Next Article in Journal
Effect of Graphene on the Microstructure and Mechanical Properties of WC-Based Cemented Carbide
Next Article in Special Issue
Ultraviolet Exposure Improves SERS Activity of Graphene-Coated Ag/ZrO2 Substrates
Previous Article in Journal
Dynamic Analysis of Multilayered Piezoelectric Quasicrystal Three-Dimensional Sector Plates with Imperfect Interfaces
Previous Article in Special Issue
Laser-Induced Graphene in Polyimide for Antenna Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Graphene Production and Biomedical Applications: A Review

by
Klaudia Malisz
and
Beata Świeczko-Żurek
*
Department of Biomaterials Technology, Faculty of Mechanical Engineering and Ship Technology, Gdansk University of Technology, Gabriela Narutowicza 11/12, 80-229 Gdansk, Poland
*
Author to whom correspondence should be addressed.
Crystals 2023, 13(10), 1413; https://doi.org/10.3390/cryst13101413
Submission received: 28 July 2023 / Revised: 21 August 2023 / Accepted: 23 August 2023 / Published: 23 September 2023
(This article belongs to the Special Issue Advanced Technologies in Graphene-Based Materials)

Abstract

:
Graphene is a two-dimensional nanomaterial composed of carbon atoms with sp2 hybrid orbitals. Both graphene and graphene-based composite have gained broad interest among researchers because of their outstanding physiochemical, mechanical, and biological properties. Graphene production techniques are divided into top-down and bottom-up synthesis methods, of which chemical vapor deposition (CVD) is the most popular. The biomedical applications of graphene and its composite include its use in sensors, implantology, and gene and drug delivery. They can be used for tissue engineering, anticancer therapies, and as antimicrobial agents in implant application. The biocompatibility of graphene-based nanomaterials enables their use in the field of biomedicine. This article reviews the properties of graphene, the methods used to produce it, the challenges associated with its use, and the potential applications of this material in biomedicine, regenerative medicine, and drug delivery systems.

1. Introduction

Carbon nanostructures have attracted a lot of attention, especially single-walled carbon nanotubes, multi-walled carbon nanotubes, graphene, and reduced graphene oxide (rGO) [1]. Graphene is a two-dimensional material composed of carbon (C) atoms with an sp2 hybrid orbital [2,3,4,5], displaying a single π orbital and three σ bonds perpendicular to the plane [6]. It is a honeycomb like structure formed with a single thick planer carbon sheet [7]. Graphene can be wrapped up into zero-dimensional fullerenes, rolled into one-dimensional carbon nanotubes, or piled together with an inter-planar spacing of 0.335 nm to form three-dimensional graphite [6,8]. Due to its outstanding physical, mechanical, electrical, and chemical properties, materials based on graphene have attracted a lot of interest. These include excellent thermal and electrical conductivity, high specific surface area, elastic moduli (about 1 TPa [9]), an intrinsic strength of 130 GPa [5], adaptability to both flat and irregular surfaces, flexibility in chemical and biological functionalization, and simplicity in mass production, [10,11,12,13,14,15]. Sensors, energy harvesting, and storage devices such as solar cells and supercapacitors can all be improved by the use of graphene, as can lightweight polymer composites, membranes, and actuators. Graphene derivatives could also be used in biomedical applications such as drug delivery systems, gene therapies, photothermal therapies, antibacterial agents, and bioimaging tools [5,10,16,17,18]. However, scientists still face numerous challenges when attempting to utilize graphene such as cytotoxicity, biodistribution, and immunological responses.
Currently, there are many ways to produce graphene. Those include chemical vapor deposition, the mechanical cleavage of highly ordered pyrolytic graphite (HOPG), the chemical reduction of chemically exfoliated graphene oxide, epitaxial growth, and chemical synthesis [11,14]. Graphene oxide (GO), reduced graphene oxide (rGO), graphene, and graphene quantum dots (GQDs) are all included in the graphene family [19]. The molecular structures are shown in Figure 1 [20]. Chemical modification can reduce GO to rGO [18]. GQDs, also known as nano-graphene (NG), are nanomaterials that can be used in a range of applications. To create NG, top-down and bottom-up techniques are used [18,21]. Furthermore, there is a separate kind of carbon-based nanomaterial called nano-graphene oxide (NGO). NGO is created by top-down methods. Graphene oxide can be oxidized to form nano-graphene oxide, which has hydrophobic properties and the qualities suitable for creating very stable aqueous colloidal suspensions [18].
Both GO and rGO are well recognized for their lower dispersion in water but also easy aggregation. This can reduce their surface areas, which would limit their usability and ability to be recycled from treated water. Additionally, graphene composites have significant limitations because of their hydrophobic qualities, which tend to result in the aggregation of sheets, along with curled, folded, and corrugated formations on base materials [19]. However, because of their enormous clinical potential in tissue regeneration therapy, GO and rGO-based scaffolds are particularly promising. The proliferation and differentiation of applied stem cells have been strongly influenced by both GO and rGO, but there are still some challenges to overcome, e.g., cytotoxicity, biodistribution, biotransformation, and immune response [17]. Graphene-reinforced polymer composites are also of great interest. The use of graphene as a reinforcing agent in the polymer matrix improves the composites’ properties (for example, adding carbon nanotubes and graphene to metals reduces the coefficient of friction and the wear rate while increasing tensile strength) [5,22,23].
In this review, we will describe the properties of graphene and graphene composites, and the methods used to manufacture them with a particular emphasis on the challenges and potential applications of these materials in biomedicine, tissue engineering, and drug carrier systems.

