition of the oxygen functional group into GO, and there was additional weight loss between 450 and 565 °C, mainly due to the burning of the graphitic regions [76].

1.2.7. Glucose

According to the literature, carbohydrates, also called glycans, are one of the most prevalent biopolymers in nature, and one of the three essential molecules of life alongside polynucleotides and polypeptides. The carbohydrate molecule’s presence is ubiquitous, found in all living beings, from animals and plants to microbes. Furthermore, its function is to play significant roles in physiological processes. In essence, the carbohydrate molecule is composed of carbon, hydrogen, and oxygen atoms, with the general chemical formula (CH2O)n. In this sense, monosaccharides such as glucose (GI) are important sugars for the energy metabolism of most living organisms. Furthermore, it is commercially available in the form of a white, crystallized, sweet, and odorless powder, with the formula C6H12O6 and a molecular weight of 180.16 g/mol [77].
With reducing capacity and non-toxic properties, GI works as a reducing agent, but with less efficiency when used to reduce GO. However, by adding elements such as ammonia, which raises the pH of the reaction or catalysts such as iron foil (Fe), it is possible to increase GI’s reducing functionality. In ammonia solution, GI is oxidized by GO and forms aldonic acid, which can be further converted into lactones. As a result, oxidized glucose products containing hydroxyl and carboxyl groups can form hydrogen bonds with residual oxygen functionalities on the GO surfaces to reduce to a stable glucose-rGO colloidal dispersion [78]. After the reduction of GO by glucose, the color of the suspension changes from light brown to black. This color change can be attributed to the partial restoration of the π–π structure between the leaves due to deoxygenation. This causes the composition of wrinkles at the edge of the rGO sheet, and further reduction can accommodate the properties of the wrinkles as well as their shrinkage [79].
The number of wrinkles present in GO nanoplates can contribute to the dissipation of frictional energy from the interface and improve the vibration damping performance. However, atomic force microscopy (AFM) shows that wrinkles with broad coefficients of friction (COFs) play an important role in delaying sliding events between rGO sheets [80]. Another morphological feature of glucose-reduced GO is the formation of a thick flake-like structure that can result in the agglomeration of much thinner flakes. This is due to end-to-end interactions between oxygen functional groups through hydrogen bonds. It is possible to visualize, in some regions, a small number of carbon spheres that contain a hydrophobic core and a hydrophilic shell [80].
The sugar-assisted thermal reduction of GO is applied to synthesize three-dimensional porous graphene (3D-PG), which can be used as an electrode for supercapacitors. It exhibits capacitance of about 115 F/g at a scan rate of 10 mV/s, and the addition of glucose in the GO reduction process strengthens the interconnectivities between the graphene sheets, making them stick together tightly and further accelerating cargo transport [81]. A green approach is applied to the fabrication of sterile nanometer-sized rGO (nrGO) for photothermal therapy and the delivery of drugs using pure glucose as a reducing agent. nGO reduced in pure glucose restores the conjugated structure and enhances the photothermal effect, making nGO more sensitive to the reduction carried out by glucose aldehyde groups. The oxidation of nitronium in OG promotes the reduction of nrGO to glucose and the degree of biocompatibility. Furthermore, the resulting nrGO has a 317% (w/w) loading of doxorubicin (DOX), a broad-spectrum chemotherapeutic agent known to adsorb on graphene-based materials via hydrophobic interactions and π–π stacking. And the release of DOX from the nrGO/DOX complex can be effectively enhanced by acidic conditions, GSH concentration, and heating, making nrGO a promising synergistic nano-platform for photothermal therapy and the controllable delivery of drugs, which can be applied to the field of nanomedicine [81].
Glucose added to the solution can increase the crosslinking of graphene sheets and the deoxygenation of GO sheets in the presence of metals such as Fe. This helps in recovering the graphitic structure of GO sheets during reduction [81]. The new findings provide relevant information on the green reduction of GO by sugarcane waste and its use to remediate metal ions [81,82]. These studies show the effectiveness of GO reduction by reducing sugars, such as glucose, as well as the applicability of rGO in environmental bioremediation. Sugarcane bagasse is mainly composed of hemicellulose (25–35%), lignin (15–35%), and cellulose (40–45%), and thermochemical treatment breaks the lignocellulosic bonds and partially solubilizes the polysaccharides to release sugars, such as glucose and sucrose [83]. Some applications require rGO with specific functional groups and this can be achieved by a simple one-step hydrothermal method to produce partially reduced GO rich in carbonyl functional groups using glucose as the sole precursor. The carbonyl functional groups present in the sample allow greater functionalization and provide good dispersibility in water, making them useful in biomedical and water treatment applications [83].

1.2.8. Hydroxycinnamic Acid

The compound 3,4-hydroxycinnamic acid (C9H8O4), also known as caffeic acid (CA) (Figure 10) or 3,4-dihydroxycinnamic acid, consists of an aromatic organic compound with a mixed and unsaturated carbon chain consisting of cinnamic acid and hydroxylated benzene [84]. This biomolecule belongs to the group of polyphenols, and more precisely, to the chemical class of phenolic acids. Caffeic acid, as well as its derivatives, are biologically active substances of great importance regarding the tolerance and control of biotic and abiotic stresses in plants, such as, for example, pathogen attacks, low- and high-temperature stress, ultraviolet (UV) light stress, drought stress, heavy metal stress, and salinity stress. It is noteworthy that this phytonutrient is one of the products generated during the secondary metabolism of plant species (phenylpropanoid pathway) [85].
Furthermore, it is worth mentioning that this substance can be found in abundance in plants, fruits, and vegetables, of which the following stand out: olives, coffee, sage, oregano, mint, rosemary, thyme, sunflower seeds, potatoes, carrots, and burdock (sticky grass) [86]. Caffeic acid (CA) can also be isolated from wines, teas, propolis, and vegetable residues. CA and its derivatives (caffeoylquinic acids and dicaffeoylquinic acids), as well as chlorogeneic acid (an ester formed between caffeic acid and quinic acid), are responsible for the enzymatic browning of plant species, thus being natural antioxidant agents [87].
Regarding its use in the production of rGO, caffeic acid is usually used in hydrothermal reduction processes. When used as a reducing agent in such a process, it allows rGO with a morphology consisting of stacked agglomerates to be obtained [88]. Such morphological behavior comes from the hydrothermal process, as the removal of the oxygen group during reduction tends to provide a hydrophobic effect to the rGO layers, agglomerating to reduce free energy, due to hydrogen bonds between the layers. The reduction effect tends to affect the arrangement of the rGO layers, where the previously stacked layers become folded and curled after hydrothermal reduction, an effect corresponding to an attempt by the structure to become thermodynamically stable. The reduction effect tends to reduce the distance between rGO planes, a behavior caused by the reduction in the presence of oxygen in the structure. During reduction, variations in acid proportions tend to promote the formation of nanoparticles, with a reduction in particle size in accordance with the reduction in the proportion of caffeic acid. This behavior occurs due to the presence of rGO, which prevents particle growth [89].
Among the class of green reducing agents, caffeic acid has one of the greatest reducing effects, when used alone; at a C/O ratio of 7.15 during reduction, a low amount of rGO layers are produced, maintaining the standard structure of wrinkling [89].

1.2.9. Ethanoic Acid

Ethanoic acid (CH₃COOH) or acetic acid (AA) is an organic compound with a saturated and straight chain, and it is an important chemical product widely used in the food industry [90]. Due to its characteristics and properties, this substance has traditionally been used in food preservation, as well as being a solvent or intermediate ingredient in a variety of chemical products of commercial interest, among which the following stand out: vinyl acetate monomers (VAM), cellulose acetate (anhydrous acetate), acetate ester, and terephthalic acid (TPA), which is the raw material of polyethylene terephthalate (PET) [91].
In general, AA is produced through an aerobic fermentation process, whose synthesis route is based on the oxidation of ethanol present in fruits (grapes, apples, and coconuts) and grains (rice and wheat). Furthermore, the process is hosted by biological agents, such as fungi and bacteria [92]. Among these are acetic acid bacteria (BAA) and the fungus Micoderma aceti, which are characterized by their versatility with regard to the conversion of a series of carbon sources into biomolecules of industrial interest; however, the route of biological synthesis of ethanoic acid represents only 10% of global production [93]. AA can also be obtained from bio-oil that is produced from the pyrolysis of lignocellulosic biomass. In addition, methanol carbonylation, acetaldehyde oxidation, butane/naphtha oxidation, and methyl acetate carbonylation are the main synthetic pathways for producing acetic acid [93].
The use of AA as a reducing agent in the production of rGO promotes the formation of ordered layers with a predominantly clear appearance. The morphological profile presents a slight kneading, an effect that may be responsible for the high surface area and nano structuring properties. During the reduction of a GO produced using the Hummers method, the structure that was initially flocked transforms into a structure with a large number of wrinkles and dents throughout the sheet, structural defects caused due to the high reduction of oxygen in the structure and restructuring of the conjugated structure [94]. AA is commonly combined with salicylic acid as a mixed reducing agent, increasing the level of gratification of the rGO produced when compared to other conventional reducing agents, and is used in the production of nanoparticles; as an additional effect, the dispersity is increased, and the morphological profile by spectroscopy transmission reveals the presence of clear areas without the presence of dents, related to the high reducing effect caused by the process [94].
The common reduction process for this hybrid reducer consists of a hydrothermal reduction that assists in the electrical capacity of the rGO. A second method used is vapor reduction that can be used to provide coiled morphologies in order to adhere to some tubular surface, such as threads and fibers. Similarly, the process also helps to reduce the electrical resistance of rGO [95]. The combination of these agents promotes the formation of a material with high electrical conductivity, promoting its use in various sectors. In order to increase the electrical properties of rGOs produced by acetic acid, a binary 3D structure can be formed through vacuum filtration; this ordering tends to increase the charge modulus of the rGO under conditions of low concentrations [96].

1.2.10. Betanin

Betanin (Figure 11) is a pigment belonging to the class of betalains (group of betacyanins), water-soluble nitrogenous pigments widely found in plants of the order Caryophyllales. Betalamic acid is the common precursor of these phenolic compounds, being synthesized from the secondary metabolism of plants and stored mainly in the vacuole of cells [2]. It is considered an excellent antioxidant, acting as a free radical scavenger and an inducer of the antioxidant defense mechanism in cultured cells. Studies indicate its effectiveness in anti-inflammatory, hematoprotective, chemopreventive, and anticancer actions. The substance is found in the root of red beetroot (Beta vulgaris L.), pitaya peel (Hylocereus polyrhizus), quinoa (Chenopodium quinoa Willd), cactus fruits (Opuntia ficus-indica), chard (Beta vulgaris L. ssp. cicla [L.] Alef. Cv. Bright Lights), and uluco (Ullucus tuberosus) [97]. Betanin is considered a promising alternative as a green reducing agent, as the pigment induces the reduction of GO through SN2 nucleophilic reactions and thermal elimination [98].
During a betanin reduction process, the reducing agent is initially extracted through a saline phosphate buffer. With the application of these reducers derived from dragon fruit peel, GO sheets change from a transparent structure covered in wrinkles to a morphology that presents a fine structure, with an increase in wrinkles [98]. The rough morphology is highly affected according to the absorption bands of its absorption spectrum, especially the D (1360 cm−1) and G (1590 cm−1) bands, which when related due to their Id/Ig intensities, a deoxygenation rate and an average reduction of the sp2 carbon structure are obtained. Increasing this ratio implies an increase in GO reduction, causing a progressive increase in the amount of wrinkle morphology on the surface [99].
The reduction time also implies additional variations in morphology, knowing that an increase in the Id/Ig ratio intensifies the formation of C=C bonds, with a high reduction time causing an extremely rough structure to form flaked structures [100]. This differentiated process led to variations in the sample’s absorption peaks, indicating the restoration of the electronic structure; however, it also presented a reduction in the Id/Ig index, leading to a theoretically smaller reduction. The morphology in this zone showed the formation of laminated aggregates in the rGO morphology, with high thickness, indicative of an inferior reduction quality caused by the excessive acidity of graphene oxide [99,100].

2. Perspectives

The growing interest in natural antioxidant agents as a strategy for reducing graphene oxide has been highlighted prominently in the scientific sphere, driven by the potential for biological and environmental applications. It is important to elucidate that these antioxidants, widely available in natural sources such as plants, fruits, and spices, exhibit chemical properties that allow them to interact with and neutralize free radicals as well as other reactive oxygen species associated with graphene oxide. Therefore, the use of compounds such as vitamins C and E, polyphenols, and flavonoids to alter or minimize the reactivity of graphene oxide not only offers a more ecological and safer alternative compared to traditional chemical reducers, but also fosters the development of new techniques for the synthesis and manipulation of graphene-based materials in a sustainable and environmentally friendly manner.

3. Conclusions

Reviewing the antioxidant properties of different natural products in reducing graphene oxide is a promising and important area of research. Graphene oxide is an oxidized form of graphene that can cause damage to cells, but the combination of natural antioxidant substances can help protect cells from the harmful effects of GO. This could have significant implications for human health, especially in relation to diseases related to oxidative stress, such as cancer and cardiovascular disease. However, more research is needed to fully understand how these compounds interact and how they can be applied for therapeutic use. Still, the current research is an important step toward advancing our knowledge about the properties of graphene and how we can use them to improve human health and well-being.

Author Contributions

Conceptualization, F.K.F.d.S. and A.A.M.P.J.; methodology, F.d.A.A. and A.A.M.P.J.; software, C.J.N.F. and D.K.M.I.; resources, F.K.F.d.S., A.A.M.P.J. and A.L.N.F.; data curation, F.K.F.d.S. and D.K.M.I.; writing—original draft preparation, F.K.F.d.S., A.A.M.P.J., P.H.A.O. and V.F.d.V.J.; writing—review and editing, F.K.F.d.S., A.A.M.P.J., P.H.A.O. and V.F.d.V.J.; visualization, F.K.F.d.S., F.d.A.A. and A.L.N.F.; supervision, V.F.d.V.J.; project administration, V.F.d.V.J.; funding acquisition, V.F.d.V.J. All authors have read and agreed to the published version of the manuscript.