2. Graphene Production Techniques

Several methods including pyrolysis, epitaxial growth, the chemical and electrochemical exfoliation of graphite, physical vapor deposition, and chemical vapor deposition have been used to produce graphene [6,11,14]. We divide these techniques into top-down (mechanical exfoliation, chemical exfoliation, chemical synthesis) and bottom-up synthesis methods (epitaxial growth, pyrolysis, CVD, etc.) [6]. To obtain graphene, it is possible to use the thermal expansion of graphite/graphite oxide [8,24], which can lead to its formation. However, this process rarely leads to results in the complete exfoliation of graphene to the atomic level of individual sheets of graphene [8]. There are also known methods such as liquid phase exfoliation of graphite employing ultrasonication, chemical exfoliation methods, or the electrochemical exfoliation method [8,25,26,27]. The disadvantages of graphene obtained by liquid phase exfoliation of graphite employing ultrasonication include the graphene sheet size because it is typically smaller than 1 μm2, and the graphene yield is too low to be utilized in technological applications. Chemical exfoliation methods involve the oxidation of graphite to graphene oxide and then reduction chemically or thermally. These methods have attracted a lot of interest due to their potentially low cost and easy production, but to obtain the graphene structure, the thermal reduction of graphite oxide is carried out at high temperatures. However, the oxidized forms of graphite oxide cannot be completely removed with a chemical reducing agent, which may cause the degradation of the electronic properties [8].
Among these technologies, CVD appears to be the most promising process for industrial development and scale-up production of graphene. Nearly all of the transition metals included in the periodic table may be used as catalysts for the production of graphene, according to recent developments in the CVD process for growing graphene [28]. Most commonly, CVD of graphene involves injecting a precursor in the gas phase into a reaction chamber. In the reaction chamber at elevated temperature, the precursor reacts with the catalyst, and graphene is produced on the surface of the catalyst. The precursor may be a hydrocarbon (for example, methane or ethylene) or also low-molecular-weight alcohols. The temperatures of growth range from a few hundred degrees Celsius to the melting point of the catalyst metal [29]. Nickel (Ni) and copper (Cu) are two catalysts mostly used to prepare homogeneous and high-quality graphene. Nickel-catalyzed graphene, consisting of a large proportion of few-layer or multilayer, could therefore be widely used as stretchable transparent electrodes. In contrast, graphene grown on Cu is a monolayer with a large area that has great potential in high-mobility field-effect transistors. The catalyst surface is where graphene grains often form during the CVD growth of graphene on metals. During graphene growth, the grains coalesce and eventually lead to the formation of grain boundaries, which are mainly dispersed in the as-grown graphene layer. As a result, CVD-grown graphene is frequently polycrystalline and composed of a patchwork of grains with different sizes and orientations. Two methods can be used to obtain monocrystalline graphene without grain boundaries using chemical vapor deposition. The first method involves complete control of the number of nucleation centers, as reducing the nucleation number to one will finally result in the formation of individual graphene crystals. The second method, on the other hand, involves controlling the orientation of the graphene grains. In the absence of grain boundaries in the stitching areas, grains with aligned crystal lattices would merge neatly, resulting in large-sized graphene crystals [28].
The catalyst-assisted CVD approach is an intriguing synthetic method for producing wafer-sized graphene. When the nucleation and growth of crystallized graphene domains take place in an atmospheric pressure CVD process, the average size of graphene domains raises with increasing temperature and CH4 partial pressure, but the domain density decreases (but it is independent of the CH4 partial pressure). In addition, research by Liu L. et al. [30] indicates that the nucleation of graphene domains on Cu depends on the initial annealing temperature [30]. Plasma-enhanced chemical vapor deposition (PECVD) has great potential for graphene production due to its low-temperature growth and fast reaction rate. The high temperature provides the energy needed for graphene nucleation and growth in the thermal CVD process. However, in the PECVD process, the growth of multilayer graphene is dominated mostly by radicals generated in the plasma, so it does not need to use high temperatures. Notwithstanding, the disadvantages of this method include poor controllability and quality of obtained graphene after the graphene is grown. Due to these limitations, this method is not widely used [11,31]. Li N. et al. [11] studied the nucleation and growth of graphene at different temperatures by PECVD. They noted that graphene could not be grown at temperatures under 600 °C. In the temperature range of 650–800 °C, they obtained a high nucleation density of nanoscale graphene. In the temperature range from 850 °C to 900 °C, they formed bigger grains of graphene, and over 950 °C, the coexistence of large and small grains was produced [11].
Electrochemical exfoliation graphene is a popular top-down graphene recovery method. It is composed of oxidizing, expanding, and exfoliating graphite to produce graphene with a lower oxygen content. The process is schematically shown in Figure 2. The electrochemical exfoliation method is an environmentally friendly technique to fabricate graphene and enables production at room temperature [8,32,33]. This technique uses many types of graphite, including graphite foils, rods, plates, and powders, as electrodes in an aqueous or non-aqueous electrolyte, together with an electric current to cause electrode expansion [6]. The surface shows high roughness after electrolytic exfoliation, which is associated with a large surface area. The highest surface area was obtained after 1 h of exfoliation, and after that it declined rapidly [8]. Applying a voltage of 10 V resulted in the formation of multilayer graphene platelets with high defect concentration. Awasthi G. et al. [33] produced multi-layered (i.e., 1–5), high-quality graphene nanoplatelets using low voltage (3–4 V) during the process. For this purpose, they used potassium hydroxide solution as an electrolyte. The stirring of an electrolyte during electrolysis affected the properties of graphene, i.e., the average size of graphene nanoplatelets increased in that case, but also a reduction in the defect concentration was observed with an increase in several layers of C atoms in graphene [33]. To create high-yield solution-processable graphene, Zhang Y. et al. [34] used an electrochemical anode and cathode co-exfoliation technique. They obtained graphene with an ultralow defect and oxygen content on electrodes. The flexible supercapacitor exhibits great electrochemical performance and can be potentially used in wearable electronics [34].

3. Application of Graphene

Graphene has received a lot of attention in a variety of sectors, including energy storage, electrochemical sensors, nanoelectronics, optical biosensors, membranes, composites, drug delivery systems, etc. [5,9,10].