Funding

CAPES (Coordination for the Improvement of Higher Education Personnel), CNPq (Brazilian National Council of Research, Grant number 310782/2022-8), and FAPERJ (Rio de Janeiro State Research Support Foundation, Grant Numbers E-26/200.512/2023 and E-26/211.315/2021).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are available on request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Yu, W.; Sisi, L.; Haiyan, Y.; Jie, L. Progress in the functional modification of graphene/graphene oxide: A review. RSC Adv. 2020, 10, 15328–15345. [Google Scholar] [CrossRef] [PubMed]
  2. Kluge, R.A.; Preczenhak, A.P. Betalaínas em beterraba minimamente processada: Perdas e formas de preservação. Rev. Iberoam. Tecnol. Postcosecha 2016, 17, 175–192. [Google Scholar]
  3. Huang, H.; Shi, H.; Das, P.; Qin, J.; Li, Y.; Wang, X.; Su, F.; Wen, P.; Li, S.; Lu, P.; et al. The chemistry and promising applications of graphene and porous graphene materials. Adv. Funct. Mater. 2020, 30, 1909035. [Google Scholar] [CrossRef]
  4. Korkmaz, S.; Kariper, İ.A. Graphene and graphene oxide based aerogels: Synthesis, characteristics and supercapacitor applications. J. Energy Storage 2020, 27, 101038. [Google Scholar] [CrossRef]
  5. Xu, B.; Zhang, J.; Gu, Y.; Zhang, Z.; Al Abdulla, W.; Kumar, N.A.; Zhao, X.S. Lithium-storage properties of gallic acid-reduced graphene oxide and silicon-graphene composites. Electrochim. Acta 2016, 212, 473–480. [Google Scholar] [CrossRef]
  6. Vu, D.L.; Seo, J.S.; Lee, H.Y.; Lee, J.W. Activated carbon with hierarchical micro–mesoporous structure obtained from rice husk and its application for lithium–sulfur batteries. RSC Adv. 2017, 7, 4144–4151. [Google Scholar] [CrossRef]
  7. Georgakilas, V.; Tiwari, J.N.; Kemp, K.C.; Perman, J.A.; Bourlinos, A.B.; Kim, K.S.; Zboril, R. Noncovalent functionalization of graphene and graphene oxide for energy materials, biosensing, catalytic, and biomedical applications. Chem. Rev. 2016, 116, 5464–5519. [Google Scholar] [CrossRef] [PubMed]
  8. Zhang, C.; Nicolosi, V. Graphene and MXene-based transparent conductive electrodes and supercapacitors. Energy Storage Mater. 2019, 16, 102–125. [Google Scholar] [CrossRef]
  9. Zhang, L.; Zhang, F.; Yang, X.; Long, G.; Wu, Y.; Zhang, T.; Leng, K.; Huang, Y.; Ma, Y.; Yu, A.; et al. Porous 3D graphene-based bulk materials with exceptional high surface area and excellent conductivity for supercapacitors. Sci Rep 2013, 3, 1408. [Google Scholar] [CrossRef]
  10. Liu, J.; Cui, L.; Losic, D. Graphene and graphene oxide as new nanocarriers for drug delivery applications. Acta Biomater. 2013, 9, 9243–9257. [Google Scholar] [CrossRef]
  11. Ramírez, I.B.; Carreón, Y.J.P.; Rodríguez-Almazán, C.; Medina-Durán, C.M.; Islas, S.R.; Saniger, J.M. Graphene-Based Biosensors for Molecular Chronic Inflammatory Disease Biomarker Detection. Biosensors 2022, 12, 244. [Google Scholar] [CrossRef] [PubMed]
  12. Kujawska, M.; Bhardwaj, S.K.; Mishra, Y.K.; Kaushik, A. Using Graphene-Based Biosensors to Detect Dopamine for Efficient Parkinson’s Disease Diagnostics. Biosensors 2021, 11, 433. [Google Scholar] [CrossRef] [PubMed]
  13. Kanchanapallya, R.; Nellorea, B.P.V.; Sinhaa, S.S.; Pedrazab, F.; Jonesa, S.J.; Pramanika, A.; Chavvaa, S.R.; Tchounwoua, C.; Shia, Y.; Vangaraa, A.; et al. Antimicrobial peptide-conjugated graphene oxide membrane for efficient removal and effective killing of multiple drug resistant bacteria. RSC Adv. 2015, 5, 18881–18887. [Google Scholar] [CrossRef]
  14. Vasil’ev, R.F.; Veprintsev, T.L.; Dolmatova, L.S.; Naumov, V.V.; Trofimov, A.V.; Tsaplev, Y.B. Kinetics of ethylbenzene oxy-chemiluminescence in the presence of antioxidants from tissues of the marine invertebrate Eupentacta fraudatrix: Estimating the concentration and reactivity of the natural antioxidants. Kinet. Catal. 2014, 55, 148–153. [Google Scholar] [CrossRef]
  15. Fedorova, G.F.; Menshov, V.A.; Trofimov, A.V.; Tsaplev, Y.B.; Vasil’ev, R.F.; Yablonskaya, O.I. Chemiluminescence of cigarette smoke: Salient features of the phenomenon. J. Photochem. Photobiol. A 2017, 93, 579–589. [Google Scholar] [CrossRef]
  16. Flohé, L. Looking back at the early stages of redox biology. Antioxidants 2020, 9, 1254. [Google Scholar] [CrossRef]
  17. Agarwal, V.; Zetterlund, P.B. Strategies for reduction of graphene oxide—A comprehensive review. Chem. Eng. J. 2021, 405, 127018. [Google Scholar] [CrossRef]
  18. Chng, E.L.K.; Pumera, M. The toxicity of graphene oxides: Dependence on the oxidative methods used. Chem. Eur. J. 2013, 19, 8227–8235. [Google Scholar] [CrossRef] [PubMed]
  19. Akhavan, O. Graphene scaffolds in progressive nanotechnology/stem cell-based tissue engineering of the nervous system. J. Mater. Chem. B 2016, 4, 3169–3190. [Google Scholar] [CrossRef]
  20. Kovbasyuk, L.; Mokhir, A. Toxicity studies and biomedical applications of graphene oxide. In Graphene Oxide: Fundamentals and Applications; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2016; pp. 364–381. [Google Scholar] [CrossRef]
  21. An, W.; Zhang, Y.; Zhang, X.; Li, K.; Kang, Y.; Akhtar, S.; Sha, X.; Gao, L. Ocular toxicity of reduced graphene oxide or graphene oxide exposure in mouse eyes. Exp. Eye Res. 2018, 174, 59–69. [Google Scholar] [CrossRef]
  22. De Silva, K.K.H.; Huang, H.H.; Joshi, R.K.; Yoshimura, M. Chemical reduction of graphene oxide using green reductants. Carbon 2017, 119, 190–199. [Google Scholar] [CrossRef]
  23. Soni, J.; Teli, P.; Agarwal, S. Recent Advances in the Green Reduction of Graphene Oxide and its Potential Applications. Curr. Nanosci. 2024, 20, 146–156. [Google Scholar] [CrossRef]
  24. Ismail, Z. Green reduction of graphene oxide by plant extracts: A short review. Ceram. Int. 2019, 45, 23857–23868. [Google Scholar] [CrossRef]
  25. Seabra, A.B.; Paula, A.J.; de Lima, R.; Alves, O.L.; Durán, N. Nanotoxicity of graphene and graphene oxide. Chem. Res. Toxicol. 2014, 27, 159–168. [Google Scholar] [CrossRef]
  26. Pei, S.; Cheng, H.M. The reduction of graphene oxide. Carbon 2012, 50, 3210–3228. [Google Scholar] [CrossRef]
  27. Thakur, S.; Karak, N. Alternative methods and nature-based reagents for the reduction of graphene oxide: A review. Carbon 2015, 94, 224–242. [Google Scholar] [CrossRef]
  28. Dhara, K.; Debiprosad, R.M. Review on nanomaterials-enabled electrochemical sensors for ascorbic acid detection. Anal. Biochem. 2019, 586, 113415. [Google Scholar] [CrossRef]