3.1. Biosensors

A biosensor is a device that detects chemical or biological reactions by producing signals proportional to the concentration of an analyte in the reaction. Applications for biosensors include the detection of pollutants, disease-causing microorganisms, and markers that are indicators of disease in bodily fluids like blood, saliva urine, or even sweat. Biosensors are also used in the discovery of new medicines and the monitoring of diseases. A typical biosensor consists of a bioreceptor, transducer, electronics, and display. A bioreceptor is a molecule that recognizes the analyte specifically [35]. A transducer turns biological interactions into physical signals (for example, optical, chemical, electrical, or thermal signals) [36]. Electronics is a part that processes the transduced signal and prepares it for display. On the display, the output signal is presented in numerical, graphical, tabular, or pictorial form, depending on the requirements of the user [35].
Due to its excellent charge transfer, high specific surface area (2620 m2/g), thermal conductivity of 500 W/mK, optical properties (visible and infrared light transmittance of 98%), and ability to immobilize molecules, graphene has become a widely used material in sensor manufacturing [7,9]. Graphene can be modified, for example, with gold or silver nanoparticles. It has been proven that the addition of nanogold resulted in greater surface area, enhanced electrochemical reactions, and interaction between elements or nanoparticles, as well as simplified charge transfer for better robustness, sensitivity, and specificity [9,37]. Narayanan J. et al. [38] successfully developed an electrochemical immunosensor for the detection of botulinum neurotoxin-E, where glassy carbon electrodes were modified using graphene nanosheets-aryldiazonium salt as a sensing platform and enzyme induced silver nanoparticles, which were deposited on gold nanoparticles as a signal amplifier [38].
Graphene-based electrochemical sensors are used for the detection of cancer biomarkers and neurotransmitters, but also as immunosensors and for the detection various other bioanalytes [39]. Graphene’s self-assembling characteristics are easily monitored, allowing it to be used in highly sensitive biosensors for DNA detection. Furthermore, due to their large surface area, superior electrical conductivity, and ability to load analytes, GO-based electrochemical biosensors have shown promising results in detecting cancers [18].
Ansari G. et al. [40] prepared a nanosensor using zinc oxide nanowires (ZnO NWs) and graphene. It can detect changes in hemoglobin content in vivo. ZnO NWs were produced using vapor-liquid–solid fabrication. The graphene layer can improve the sensitivity and response time of the biosensor, while the ZnO NWs can offer a significant surface area for immobilizing the biological molecules. The biosensor’s sensitivity is improved by graphene’s changed chemical potential. Furthermore, combining these two materials raises the selectivity of this device [40]. Soman G. et al. [41] designed a GO-based molecularly imprinted polymer for selective uric acid detection in blood serum. The main advantages of the manufactured sensor include the possibility of direct analysis of blood samples without requiring any pre-treatment, as well as extraordinary stability and reproducibility. The findings showed the potential of the developed sensor in uric acid analysis [41].
Additionally, Mubarakali A. et al. [42] worked on a sensor to detect glucose. They obtained that using colloidal Cu nanoparticles modified with graphene coated on indium-coated tin oxide glass substrate as a working electrode. The results indicate that the bioelectrode shows adequate stability and repeatability with a shorter response time. Because of those properties, the electrode may be one option for a non-enzymatic glucose biosensor [42]. Li B. et al. [43] also studied biosensors for glucose detection. A flexible glucose oxidase/chitosan/graphene sponge/Prussian blue biosensor for sweat glucose detection was designed (fabrication process is presented in Figure 3). The immobilization of glucose oxidase with chitosan offers glucose oxidase good isolation from the environment and protection, which increases glucose oxidase’s stability, reusability, and activity. The gaphene sponge (graphene aerogel) is characterized by a large specific surface area, high porosity, excellent electrical conductivity, and biocompatibility. As an electron transport medium, Prussian blue was electrochemically deposited on the working electrode. The results presented that the biosensor has a high selectivity for glucose and responds well to glucose in human sweat. In addition, the many binding sites of the graphene sponge allow for further composites with other functional materials to improve the flexible sweat glucose biosensors [43].
However, Pareek S. et al. [44] fabricated an electrochemical DNA biosensor for human papillomavirus-16 (HPV-16) detection. Cervical cancer is caused by the HPV-16 virus. GO and silver-coated gold nanoparticles were used to modify an indium tin oxide-coated glass electrode. The results indicate that the biosensor has a high sensitivity for the detection of HPV-16, and it is possible to detect it in the early stage, which can be crucial in developing point-of-care devices [44]. Bao J. et al. [45] developed an electrochemical biosensor for methylated DNA detection. DNA methylation is related to cell proliferation and differentiation. More and more often, there is information that abnormal methylation contributes to the occurrence of diseases, especially in tumorigenesis. It can also be used as biomarkers to predict response to chemotherapy strategies [45,46]. They proposed the biosensor, which included gold electrodes and nanocomposite based on gold nanoparticles, rGO, and graphite carbon nitride. The nanocomposite showed excellent electrochemical properties by the synergistic effects of each component. The proposed biosensor also displayed features of high specificity, stability, and reproducibility. Because of the synergistic effect of each component, the nanocomposite demonstrated good electrochemical characteristics [45].