  29. Santos, M.A.T.D. Efeito do cozimento sobre alguns fatores antinutricionais em folhas de brócoli, couve-flor e couve. Ciência Agrotecnol. 2006, 30, 294–301. [Google Scholar] [CrossRef]
  30. Zhang, L.; Li, X.; Huang, Y.; Ma, Y.; Wan, X.; Chen, Y. Controlled synthesis of few-layered graphene sheets on a large scale using chemical exfoliation. Carbon 2010, 48, 2367–2371. [Google Scholar] [CrossRef]
  31. Kurian, M. Recent progress in the chemical reduction of graphene oxide by green reductants—A Mini review. Carbon Trends 2021, 5, 100120. [Google Scholar] [CrossRef]
  32. Xu, Y.; Shi, G.; Duan, X. Self-assembled three-dimensional graphene macrostructures: Synthesis and applications in supercapacitors. Acc. Chem. Res. 2015, 48, 1666–1675. [Google Scholar] [CrossRef] [PubMed]
  33. Feng, J.; Ye, Y.; Xiao, M.; Wu, G.; Ke, Y. Synthetic routes of the reduced graphene oxide. Chem. Pap. 2020, 74, 3767–3783. [Google Scholar] [CrossRef]
  34. Valipour, A.; Ayyaru, S.; Ahn, Y. Application of graphene-based nanomaterials as novel cathode catalysts for improving power generation in single chamber microbial fuel cells. J. Power Sources 2016, 327, 548–556. [Google Scholar] [CrossRef]
  35. Lee, T.W.; Tsai, I.C.; Liu, Y.F.; Chen, C. Upcycling fruit peel waste into a green reductant to reduce graphene oxide for fabricating an electrochemical sensing platform for sulfamethoxazole determination in aquatic environments. Sci. Total Environ. 2022, 812, 152273. [Google Scholar] [CrossRef]
  36. Khan, M.I. Plant betalains: Safety, antioxidant activity, clinical efficacy, and bioavailability. Compr. Rev. Food Sci. Food Saf. 2016, 15, 316–330. [Google Scholar] [CrossRef]
  37. Tang, Y.; Li, X.; Zhang, B.; Chen, P.X.; Liu, R.; Tsao, R. Characterisation of phenolics, betanins and antioxidant activities in seeds of three Chenopodium quinoa Willd. genotypes. Food Chem. 2015, 166, 380–388. [Google Scholar] [CrossRef] [PubMed]
  38. Esatbeyoglu, T.; Wagner, A.E.; Motafakkerazad, R.; Nakajima, Y.; Matsugo, S.; Rimbach, G. Free radical scavenging and antioxidant activity of betanin: Electron spin resonance spectroscopy studies and studies in cultured cells. Food Chem. Toxicol. 2014, 73, 119–126. [Google Scholar] [CrossRef]
  39. Dorniani, D.; Saifullah, B.; Barahuie, F.; Arulselvan, P.; Hussein, M.Z.B.; Fakurazi, S.; Twyman, L.J. Graphene oxide-gallic acid nanodelivery system for cancer therapy. Nanoscale Res. Lett. 2016, 11, 491. [Google Scholar] [CrossRef]
  40. Pham, T.A.; Kim, J.S.; Kim, J.S.; Jeong, Y.T. One-step reduction of graphene oxide with L-glutathione. Colloids Surf. A Physicochem. Eng. Asp. 2011, 384, 543–548. [Google Scholar] [CrossRef]
  41. Chhetri, S.; Adak, N.C.; Samanta, P.; Murmu, N.C.; Hui, D.; Kuila, T.; Lee, J.H. Investigation of the mechanical and thermal properties of l-glutathione modified graphene/epoxy composites. Compos. Part B Eng. 2018, 143, 105–112. [Google Scholar] [CrossRef]
  42. Dhanalakshmi, N.; Priya, T.; Thennarasu, S.; Sivanesan, S.; Thinakaran, N. Synthesis and electrochemical properties of environmental free l-glutathione grafted graphene oxide/ZnO nanocomposite for highly selective piroxicam sensing. J. Pharm. Anal. 2021, 11, 48–56. [Google Scholar] [CrossRef] [PubMed]
  43. Gaucher, C.; Boudier, A.; Bonetti, J.; Clarot, I.; Leroy, P.; Parent, M. Glutathione: Antioxidant properties dedicated to nanotechnologies. Antioxidants 2018, 7, 62. [Google Scholar] [CrossRef]
  44. Cominetti, C.; Bortoli, M.D.; Abdalla, D.S.P.; Cozzolino, S.M.F. Considerations about oxidative stress, selenium and nutrigenetics. Nutr. Rev. Soc. Bras. Aliment. Nutr. 2011, 36, 131–153. [Google Scholar]
  45. Wang, Z.; Xu, C.; Li, X.; Liu, Z. In situ green synthesis of Ag nanoparticles on tea polyphenols-modified graphene and their catalytic reduction activity of 4-nitrophenol. Colloids Surf. A Physicochem. Eng. Asp. 2015, 485, 102–110. [Google Scholar] [CrossRef]
  46. Li, B.; Jin, X.; Lin, J.; Chen, Z. Green reduction of graphene oxide by sugarcane bagasse extract and its application for the removal of cadmium in aqueous solution. J. Clean. Prod. 2018, 189, 128–134. [Google Scholar] [CrossRef]
  47. dos Santos, F.K.F.; de Almeida Cavalcante, S.F.; Rezende, C.M.; Veiga-Junior, V.F. Extraction and enrichment of furanoid glycosylated diterpenes present in green coffee beans: New methodology in the QuEChERS concept. J. Food Compos. Anal. 2021, 104, 104142. [Google Scholar] [CrossRef]
  48. Liu, H.; Zou, T.; Gao, J.M.; Gu, L. Depolymerization of cranberry procyanidins using (+)-catechin,(−)-epicatechin, and (−)-epigallocatechin gallate as chain breakers. Food Chem. 2013, 141, 488–494. [Google Scholar] [CrossRef]
  49. Fan, X.; Peng, W.; Li, Y.; Li, X.; Wang, S.; Zhang, G.; Zhang, F. Deoxygenation of exfoliated graphite oxide under alkaline conditions: A green route to graphene preparation. Adv. Mater. 2008, 20, 4490–4493. [Google Scholar] [CrossRef]
  50. Iqbal, K.; Khan, A.; Khattak, M.M.A.K. Biological significance of ascorbic acid (vitamin C) in human health-a review. Pak. J. Nutr. 2004, 3, 5–13. [Google Scholar]
  51. Gęgotek, A.; Skrzydlewska, E. Antioxidative and Anti-Inflammatory Activity of Ascorbic Acid. Antioxidants 2022, 11, 1993. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  52. Smirnoff, N. Ascorbic acid metabolism and functions: A comparison of plants and mammals. Free Radic. Biol. Med. 2018, 122, 116–129. [Google Scholar] [CrossRef] [PubMed]
  53. Zainuddin, M.F.; Nik Raikhan, N.H.; Othman, N.H.; Abdullah, W.F.H. Synthesis of reduced Graphene Oxide (rGO) using different treatments of Graphene Oxide (GO). IOP Conf. Ser. Mater. Sci. Eng. 2018, 358, 012046. [Google Scholar] [CrossRef]
  54. Yao, L.; Zhou, C.; Hu, N.; Hu, J.; Hong, M.; Zhang, L.; Zhang, Y. Flexible graphene/carbon nanotube hybrid papers chemical-reduction-tailored by gallic acid for high-performance electrochemical capacitive energy storages. Appl. Surf. Sci. 2018, 435, 699–707. [Google Scholar] [CrossRef]
  55. Verma, S.; Singh, A.; Mishra, A. Gallic acid: Molecular rival of cancer. Environ. Toxicol. Pharmacol. 2013, 35, 473–485. [Google Scholar] [CrossRef]
  56. de Jesus Benevides, C.M.; Souza, M.V.; Souza, R.D.B.; Lopes, M.V. Fatores antinutricionais em alimentos: Revisão. Segurança Aliment. Nutr. 2011, 18, 67–79. [Google Scholar] [CrossRef]