3.2. Tissue Engineering

Tissue engineering is “an interdisciplinary field that applies the principles of engineering and the life sciences toward the development of biological substitutes that restore, maintain, or improve tissue function” [47]. Tissue engineering is often known as “regenerative medicine”. It offers a variety of solutions, from building new organs for transplant surgery to repairing damaged tissues. Tissue engineering, apart from cell culture and stem cell differentiation, also uses biodegradable and biologically safe materials to create structures called “scaffolds” [48]. Due to different design restrictions related to tissue engineering, features including mass transport, scaffold degradation, and biocompatibility must be taken into consideration while designing scaffolds [49]. For successful tissue engineering, it is essential to produce a biocompatible porous scaffold to support the cells, delivering bioactive molecules (for example, growth factors); as a result, they can produce physical and chemical cues that will ultimately determine the cell fate and cellular organization [50]. The cell microenvironment has an impact on the adhesion, migration, differentiation, communication, and proliferation of cells on the surface, biomaterials, or inside the extracellular matrix (ECM). A perfect biomaterial for tissue engineering and regenerative medicine often aims to imitate the corresponding ECM, providing cells with the right microenvironment [51].
The commonly used materials for tissue engineering are presented in Table 1. In addition, 2D nanomaterials have recently attracted great interest. 2D nanomaterials have a thickness of a few nanometers but a large lateral size. The unique energy level structure and optical properties, controllable thickness, simplicity in modification and doping, inherent bioactivity, high biocompatibility, and biodegradability are just a few of the distinctive properties of 2D nanomaterials. Additionally, 2D nanomaterials’ large specific surface area makes them good carriers of nanoparticles or medicines [52]. In addition to graphene, scientists in recent years have proposed the following: 2D oxide and hydroxide nanosheets [53], black phosphorus nanoparticles, nanodots, nanosheets [54], non-spherical metal nanomaterials (noble metal or transition metal dichalcogenides) [55], or hexagonal boron nitride (hBN) nanoparticle [56]. Tarhan T. et al. [56] used hexagonal boron nitride nanoparticles, silver nanoparticles, and polydopamine to stimulate wound healing. Silver nanoparticles were chosen because of their anti-inflammatory properties and hBN for induced proliferation and migration of cells. Even the low concentration of the obtained agent enhances wound healing besides it reduces ROS production, promotes wound closure, and reorganizes tube formation in cells [56]. Whereas, non-spherical metal nanoparticles are used in biomedicine, including regenerative medicine and cancer therapy. What is important, the architecture of metal-based nanomaterials influences their interaction with biological systems, i.e., particle geometry is an important factor and can affect their biodistribution, interactions with blood vessels, tumor penetration, transport across endothelial cells, etc. [56,57].
Sharifi S. et al. [59] proposed magnesium–zinc–graphene oxide nanocomposite scaffolds for bone tissue engineering. Scaffolds were fabricated by using powder metallurgy method. Because of their biodegradability, adequate mechanical qualities, and required biocompatibility, magnesium (Mg) scaffolds are an appropriate option for bone tissue regeneration in several studies. The addition of Zn to Mg increases its corrosion resistance and mechanical qualities. Furthermore, zinc is an element crucial for the human organism; therefore, the released products should not be toxic to the body. Among the known alloys, Mg-6Zn is interesting because of its corrosion resistance, mechanical properties, and cell compatibility. However, GO indicates a positive effect on the adhesion and differentiation of osteoblasts and stimulates biomineralization [59,60,61,62]. The addition of graphene oxide made the mechanical strength and corrosion resistance significantly improved compared to Mg-Zn scaffolds. Additionally, Mg-Zn-GO scaffolds were biocompatible and not cytotoxic in contact with L-929 cells [59].
Chitosan is an amino polysaccharide that is largely found in the cell walls of fungi, plants, insects, and marine invertebrates. Due to its chemical stability, high biocompatibility, antibacterial activity, biodegradability, ability to adsorb proteins, and to accelerate wound healing, CS is used in implantology, as a drug delivery system or scaffold [58,63,64]. Valencia A.M. [65] proposed to use chitosan functionalized GO by a covalent bond (CS-GO). The results show the potential application of the CS-GO compound in tissue engineering because of a low inflammatory response in vivo test and advanced resorption at 60 days of implantation [65]. Nakagawa de Arruda M. et al. [32] designed a few-layer graphene using the electrochemical exfoliation of graphite. It was supplemented with chitosan to form a homogeneous composite with low oxygen content. In addition to increasing surface area and enhancing chemical interactions, chitosan creates homogenous coatings with graphitic materials that improve the percolation of charges and electrical signals, increasing conductivity or capacitance. Because of its higher dispersion and resonance active groups, the graphene composite’s electrical conductivity increased. Furthermore, it is environmentally friendly because it uses less water and produces less toxic waste [32].
Hydroxyapatite (Ca10(PO4)6(OH)2) is a commonly used material because of its similarity with natural bone in terms of chemistry and structure. It has high biocompatibility, bioactivity, and osteoconductivity. HAp also shows osteoinductive properties and it promotes bone regeneration [66,67,68,69]. However, HAp’s limited mechanical strength and fracture toughness limit its clinical applicability because they could cause the propagation of cracks and an increase in corrosion rate [66,67,69,70]. To improve mechanical properties, some materials are used as reinforcements of Hap, i.e., polyethylene, Al2O3, TiO2, and carbon nanotubes. Unfortunately, these reinforcements may reduce the biological properties of hydroxyapatite and affect adjacent tissues [66]. However, HAp/GO composite showed improved in vitro osteoblast adhesion and apatite mineralization [71].
Liu Y. et al. [72] proposed HAp and HAp–graphene nanosheet (GN) composites synthesized using a liquid precipitation approach and deposited using vacuum cold spraying. The results demonstrated that the osteoblast cells spread and proliferated more readily on the GN-containing coatings [72]. Ramadas M. et al. [71] proposed hydroxyapatite nanorods grown on a graphene oxide sheet using a hydrothermal process. To assess cytotoxicity at various concentrations of material, they used human skin cancer cells. The nanocomposites presented no cytotoxicity effects on cancer cell lines. Although HAp/GO provides excellent biocompatibility [71]. However, Sánchez-Campos D. et al. [73] prepared a composite containing HAp, GO, and silver nanoparticles using a microwave-assisted hydrothermal method, a modified Hummer’s synthesis, and by using dietary quercetin for silver nanoparticles formation and deposition. A dose-dependent toxic response was found in eukaryotic cells. It depended on the presence of GO, but superior cytocompatibility was observed for the samples with HAp. Furthermore, the composite increased the inhibition zone for S. aureus and E. coli compared to silver nanoparticles. The results of the antibacterial study are presented in Figure 4. These indicate that the material can be applied as a bactericidal agent for tissue replacement and surface coating [73].
Han W. et al. [66] prepared graphene oxide/hydroxyapatite composite coatings on a titanium substrate using the electrophoretic deposition method. GO/HAp composite coatings outperformed HAp coatings in terms of adhesion strength and corrosion resistance. An appropriate amount of graphene oxide promoted the proliferation of the mouse fibroblast cells and enhanced the mineralization properties. The results showed that 5 wt% GO exhibited optimal mechanical properties, bioactivity, corrosion resistance, and good biocompatibility [66]. Fardi S.R. et al. [67] deposited hydroxyapatite–graphene oxide nanocomposites on titanium using an ultrasound-assisted electrophoretic technique. Ti sheets were anodized and then coated with HAp-GO with various compositions (0, 1, and 3 wt% of GO). They improved the mechanical properties with the best reported adhesion strength demonstrated by coating with 1 wt% GO. This coating showed excellent bioactivity and corrosion resistance [67]. Daulbayev C. et al. [74] fabricated a graphene oxide/calcium hydroxyapatite/polycaprolactone composite using an electrospun technique. Graphene oxide/hydroxyapatite composite was dispersed in biodegradable PLC. The results showed good antimicrobial activity against Gram-positive and Gram-negative bacterial strains, which depended on the GO loading. Furthermore, the composite had no cytotoxic effect on the preosteoblast MC3T3-E1 cell line. This indicates that this composite scaffold can be used for potential bone tissue regeneration [74]. Zhao H. et al. [75] fabricated porous graphene oxide/hydroxyapatite composite ceramic scaffolds using digital light processing technology. They demonstrated that adding a small amount of GO (0.1–0.4 wt%) to composite ceramics can improve their mechanical properties. The findings revealed that no scaffolds are cytotoxic. The composite scaffold increased cell adhesion, proliferation, and the expression of osteogenesis-related genes, while GO (0.1–0.2 wt%)/HAp scaffolds demonstrated superior alkaline phosphatase activity and more effective bone mineralization, as well as osteoinductivity [75].

3.3. Drug Delivery

Most conventional therapies and drug delivery systems have disadvantages such as rapid metabolism and excretion of drugs before reaching the target, poor water solubility, non-specificity to the target site, and opposing impacts on normal tissues. Therefore, nanotechnology and nanomaterials are becoming more and more popular in the development of innovative drug transport mechanisms. The efficiency of intelligent drug delivery systems is continually being improved. This is to maximize therapeutic activity and reduce side effects [76,77,78].
Li W. et al. [79] constructed a rectal delivery system using temperature-sensitive hydroxybutyl chitosan gel cross-linked with GO with pingyangmycin as a model drug. The temperature-sensitive gels are mostly based on polyisopropyl acrylamide, poloxamer, and CS. Because of their good biocompatibility, high permeability, and low drug toxicity, they are frequently used for injectable drug administration, oral drug delivery, rectal drug delivery, or mucosal drug delivery. The obtained material presented an effectively prolonged action time of the drug in vivo with rectal administration, and a sustained release effect [79]. Whereas, pH-responsive chitosan-based hydrogel can also be used as a drug delivery system. Hydrogel was efficiently fabricated by varying amounts of GO for controlled cephradine release. The content of GO proportionally affected the thermal stability. The material was most effective against Escherichia coli and Pseudomonas aeruginosa. Furthermore, pH-responsive swelling (schematic diagram presented in Figure 5) and biodegradability were confirmed in phosphate-buffered saline and proteinase K solutions [80].
Breast cancer is one of the most common cancers in women and has a high mortality rate. Chemotherapy, immunotherapy, radiotherapy, and surgery are just a few of the methods that have been used to treat cancer, and combining these methods typically results in a more successful treatment [81,82,83]. However, scientists are still working on improving therapeutic methods. Matiyani M. et al. [84] developed a multi-functionalized GO-based novel drug nanocarrier for the delivery of the chemotherapeutic. They used quercetin and curcumin, which are potent anticancer drugs. The drug delivery system was created by combining graphene oxide with a hydrophilic polymer and metal oxides, zinc oxide and titanium dioxide. The results showed that the nanocarrier exhibits a pH-sensitive release of loaded drugs, besides demonstrating cytotoxicity towards breast cancer cells [84]. Whereas, Rajaei M. et al. [81] synthesized pH-sensitive hydrogel of chitosan/agarose/graphene oxide with glyoxal as the cross-linker. It was prepared using the water-in-oil-in-water emulsification technique. The results indicate that the loading and entrapment efficiencies of the drug are satisfactory. A highly effective and sustained medicine release profile was observed at pH 5.4; within 48 hours, almost the entire content of the drug model was released. In addition, effective cytotoxicity was observed against breast cancer cell lines (MCF-7) [81].
Quantum dots can be used to minimize the size of sheets to less than 100 nm. Moreover, quantum dots show better properties with more active groups on their surface than typical graphene and graphene oxide. They exhibit unique properties, i.e., water solubility, non-toxicity, high biocompatibility, a small size with a large surface area, high drug loading capacity, and better cellular uptake [85,86,87]. Mohammed-Ahmed H.K. et al. [85] examined graphene oxide quantum dots conjugated with glucosamine and boric acid with and without doxorubicin (DOX) for cancer therapy. The results showed that nanocomposite with boric acid improved the loading and release of DOX, with higher cellular internalization. Seyyedi Zadeh et al. [86] also worked on the production of a drug carrier based on quantum dots. They synthesized a magnetic graphene quantum dots-Fe3O4 nanocarrier targeted with folic acid (FA) and loaded with curcumin. pH-dependent drug release was observed for both GQDs-Fe3O4 and GQDs-Fe3O4-FA. The bare nanocarrier had no cytotoxic effect on normal cells. In addition, folic acid acted as a selective targeting agent in the nanocarriers [86].

4. Conclusions

Single-walled carbon nanotubes, multiwalled carbon nanotubes, graphene, and reduced graphene oxide are just a few of the carbon nanostructures that have drawn a lot of research interest. Graphene is a 2D sheet of sp2-hybridized carbon atoms of atomic thickness that are organized in a honeycomb crystal lattice. Graphene is attracting a lot of interest due to its unique structural, optical, chemical, thermal, mechanical, electrical, and biological properties. Its derivatives are graphene oxide and reduced graphene oxide and graphene quantum dots. Graphene can be created using a variety of processes, including pyrolysis, epitaxial growth, chemical and electrochemical exfoliation of graphite, physical vapor deposition, and chemical vapor deposition. But the most promising method for industrial development and the production of graphene among those approaches is CVD.
Biomedical applications such as biosensing, medication delivery, regenerative medicine, and diagnostic tools are rapidly expanding. Graphene-based scaffolds are a particularly promising technology and have attracted interest due to their strong clinical potential in tissue regeneration therapies due to their substantial influence on the proliferation and differentiation of used stem cells. However, scientists still have many problems to solve, the biggest challenges are as follows: cytotoxicity, biodistribution, biotransformation, and immune response.
In conclusion, graphene nanomaterials have a bright future in biomedical applications. However, scientists still have many problems to solve, the biggest challenges being cytotoxicity, biotransformation, and immune response.

Author Contributions

Conceptualization, K.M.; writing—original draft preparation, K.M.; writing and editing, B.Ś.-Ż. supervision, B.Ś.-Ż.—substantial assistance in writing the article, methodology. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