  57. Serna-Escolano, V.; Giménez, M.J.; Castillo, S.; Valverde, J.M.; Martínez-Romero, D.; Guillén, F.; Serrano, M.; Valero, D.; Zapata, P.J. Preharvest treatment with oxalic acid improves postharvest storage of lemon fruit by stimulation of the antioxidant system and phenolic content. Antioxidants 2021, 10, 963. [Google Scholar] [CrossRef] [PubMed]
  58. Song, P.; Zhang, X.; Sun, M.; Cui, X.; Lin, Y. Synthesis of graphene nanosheets via oxalic acid-induced chemical reduction of exfoliated graphite oxide. RSC Adv. 2012, 2, 1168–1173. [Google Scholar] [CrossRef]
  59. El-Hout, S.I.; El-Sheikh, S.M.; Hassan, H.M.; Harraz, F.A.; Ibrahim, I.A.; El-Sharkawy, E.A. A green chemical route for synthesis of graphene supported palladium nanoparticles: A highly active and recyclable catalyst for reduction of nitrobenzene. Appl. Catal. A Gen. 2015, 503, 176–185. [Google Scholar] [CrossRef]
  60. Diggikar, R.S.; Late, D.J.; Kale, B.B. Unusual morphologies of reduced graphene oxide and polyaniline nanofibers-reduced graphene oxide composites for high performance supercapacitor applications. RSC Adv. 2014, 4, 22551–22560. [Google Scholar] [CrossRef]
  61. Wei, T. Oral delivery of glutathione: Antioxidant function, barriers and strategies. Acta Mater. Med. 2022, 1, 177–192. [Google Scholar] [CrossRef]
  62. Barbosa, D.S. Green Tea Polyphenolic Compounds and Human Health. J. Verbraucherschutz Leb. 2007, 2, 407–413. [Google Scholar] [CrossRef]
  63. Yan, Z.; Zhong, Y.; Duan, Y.; Chen, Q.; Li, F. Antioxidant mechanism of tea polyphenols and its impact on health benefits. Anim. Nutr. 2020, 6, 115–123. [Google Scholar] [CrossRef] [PubMed]
  64. Patel, R.; Salamone, G.; Macwan, I. The Role of Graphene Monolayers in Enhancing the Yield of Bacteriorhodopsin Photostates for Optical Memory Applications. Appl. Sci. 2021, 11, 9698. [Google Scholar] [CrossRef]
  65. Wong, X.Y.; Quesada-González, D.; Manickam, S.; New, S.Y.; Muthoosamy, K.; Merkoçi, A. Integrating gold nanoclusters, folic acid and reduced graphene oxide for nanosensing of glutathione based on “turn-off” fluorescence. Sci. Rep. 2021, 11, 2375. [Google Scholar] [CrossRef]
  66. Akhavan, O.; Kalaee, M.; Alavi, Z.S.; Ghiasi, S.M.A.; Esfandiar, A. Increasing the antioxidant activity of green tea polyphenols in the presence of iron for the reduction of graphene oxide. Carbon 2012, 50, 3015–3025. [Google Scholar] [CrossRef]
  67. Singh, P.; Singh, S.; Jaggi, N.; Kim, K.H.; Devi, P. Recent advances in bacteriorhodopsin-based energy harvesters and sensing devices. Nano Energy 2021, 79, 105482. [Google Scholar] [CrossRef]
  68. Akhavan, O. Bacteriorhodopsin as a superior substitute for hydrazine in chemical reduction of single-layer graphene oxide sheets. Carbon 2015, 81, 158–166. [Google Scholar] [CrossRef]
  69. Bao, Q.X.; Liu, Y.; Liang, Y.Q.; Weerasooriya, R.; Li, H.; Wu, Y.C.; Chen, X. Tea polyphenols mediated Zero-valent Iron/Reduced graphene oxide nanocomposites for electrochemical determination of Hg2+. J. Electroanal. Chem. 2022, 917, 116428. [Google Scholar] [CrossRef]
  70. Abdullah, M.F.; Zakaria, R.; Zein, S.H.S. Green tea polyphenol–reduced graphene oxide: Derivatisation, reduction efficiency, reduction mechanism and cytotoxicity. RSC Adv. 2014, 4, 34510–34518. [Google Scholar] [CrossRef]
  71. Liu, H.; Li, T.; Liu, Y.; Qin, G.; Wang, X.; Chen, T. Glucose-reduced graphene oxide with excellent biocompatibility and photothermal efficiency as well as drug loading. Nanoscale Res. Lett. 2016, 11, 211. [Google Scholar] [CrossRef] [PubMed]
  72. Sunil, K.C.; Suvarna, S.; Nairy, R.K.; Chethan, G.; Mustak, M.S.; Yerol, N. Facile, cost-effective and eco-friendly synthesis of carbonyl-rich partially reduced graphene oxide using glucose as a sole precursor. SN Appl. Sci. 2020, 2, 1323. [Google Scholar] [CrossRef]
  73. Lu, W.; Qin, F.; Wang, Y.; Luo, Y.; Wang, H.; Scarpa, F.; Li, J.; Sesana, R.; Cura, F.; Peng, H.X. Engineering graphene wrinkles for large enhancement of interlaminar friction enabled damping capability. ACS Appl. Mater. Interfaces 2019, 11, 30278–30289. [Google Scholar] [CrossRef] [PubMed]
  74. Han, F.; Qian, O.; Chen, B.; Tang, H.; Wang, M. Sugar blowing-assisted reduction and interconnection of graphene oxide into three-dimensional porous graphene. J. Alloys Compd. 2018, 730, 386–391. [Google Scholar] [CrossRef]
  75. Wang, X.; Zhang, Y.; Zhi, C.; Wang, X.; Tang, D.; Xu, Y.; Weng, Q.; Jiang, X.; Mitome, M.; Golberg, D.; et al. Three-dimensional strutted graphene grown by substrate-free sugar blowing for high-power-density supercapacitors. Nat. Commun. 2013, 4, 2905. [Google Scholar] [CrossRef] [PubMed]
  76. Espíndola, K.M.M.; Ferreira, R.G.; Narvaez, L.E.M.; Silva Rosario, A.C.R.; Da Silva, A.H.M.; Silva, A.G.B.; Vieira, A.P.O.; Monteiro, M.C. Chemical and pharmacological aspects of caffeic acid and its activity in hepatocarcinoma. Front. Oncol. 2019, 9, 541. [Google Scholar] [CrossRef] [PubMed]
  77. dos Santos, F.K.F.; Silva, I.G.C.B.; Nunes Filho, A.L.; da Veiga Júnior, V.F. Coffee and Folk Medicine: Mechanisms and Activities. In Ethnobotany; CRC Press: Boca Raton, FL, USA, 2023; pp. 65–80. [Google Scholar]
  78. Sova, M. Antioxidant and antimicrobial activities of cinnamic acid derivatives. Mini Rev. Med. Chem. 2012, 12, 749–767. [Google Scholar] [CrossRef]
  79. Silva, R.O.; Freitas Filho, J.D.; Freitas, J. glicose, uma biomolécula fascinante: História, propriedades, produção e aplicação. Rev. Virtual Quim. 2018, 10, 875–891. [Google Scholar]
  80. Riaz, U.; Kharal, M.A.; Murtaza, G.; uz Zaman, Q.; Javaid, S.; Malik, H.A.; Aziz, H.; Abbas, Z. Prospective roles and mechanisms of caffeic acid in counter plant stress: A mini review. Pak. J. Agric. Res. 2019, 32, 8–19. [Google Scholar] [CrossRef]
  81. Hernández-Chávez, G.; Martinez, A.; Gosset, G. Metabolic engineering strategies for caffeic acid production in Escherichia coli. Electron. J. Biotechnol. 2019, 38, 19–26. [Google Scholar] [CrossRef]
  82. Mirzaei, S.; Gholami, M.H.; Zabolian, A.; Saleki, H.; Farahani, M.V.; Hamzehlou, S.; Far, F.B.; Sharifzadeh, S.O.; Samarghandian, S.; Khan, H.; et al. Caffeic acid and its derivatives as potential modulators of oncogenic molecular pathways: New hope in the fight against cancer. Pharmacol. Res. 2021, 171, 105759. [Google Scholar] [CrossRef] [PubMed]
  83. Muhammad Abdul Kadar, N.N.; Ahmad, F.; Teoh, S.L.; Yahaya, M.F. Caffeic acid on metabolic syndrome: A review. Molecules 2021, 26, 5490. [Google Scholar] [CrossRef]
  84. Sato, Y.; Itagaki, S.; Kurokawa, T.; Ogura, J.; Kobayashi, M.; Hirano, T.; Sugawara, M.; Iseki, K. In vitro and in vivo antioxidant properties of chlorogenic acid and caffeic acid. Int. J. Pharm. 2011, 403, 136–138. [Google Scholar] [CrossRef]
  85. dos Santos, F.K.F.; de Rezende, C.M.; da Veiga Júnior, V.F. Macroporous polymeric resins as a tool to obtain bioactive compounds in food and food-waste: A review. J. Food Compos. Anal. 2022, 114, 104703. [Google Scholar] [CrossRef]
  86. Liu, H.; Wang, L.; Li, R.; Shi, B.; Wang, P.; Zhao, Y.; Zhang, X. Modulated Crystallization and Reduced V OC Deficit of Mixed Lead–Tin Perovskite Solar Cells with Antioxidant Caffeic Acid. ACS Energy Lett. 2021, 6, 2907–2916. [Google Scholar] [CrossRef]
  87. Sakthinathan, S.; Keyan, A.K.; Rajakumaran, R.; Chen, S.M.; Chiu, T.W.; Dong, C.; Vinothini, S. Synthesis of N-rGO-MWCNT/CuCrO2 catalyst for the bifunctional application of hydrogen evolution reaction and electrochemical detection of Bisphenol-A. Catalysts 2021, 11, 301. [Google Scholar] [CrossRef]
  88. Mei, X.; Meng, X.; Wu, F. Hydrothermal method for the production of reduced graphene oxide. Phys. E Low-Dimens. Syst. Nanostruct. 2015, 68, 81–86. [Google Scholar] [CrossRef]
  89. Bo, Z.; Shuai, X.; Mao, S.; Yang, H.; Qian, J.; Chen, J.; Yan, J.; Cen, K. Green preparation of reduced graphene oxide for sensing and energy storage applications. Sci. Rep. 2014, 4, 46848. [Google Scholar] [CrossRef] [PubMed]
  90. Budiman, A.W.; Nam, J.S.; Park, J.H.; Mukti, R.I.; Chang, T.S.; Bae, J.W.; Choi, M.J. Review of acetic acid synthesis from various feedstocks through different catalytic processes. Catal. Surv. Asia 2016, 20, 173–193. [Google Scholar] [CrossRef]
  91. Pal, P.; Nayak, J. Acetic acid production and purification: Critical review towards process intensification. Sep. Purif. Rev. 2017, 46, 44–61. [Google Scholar] [CrossRef]
  92. Sarchami, T.; Batta, N.; Berruti, F. Production and separation of acetic acid from pyrolysis oil of lignocellulosic biomass: A review. Biofuels Bioprod. Biorefin. 2021, 15, 1912–1937. [Google Scholar] [CrossRef]
  93. Bartowsky, E.J.; Henschke, P.A. Acetic acid bacteria spoilage of bottled red wine—A review. Int. J. Food Microbiol. 2008, 125, 60–70. [Google Scholar] [CrossRef]
  94. Santos, H.O.; de Moraes, W.M.; da Silva, G.A.; Prestes, J.; Schoenfeld, B.J. Vinegar (acetic acid) intake on glucose metabolism: A narrative review. Clin. Nutr. ESPEN 2019, 32, 1–7. [Google Scholar] [CrossRef]
  95. La China, S.; Zanichelli, G.; De Vero, L.; Gullo, M. Oxidative fermentations and exopolysaccharides production by acetic acid bacteria: A mini review. Biotechnol. Lett. 2018, 40, 1289–1302. [Google Scholar] [CrossRef]
  96. Valipour, A.; Hamnabard, N.; Ahn, Y.H. Performance evaluation of highly conductive graphene (RGO HI–AcOH) and graphene/metal nanoparticle composites (RGO/Ni) coated on carbon cloth for supercapacitor applications. RSC Adv. 2015, 5, 92970–92979. [Google Scholar] [CrossRef]
  97. Wang, D.; Li, D.; Zhao, M.; Xu, Y.; Wei, Q. Multifunctional wearable smart device based on conductive reduced graphene oxide/polyester fabric. Appl. Surf. Sci. 2018, 454, 218–226. [Google Scholar] [CrossRef]
  98. Moon, I.K.; Lee, J.; Ruoff, R.S.; Lee, H. Reduced graphene oxide by chemical graphitization. Nat. Commun. 2010, 1, 73. [Google Scholar] [CrossRef] [PubMed]
  99. Zheng, Y.; Chen, W.; Wang, Z.; Wang, Q. Polystyrene/rGO composites with orientation-3d network binary structure and its surprising conductivity. Macromolecules 2018, 51, 7993–8000. [Google Scholar] [CrossRef]
  100. Lourith, N.; Kanlayavattanakul, M. Antioxidant and stability of dragon fruit peel colour. Agro Food Ind. Hi Tech 2013, 24, 56–58. [Google Scholar]
Figure 1. Chemical reduction of GO. Elaborated by the authors.
Figure 1. Chemical reduction of GO. Elaborated by the authors.
Ijms 25 05182 g001
Figure 2. Chemical structure of ascorbic acid. Elaborated by the authors.
Figure 2. Chemical structure of ascorbic acid. Elaborated by the authors.
Ijms 25 05182 g002
Figure 3. Chemical structure of gallic acid. Elaborated by the authors.
Figure 3. Chemical structure of gallic acid. Elaborated by the authors.
Ijms 25 05182 g003
Figure 4. Chemical structure of oxalic acid. Elaborated by the authors.
Figure 4. Chemical structure of oxalic acid. Elaborated by the authors.
Ijms 25 05182 g004
Figure 5. Chemical structure of L-glutathione. Elaborated by the authors.
Figure 5. Chemical structure of L-glutathione. Elaborated by the authors.
Ijms 25 05182 g005
Figure 6. FTIR spectrum of GO, graphene, and other elements. Elaborated by the authors.
Figure 6. FTIR spectrum of GO, graphene, and other elements. Elaborated by the authors.
Ijms 25 05182 g006
Figure 7. Morphology of GO. Elaborated by the authors.
Figure 7. Morphology of GO. Elaborated by the authors.
Ijms 25 05182 g007
Figure 8. Main polyphenols in green tea. Elaborated by the authors.
Figure 8. Main polyphenols in green tea. Elaborated by the authors.
Ijms 25 05182 g008
Figure 9. Illustration of process of reduction of GO with green tea polyphenols (GTPs). Elaborated by the authors.
Figure 9. Illustration of process of reduction of GO with green tea polyphenols (GTPs). Elaborated by the authors.
Ijms 25 05182 g009
Figure 10. Main hydroxycinnamic acids described in coffee. Elaborated by the authors.
Figure 10. Main hydroxycinnamic acids described in coffee. Elaborated by the authors.
Ijms 25 05182 g010
Figure 11. Chemical structure of betanin. Elaborated by the authors.
Figure 11. Chemical structure of betanin. Elaborated by the authors.
Ijms 25 05182 g011
Table 1. Examples of the application of ascorbic acid as a reducing agent.
Table 1. Examples of the application of ascorbic acid as a reducing agent.
Ref.SampleReduction MethodReduction TemperatureEquipmentObservations[AA][GO]
[1]Graphite powderChemical70 °C, 120 °C, and 300 °CXRD, XPS, UV-Vis-NIROptical performance analysis of rGO thin films with AA comparing them with hydrazine at the same concentrations.5, 25, 50, 100, 200, 400, and 800 mM10 mg/mL
[3]Graphite flakesChemical90 °C/24 hXRD, XPS, FE-SEM, DLSStructural, electrical, and optical analysis of rGO films with AA.30 mg/mL32 and 100 mg/mL