We thank Michael Tokarz for checking the correctness of the English language.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chakrabarty, N.; Chakraborty, A.K.; Kumar, H. Nickel-Hydroxide-Nanohexagon-Based High-Performance Electrodes for Supercapacitors: A Systematic Investigation on the Influence of Six Different Carbon Nanostructures. J. Phys. Chem. C 2019, 123, 29104–29115. [Google Scholar] [CrossRef]
  2. Meng, X.; Hou, L.; Jin, H.; Li, W.; Wang, S.; Wang, Z.; An, J.; Wen, C.; Ji, G.; Xu, X.; et al. Study on Corrosion Protection Properties of PANI/ZnO/Zn/Graphene Coating on Aluminum Alloy. Diam. Relat. Mater. 2023, 136, 110067. [Google Scholar] [CrossRef]
  3. Geim, A.K.; Novoselov, K.S. The Rise of Graphene. Nat. Mater. 2007, 6, 183–191. [Google Scholar] [CrossRef] [PubMed]
  4. Abergel, D.S.L.; Apalkov, V.; Berashevich, J.; Ziegler, K.; Chakraborty, T. Properties of Graphene: A Theoretical Perspective. Adv. Phys. 2010, 59, 261–482. [Google Scholar] [CrossRef]
  5. Mohan, V.B.; Lau, K.; Hui, D.; Bhattacharyya, D. Graphene-Based Materials and Their Composites: A Review on Production, Applications and Product Limitations. Compos. Part B Eng. 2018, 142, 200–220. [Google Scholar] [CrossRef]
  6. Mbayachi, V.B.; Ndayiragije, E.; Sammani, T.; Taj, S.; Mbuta, E.R.; khan, A.U. Graphene Synthesis, Characterization and Its Applications: A Review. Results Chem. 2021, 3, 100163. [Google Scholar] [CrossRef]
  7. Shahdeo, D.; Roberts, A.; Abbineni, N.; Gandhi, S. Chapter Eight—Graphene Based Sensors. In Comprehensive Analytical Chemistry; Hussain, C.M., Ed.; Elsevier: Amsterdam, The Netherlands, 2020; Volume 91, pp. 175–199. [Google Scholar]
  8. Bose, S.; Kuila, T.; Kim, N.H.; Lee, J.H. 4—Graphene Produced by Electrochemical Exfoliation. In Graphene: Properties, Preparation, Characterisation and Devices; Skákalová, V., Kaiser, A.B., Eds.; Woodhead Publishing: Sawston, UK, 2014; pp. 81–98. [Google Scholar]
  9. Lin, L.P.; Tan, M.T.T. Biosensors for the Detection of Lung Cancer Biomarkers: A Review on Biomarkers, Transducing Techniques and Recent Graphene-Based Implementations. Biosens. Bioelectron. 2023, 237, 115492. [Google Scholar] [CrossRef]
  10. Kumar, Y.; Sahoo, S.; Chakraborty, A.K. Mechanical Properties of Graphene, Defective Graphene, Multilayer Graphene and SiC-Graphene Composites: A Molecular Dynamics Study. Phys. B Condens. Matter 2021, 620, 413250. [Google Scholar] [CrossRef]
  11. Li, N.; Li, D.; Zhen, Z.; Zhang, R.; Mu, R.; Xu, Z.; He, L. Nucleation and Growth of Graphene at Different Temperatures by Plasma Enhanced Chemical Vapor Deposition. Mater. Today Commun. 2023, 36, 106568. [Google Scholar] [CrossRef]
  12. Kenry; Lee, W.C.; Loh, K.P.; Lim, C.T. When Stem Cells Meet Graphene: Opportunities and Challenges in Regenerative Medicine. Biomaterials 2018, 155, 236–250. [Google Scholar] [CrossRef]
  13. Sanchez, V.C.; Jachak, A.; Hurt, R.H.; Kane, A.B. Biological Interactions of Graphene-Family Nanomaterials: An Interdisciplinary Review. Chem. Res. Toxicol. 2012, 25, 15–34. [Google Scholar] [CrossRef]
  14. Wang, Y.; Li, Z.; Wang, J.; Li, J.; Lin, Y. Graphene and Graphene Oxide: Biofunctionalization and Applications in Biotechnology. Trends Biotechnol. 2011, 29, 205–212. [Google Scholar] [CrossRef] [PubMed]
  15. Balandin, A.A.; Ghosh, S.; Bao, W.; Calizo, I.; Teweldebrhan, D.; Miao, F.; Lau, C.N. Superior Thermal Conductivity of Single-Layer Graphene. Nano Lett. 2008, 8, 902–907. [Google Scholar] [CrossRef]
  16. Tale, B.; Nemade, K.R.; Tekade, P. V Graphene Based Nano-Composites for Efficient Energy Conversion and Storage in Solar Cells and Supercapacitors: A Review. Polym. Technol. Mater. 2021, 60, 784–797. [Google Scholar] [CrossRef]
  17. Raslan, A.; Saenz del Burgo, L.; Ciriza, J.; Pedraz, J.L. Graphene Oxide and Reduced Graphene Oxide-Based Scaffolds in Regenerative Medicine. Int. J. Pharm. 2020, 580, 119226. [Google Scholar] [CrossRef]
  18. Arshad, F.; Nabi, F.; Iqbal, S.; Khan, R.H. Applications of Graphene-Based Electrochemical and Optical Biosensors in Early Detection of Cancer Biomarkers. Colloids Surf. B Biointerfaces 2022, 212, 112356. [Google Scholar] [CrossRef]
  19. Ndagijimana, P.; Rong, H.; Ndokoye, P.; Mwizerwa, J.P.; Nkinahamira, F.; Luo, S.; Guo, D.; Cui, B. A Review on Activated Carbon/Graphene Composite-Based Materials: Synthesis and Applications. J. Clean. Prod. 2023, 417, 138006. [Google Scholar] [CrossRef]
  20. 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, 2717–2720. [Google Scholar] [CrossRef]
  21. Hulagabali, M.M.; Vesmawala, G.R.; Patil, Y.D. Synthesis, Characterization, and Application of Graphene Oxide and Reduced Graphene Oxide and Its Influence on Rheology, Microstructure, and Mechanical Strength of Cement Paste. J. Build. Eng. 2023, 71, 106586. [Google Scholar] [CrossRef]
  22. Dorri Moghadam, A.; Omrani, E.; Menezes, P.L.; Rohatgi, P.K. Mechanical and Tribological Properties of Self-Lubricating Metal Matrix Nanocomposites Reinforced by Carbon Nanotubes (CNTs) and Graphene—A Review. Compos. Part B Eng. 2015, 77, 402–420. [Google Scholar] [CrossRef]
  23. Kuilla, T.; Bhadra, S.; Yao, D.; Kim, N.H.; Bose, S.; Lee, J.H. Recent Advances in Graphene Based Polymer Composites. Prog. Polym. Sci. 2010, 35, 1350–1375. [Google Scholar] [CrossRef]
  24. Schniepp, H.C.; Li, J.-L.; McAllister, M.J.; Sai, H.; Herrera-Alonso, M.; Adamson, D.H.; Prud’homme, R.K.; Car, R.; Saville, D.A.; Aksay, I.A. Functionalized Single Graphene Sheets Derived from Splitting Graphite Oxide. J. Phys. Chem. B 2006, 110, 8535–8539. [Google Scholar] [CrossRef] [PubMed]
  25. Hernandez, Y.; Nicolosi, V.; Lotya, M.; Blighe, F.M.; Sun, Z.; De, S.; McGovern, I.T.; Holland, B.; Byrne, M.; Gun’Ko, Y.K.; et al. High-Yield Production of Graphene by Liquid-Phase Exfoliation of Graphite. Nat. Nanotechnol. 2008, 3, 563–568. [Google Scholar] [CrossRef] [PubMed]
  26. Stankovich, S.; Dikin, D.A.; Piner, R.D.; Kohlhaas, K.A.; Kleinhammes, A.; Jia, Y.; Wu, Y.; Nguyen, S.T.; Ruoff, R.S. Synthesis of Graphene-Based Nanosheets via Chemical Reduction of Exfoliated Graphite Oxide. Carbon N. Y. 2007, 45, 1558–1565. [Google Scholar] [CrossRef]