[4]Monolayer grapheneChemical70 °C/5 minSEM, Raman, FTIR, XRDProduction of centimeter-scale semipermeable membranes with ion blocking efficiency of up to 9.0.4 g/mL4 mg/mL
[34]Natural graphiteChemicalRoom temperatureUV-Vis-NIR, FT-IR, XPSAA as photosensitive reductant of GO5 mg/mL0.1 mg/mL
[6]Natural graphiteChemical65 °C/1 hDRX, XPS, SEM, AFM, ATR-FT-IRCharacterization of reduced GO by AA in different periods100 mg/mL0.1 mg/mL
[7]Multilayer GOHydrothermal95 °C/6 hSEM, XRD, FT-IRFabrication of three-dimensional reduced GO aerogels (rGOAs) using AA as a reductant240 mg/mL4.0 mg/mL
[28]Expanded graphiteHydrothermal80 °C/24 hXPS, Raman, TEM, SEMHydrogel with piezoresistive property manufactured with GO reduced with Pt and AA nanoparticles10 mg/mL5, 10, and 15 mg/mL
Table 2. Examples of the application of oxalic acid as a reducing agent.
Table 2. Examples of the application of oxalic acid as a reducing agent.
Ref.SampleReduction MethodReduction TemperatureEquipmentObservations[OA][GO]
[35]Graphite powderChemical75 °C/18 hTEM, SEM, FT-IR, XRDSynthesis of graphene nanosheets via oxalic acid-induced chemical reduction75.6 mg/mL2.5 mg/mL
[36]Natural graphiteChemical900 °C/2 hXRDExfoliation and reduction method10 mg/mL2 mg/mL
[29]Natural graphite powderHydrothermal160 °C/5 minSEM, Raman, FTIR, XRDProduction of centimeter-scale semipermeable membranes with ion blocking efficiency of up to 90%6, 12, and 25 g/mL1.5 mg/mL
[33]Synthetic graphiteHydrothermal100 °C/1 hDRX, FT-IR, SEM, XPSComposite of rGO and bismuth as electrode for super capacitor18 mg/mL1.5 mg/mL
[37]GOHydrothermal Chemical180 °C/24 hUV-Vis-NIR, FT-IR, XPSAA as photosensitive reductant of GO4.4 mg/mL0.8 mg/mL
[38]Natural graphiteHydrothermal 200 °C/12 h DRX, XPS, SEM, AFM, ATR-FT-IRCharacterization of reduced GO, by AA in different periodsNINI
[39]Multilayer GOChemical 120 °C/13 h SEM, XRD, FT-IRFabrication of three-dimensional reduced GO aerogels (rGOAs) using AA as a reductantNINI
[40]Expanded graphiteHydrothermal50 °C/12 hXPS, Raman, TEM, SEMHydrogel with piezoresistive property manufactured with GO reduced with Pt and AA nanoparticlesNINI
Table 3. Examples of the application of gallic acid as a reducing agent.
Table 3. Examples of the application of gallic acid as a reducing agent.
Ref.SampleReduction MethodReduction TemperatureEquipmentObservations[GA][GO]
[1]Graphite powderHydrothermal 50 °C/12 h FT-IR, SEMNanoformulation based on graphene oxide loaded with gallic acid (GAGO)0.25 g/mL0.05 g/mL
[3]Graphite powderChemical 45 °C/2 h FT-IR, TEM, XRD, DLSGreen synthesis of metal nanoparticles10 mL of 2 mM GA3.4 mL of 1 mg/mL GO
[4]Graphite powderChemicalRoom temperature 24 hUV-Vis, IR, Raman, XPSGallic acid as a GO reducer and stabilizer4000 mg/mL4000 mg/mL
[7]Graphite powderChemical100 °C/8 hFESEM, XPS, XRD, TGA, UV-Vis DRSGraphene hydrogel with gallic acid as an adsorbent of wastewater pollutants2 mg/mL2 mg/mL
[28]GOHydrothermal190 °C/5 hTEM, DRX, XPSAg nanoparticles (AgNP) are added to rGO sheets using gallic acid-capped AuNPs.5 mg/mL2.5 mg/mL
[29]Carbon nanotubesChemical Room temperature UV-Vis TEM, FT-IR, FT-RamanMWCNT-rGO nanocomposite electrode for sensitive detection of gallic acid capped AuNPs10 mM and 100 mM0.2 mg/mL
[40]Multilayer GOChemical 80 °C/24 h FE-SEM, TEM, FT-IR, XRD, XPSGraphene nanotube hybrid papers for high-performance electrochemical capacitive energy storage480 mg/mL0.1 mg/mL
Table 4. Examples of the application of GSH as a reducing agent.
Table 4. Examples of the application of GSH as a reducing agent.
Ref.SampleReduction MethodReduction TemperatureEquipmentObservations[GSH][GO]
[41]Natural graphite powderChemical 50 °C/6 h TEM, SEM, FT-IR, XRDPreparation of graphene nanosheets from GO and l-glutathione2 mg/mL0.1 mg/mL
[42]Natural graphite flakesChemical 50 °C/5 h FE-SEM, FT-IR, XRD, TGAL-glutathione modified graphene/epoxy composites1200 mg/mL400 mg/mL
[43]Natural graphite powderChemical60 °C/2 hXRD, FT-IR, XPS, FE-SEM, EIS, DPVGO/ZnO nanocomposite with glutathione as electrode for piroxicam sensor1 g/mL1 mg/mL
[44]Synthetic graphiteChemical Room temperature EIS, SEM, EDXComposite of rGO and bismuth as electrode for super capacitor0.03 mg/mL0.05 mg/mL
Table 5. Examples of the application of polyphenols as a reducing agent.
Table 5. Examples of the application of polyphenols as a reducing agent.
Ref.SampleReduction MethodReduction TemperatureEquipmentObservations[Polyphenol][GO]
[45]Natural graphite powderChemical 80 °C/8 h FE-SEM, FT-IR, UV-Vis, XRDHigh-efficiency anticancer drug delivery systems using functionalized RGO tea polyphenolsNI500 mg/L
[46]Natural graphite flakesChemical Room temperature XPS, AFM, RamanAntioxidant activity of green tea polyphenols in the presence of iron for the reduction of GO100 mg/L100 mg/L
[47]Natural graphite powderChemical90 °C/8 hUv-Vis, FT-IR, XPS, TGAMechanism of rGO reduction and cytotoxicity by green tea polyphenol (GPT-rGO)NI300 mg/L
[38]Synthetic graphiteChemical Room temperature/24 h FE-SEM, TEM, EIS, SWASV, XPSTea polyphenol-mediated rGO/iron nanocomposite for electrochemical determination of Hg2+5000 mg/mL10,000 mg/mL
[39]GO powderChemical80 °C/2 hUv-Vis, TEM, FEL, WXRD, FT-IR, XPSIn situ synthesis of Ag nanoparticles in GO modified with tea polyphenols2835 mg/L500 mg/L
Table 6. Examples of glucose application as a reducing agent.
Table 6. Examples of glucose application as a reducing agent.
Ref.SampleReduction MethodReduction TemperatureEquipmentObservations[Glucose][GO]
[47]GOChemical 95 °C/1 h TEM, FTIR, XRD, TGASynthesis of chemically converted GNS based on reducing sugars and GO as precursor0.1 mg/mL0.1 mg/L
[40]Natural graphiteChemical 95 °C/30 min AFM, UV-Vis-NIR, XPSGlucose-reduced GO sheets in photothermal therapy of prostate cancer cells100 μL1 and 0.05 mg/L
[1]Pure graphiteChemicalRoom temperature/24 hUv-Vis, FT-IR, XPS, EDX, XRDAg-Cu2O nanocomposites supported on GO with glucose reduction (rGO) and visible light photocatalytic activity15 mL50 mg/L
[32]Graphite flakesChemical135 °C/30 minAFM, FT-IR, UV-Vis, NIRrGO for glucose excellent biocompatibility and photothermal efficiency and drug loading22 mg/mL1 mg/mL
[33]GOHydrothermal160 °C/12 hXRD, XPS, FT-IR, UV-Vis, Raman, SEM, and DLSGO rich in carbonyl functional groups partially reduced using glucose as the only precursor50 mLNI
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