  27. Liu, F.; Wang, C.; Sui, X.; Riaz, M.A.; Xu, M.; Wei, L.; Chen, Y. Synthesis of Graphene Materials by Electrochemical Exfoliation: Recent Progress and Future Potential. Carbon Energy 2019, 1, 173–199. [Google Scholar] [CrossRef]
  28. Geng, D.; Wang, H.; Yu, G. Graphene Single Crystals: Size and Morphology Engineering. Adv. Mater. 2015, 27, 2821–2837. [Google Scholar] [CrossRef]
  29. Frank, O.; Kalbac, M. 2—Chemical Vapor Deposition (CVD) Growth of Graphene Films. In Graphene: Properties, Preparation, Characterisation and Devices; Skákalová, V., Kaiser, A.B., Eds.; Woodhead Publishing: Sawston, UK, 2014; pp. 27–49. [Google Scholar]
  30. Liu, L.; Zhou, H.; Cheng, R.; Chen, Y.; Lin, Y.-C.; Qu, Y.; Bai, J.; Ivanov, I.A.; Liu, G.; Huang, Y.; et al. A Systematic Study of Atmospheric Pressure Chemical Vapor Deposition Growth of Large-Area Monolayer Graphene. J. Mater. Chem. 2012, 22, 1498–1503. [Google Scholar] [CrossRef]
  31. Terasawa, T.; Saiki, K. Growth of Graphene on Cu by Plasma Enhanced Chemical Vapor Deposition. Carbon N. Y. 2012, 50, 869–874. [Google Scholar] [CrossRef]
  32. de Arruda, M.N.; Palley, B.F.; de Souza, G.F.; Lopez, L.; Pontes, M.A.P.; Fernandes, W.P.; Gandara, M.; Machado, F.B.C.; Ferrão, L.F.D.A.; Gonçalves, E.S. Physicochemical Characterization of a Low-Oxygen Composite Based on Chitosan and Electrochemically Exfoliated Graphene. Mater. Chem. Phys. 2023, 306, 128049. [Google Scholar] [CrossRef]
  33. Awasthi, G.; Mistry, K.; Jamnapara, N.; Salot, M.; Santhy, K.; Mandal, D.; Chaudhury, S.K. Effect of Stirring on Characteristics of Electrochemically Exfoliated Graphene. Materialia 2023, 30, 101818. [Google Scholar] [CrossRef]
  34. Zhang, Y.; Xu, Y.; Liu, R.; Niu, Y. Synthesis of High-Quality Graphene by Electrochemical Anodic and Cathodic Co-Exfoliation Method. Chem. Eng. J. 2023, 461, 141985. [Google Scholar] [CrossRef]
  35. Bhalla, N.; Jolly, P.; Formisano, N.; Estrela, P. Introduction to Biosensors. Essays Biochem. 2016, 60, 1–8. [Google Scholar] [CrossRef] [PubMed]
  36. Prasad, S.V.S.; Kumar, M.; Arulananth, T.S.; Ravi, B.; Kumar, B.; Kiran Kumar, B. Graphene/ZnO Nanocomposite Based Optical Biosensors. Mater. Today Proc. 2023. [Google Scholar] [CrossRef]
  37. O’Brien, C.; Varty, K.; Ignaszak, A. The Electrochemical Detection of Bioterrorism Agents: A Review of the Detection, Diagnostics, and Implementation of Sensors in Biosafety Programs for Class A Bioweapons. Microsyst. Nanoeng. 2021, 7, 16. [Google Scholar] [CrossRef]
  38. Narayanan, J.; Sharma, M.K.; Ponmariappan, S.; Sarita; Shaik, M.; Upadhyay, S. Electrochemical Immunosensor for Botulinum Neurotoxin Type-E Using Covalently Ordered Graphene Nanosheets Modified Electrodes and Gold Nanoparticles-Enzyme Conjugate. Biosens. Bioelectron. 2015, 69, 249–256. [Google Scholar] [CrossRef]
  39. Ashraf, G.; Aziz, A.; Iftikhar, T.; Zhong, Z.-T.; Asif, M.; Chen, W. The Roadmap of Graphene-Based Sensors: Electrochemical Methods for Bioanalytical Applications. Biosensors 2022, 12, 1183. [Google Scholar] [CrossRef] [PubMed]
  40. Ansari, G.; Pal, A.; Srivastava, A.K.; Verma, G. Detection of Hemoglobin Concentration in Human Blood Samples Using a Zinc Oxide Nanowire and Graphene Layer Heterostructure Based Refractive Index Biosensor. Opt. Laser Technol. 2023, 164, 109495. [Google Scholar] [CrossRef]
  41. Soman, G.; Molahalli, V.; Hegde, G. Molecularly Imprinted Graphene Based Biosensor as Effective Tool for Electrochemical Sensing of Uric Acid. Sens. Int. 2023, 4, 100243. [Google Scholar] [CrossRef]
  42. Mubarakali, A.; Gopinath, S.; Parthasarathy, P.; Arun Kumar, U.; Alavudeen Basha, A. Highly Efficient and Sensitive Non-Enzymatic Glucose Biosensor Based on Flower-Shaped CuO-Colloid Nanoparticles Decorated with Graphene-Modified Nanocomposite Electrode. Measurement 2023, 217, 113145. [Google Scholar] [CrossRef]
  43. Li, B.; Wu, X.; Shi, C.; Dai, Y.; Zhang, J.; Liu, W.; Wu, C.; Zhang, Y.; Huang, X.; Zeng, W. Flexible Enzymatic Biosensor Based on Graphene Sponge for Glucose Detection in Human Sweat. Surf. Interfaces 2023, 36, 102525. [Google Scholar] [CrossRef]
  44. Pareek, S.; Jain, U.; Bharadwaj, M.; Saxena, K.; Roy, S.; Chauhan, N. An Ultrasensitive Electrochemical DNA Biosensor for Monitoring Human Papillomavirus-16 (HPV-16) Using Graphene Oxide/Ag/Au Nano-Biohybrids. Anal. Biochem. 2023, 663, 115015. [Google Scholar] [CrossRef]
  45. Bao, J.; Ding, K.; Zhu, Y. An Electrochemical Biosensor for Detecting DNA Methylation Based on AuNPs/RGO/g-C3N4 Nanocomposite. Anal. Biochem. 2023, 673, 115180. [Google Scholar] [CrossRef]
  46. Heyn, H.; Esteller, M. DNA Methylation Profiling in the Clinic: Applications and Challenges. Nat. Rev. Genet. 2012, 13, 679–692. [Google Scholar] [CrossRef] [PubMed]
  47. Langer, R.; Vacanti, J. Advances in Tissue Engineering. J. Pediatr. Surg. 2016, 51, 8–12. [Google Scholar] [CrossRef] [PubMed]
  48. Elitok, M.S.; Gunduz, E.; Gurses, H.E.; Gunduz, M. Chapter 20—Tissue Engineering: Towards Development of Regenerative and Transplant Medicine. In Omics Technologies and Bio-Engineering: Towards Improving Quality of Life; Barh, D., Azevedo, V., Eds.; Academic Press: Cambridge, MA, USA, 2018; pp. 471–495. [Google Scholar]
  49. Webber, M.J.; Khan, O.F.; Sydlik, S.A.; Tang, B.C.; Langer, R. A Perspective on the Clinical Translation of Scaffolds for Tissue Engineering. Ann. Biomed. Eng. 2015, 43, 641–656. [Google Scholar] [CrossRef]
  50. Ghosal, K.; Mondal, P.; Bera, S.; Ghosh, S. Graphene Family Nanomaterials- Opportunities and Challenges in Tissue Engineering Applications. FlatChem 2021, 30, 100315. [Google Scholar] [CrossRef]
  51. Gao, J.; Yu, X.; Wang, X.; He, Y.; Ding, J. Biomaterial–Related Cell Microenvironment in Tissue Engineering and Regenerative Medicine. Engineering 2022, 13, 31–45. [Google Scholar] [CrossRef]
  52. Cao, Z.; Bian, Y.; Hu, T.; Yang, Y.; Cui, Z.; Wang, T.; Yang, S.; Weng, X.; Liang, R.; Tan, C. Recent Advances in Two-Dimensional Nanomaterials for Bone Tissue Engineering. J. Mater. 2023, 9, 930–958. [Google Scholar] [CrossRef]
  53. Ma, R.; Sasaki, T. Two-Dimensional Oxide and Hydroxide Nanosheets: Controllable High-Quality Exfoliation, Molecular Assembly, and Exploration of Functionality. Acc. Chem. Res. 2015, 48, 136–143. [Google Scholar] [CrossRef]