dos Santos, F.K.F.; Júnior, A.A.M.P.; Filho, A.L.N.; Fonseca, C.J.N.; Isidorio, D.K.M.; Araújo, F.d.A.; Oliveira, P.H.A.; Veiga Júnior, V.F.d. Graphene and Natural Products: A Review of Antioxidant Properties in Graphene Oxide Reduction. Int. J. Mol. Sci. 2024, 25, 5182. https://doi.org/10.3390/ijms25105182

AMA Style

dos Santos FKF, Júnior AAMP, Filho ALN, Fonseca CJN, Isidorio DKM, Araújo FdA, Oliveira PHA, Veiga Júnior VFd. Graphene and Natural Products: A Review of Antioxidant Properties in Graphene Oxide Reduction. International Journal of Molecular Sciences. 2024; 25(10):5182. https://doi.org/10.3390/ijms25105182

Chicago/Turabian Style

dos Santos, Filipe Kayodè Felisberto, Antônio Augusto Martins Pereira Júnior, Arquimedes Lopes Nunes Filho, Clícia Joanna Neves Fonseca, Daysianne Kessy Mendes Isidorio, Filipe de Almeida Araújo, Pablo Henrique Ataide Oliveira, and Valdir Florêncio da Veiga Júnior. 2024. "Graphene and Natural Products: A Review of Antioxidant Properties in Graphene Oxide Reduction" International Journal of Molecular Sciences 25, no. 10: 5182. https://doi.org/10.3390/ijms25105182

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

Article Metrics

Back to TopTop