  54. Deng, L.; Xu, Y.; Sun, C.; Yun, B.; Sun, Q.; Zhao, C.; Li, Z. Functionalization of Small Black Phosphorus Nanoparticles for Targeted Imaging and Photothermal Therapy of Cancer. Sci. Bull. 2018, 63, 917–924. [Google Scholar] [CrossRef]
  55. Makvandi, P.; Zarepour, A.; Zheng, X.; Agarwal, T.; Ghomi, M.; Sartorius, R.; Zare, E.N.; Zarrabi, A.; Wu, A.; Maiti, T.K.; et al. Non-Spherical Nanostructures in Nanomedicine: From Noble Metal Nanorods to Transition Metal Dichalcogenide Nanosheets. Appl. Mater. Today 2021, 24, 101107. [Google Scholar] [CrossRef]
  56. Tarhan, T.; Şen, Ö.; Ciofani, M.E.; Yılmaz, D.; Çulha, M. Synthesis and Characterization of Silver Nanoparticles Decorated Polydopamine Coated Hexagonal Boron Nitride and Its Effect on Wound Healing. J. Trace Elem. Med. Biol. 2021, 67, 126774. [Google Scholar] [CrossRef]
  57. Zhu, X.; Vo, C.; Taylor, M.; Smith, B.R. Non-Spherical Micro- and Nanoparticles in Nanomedicine. Mater. Horiz. 2019, 6, 1094–1121. [Google Scholar] [CrossRef]
  58. Ma, P.; Li, M.; Hu, J.; Li, D.; Tao, Y.; Xu, L.; Zhao, H.; Da, J.; Li, L.; Zhao, G.; et al. New Skin Tissue Engineering Scaffold with Sulfated Silk Fibroin/Chitosan/Hydroxyapatite and Its Application. Biochem. Biophys. Res. Commun. 2023, 640, 117–124. [Google Scholar] [CrossRef]
  59. Sharifi, S.; Ebrahimian-Hosseinabadi, M.; Dini, G.; Toghyani, S. Magnesium-Zinc-Graphene Oxide Nanocomposite Scaffolds for Bone Tissue Engineering. Arab. J. Chem. 2023, 16, 104715. [Google Scholar] [CrossRef]
  60. Wen, C.; Zhan, X.; Huang, X.; Xu, F.; Luo, L.; Xia, C. Characterization and Corrosion Properties of Hydroxyapatite/Graphene Oxide Bio-Composite Coating on Magnesium Alloy by One-Step Micro-Arc Oxidation Method. Surf. Coat. Technol. 2017, 317, 125–133. [Google Scholar] [CrossRef]
  61. Toghyani, S.; Khodaei, M. Fabrication and Characterization of Magnesium Scaffold Using Different Processing Parameters. Mater. Res. Express 2018, 5, 35407. [Google Scholar] [CrossRef]
  62. Seyedraoufi, Z.S.; Mirdamadi, S. In Vitro Biodegradability and Biocompatibility of Porous Mg-Zn Scaffolds Coated with Nano Hydroxyapatite via Pulse Electrodeposition. Trans. Nonferrous Met. Soc. China 2015, 25, 4018–4027. [Google Scholar] [CrossRef]
  63. Jugowiec, D.; Łukaszczyk, A.; Cieniek, Ł.; Kot, M.; Reczyńska, K.; Cholewa-Kowalska, K.; Pamuła, E.; Moskalewicz, T. Electrophoretic Deposition and Characterization of Composite Chitosan-Based Coatings Incorporating Bioglass and Sol-Gel Glass Particles on the Ti-13Nb-13Zr Alloy. Surf. Coat. Technol. 2017, 319, 33–46. [Google Scholar] [CrossRef]
  64. Kumar, S.; Deepak, V.; Kumari, M.; Dutta, P.K. Antibacterial Activity of Diisocyanate-Modified Chitosan for Biomedical Applications. Int. J. Biol. Macromol. 2016, 84, 349–353. [Google Scholar] [CrossRef]
  65. Valencia, A.M.; Valencia, C.H.; Zuluaga, F.; Grande-Tovar, C.D. Synthesis and Fabrication of Films Including Graphene Oxide Functionalized with Chitosan for Regenerative Medicine Applications. Heliyon 2021, 7, e07058. [Google Scholar] [CrossRef] [PubMed]
  66. Han, W.; Liu, L.; Yu, B.; Zhu, Q. Electrodeposition of Graphene Oxide-Hydroxyapatite Composite Coating on Titanium Substrate. Ceram. Int. 2023, 49, 9647–9656. [Google Scholar] [CrossRef]
  67. Fardi, S.R.; khorsand, H.; Askarnia, R.; Pardehkhorram, R.; Adabifiroozjaei, E. Improvement of Biomedical Functionality of Titanium by Ultrasound-Assisted Electrophoretic Deposition of Hydroxyapatite-Graphene Oxide Nanocomposites. Ceram. Int. 2020, 46, 18297–18307. [Google Scholar] [CrossRef]
  68. Erakovic, S.; Jankovic, A.; Tsui, G.C.P.; Tang, C.-Y.; Miskovic-Stankovic, V.; Stevanovic, T. Novel Bioactive Antimicrobial Lignin Containing Coatings on Titanium Obtained by Electrophoretic Deposition. Int. J. Mol. Sci. 2014, 15, 12294–12322. [Google Scholar] [CrossRef] [PubMed]
  69. Bensalem, A.; Kucukosman, O.K.; Raszkiewicz, J.; Topkaya, F. Synthesis, Characterization, Bactericidal Activity, and Mechanical Properties of Hydroxyapatite Nano Powders Impregnated with Silver and Zinc Oxide Nanoparticles (Ag-ZnO-Hap). Ceram. Int. 2021, 47, 21319–21324. [Google Scholar] [CrossRef]
  70. Akao, M.; Aoki, H.; Kato, K. Mechanical Properties of Sintered Hydroxyapatite for Prosthetic Applications. J. Mater. Sci. 1981, 16, 809–812. [Google Scholar] [CrossRef]
  71. Ramadas, M.; Bharath, G.; Ponpandian, N.; Ballamurugan, A.M. Investigation on Biophysical Properties of Hydroxyapatite/Graphene Oxide (HAp/GO) Based Binary Nanocomposite for Biomedical Applications. Mater. Chem. Phys. 2017, 199, 179–184. [Google Scholar] [CrossRef]
  72. Liu, Y.; Dang, Z.; Wang, Y.; Huang, J.; Li, H. Hydroxyapatite/Graphene-Nanosheet Composite Coatings Deposited by Vacuum Cold Spraying for Biomedical Applications: Inherited Nanostructures and Enhanced Properties. Carbon N. Y. 2014, 67, 250–259. [Google Scholar] [CrossRef]
  73. Sánchez-Campos, D.; Peña-Juárez, M.G.; Salado-Leza, D.; Mendoza-Anaya, D.; Rodríguez-Lugo, V.; González-Calderón, J.A.; Pérez, E. A Non-Toxic Synergistic Antibacterial Platform Based on Green Silver Nanoparticles Deposited on Hydroxyapatite/Graphene Oxide Composites. Mater. Today Commun. 2023, 35, 106097. [Google Scholar] [CrossRef]
  74. Daulbayev, C.; Sultanov, F.; Korobeinyk, A.V.; Yeleuov, M.; Taurbekov, A.; Bakbolat, B.; Umirzakov, A.; Baimenov, A.; Daulbayev, O. Effect of Graphene Oxide/Hydroxyapatite Nanocomposite on Osteogenic Differentiation and Antimicrobial Activity. Surf. Interfaces 2022, 28, 101683. [Google Scholar] [CrossRef]
  75. Zhao, H.; Xing, H.; Lai, Q.; Zhao, Y.; Chen, Q.; Zou, B. Additive Manufacturing of Graphene Oxide/Hydroxyapatite Bioceramic Scaffolds with Reinforced Osteoinductivity Based on Digital Light Processing Technology. Mater. Des. 2022, 223, 111231. [Google Scholar] [CrossRef]
  76. Saharan, R.; Paliwal, S.K.; Tiwari, A.; Tiwari, V.; Singh, R.; Beniwal, S.K.; Dahiya, P.; Sagadevan, S. Exploring Graphene and Its Potential in Delivery of Drugs and Biomolecules. J. Drug Deliv. Sci. Technol. 2023, 84, 104446. [Google Scholar] [CrossRef]
  77. Safari, J.; Zarnegar, Z. Advanced Drug Delivery Systems: Nanotechnology of Health Design A Review. J. Saudi Chem. Soc. 2014, 18, 85–99. [Google Scholar] [CrossRef]
  78. Goodarzi, S.; Da Ros, T.; Conde, J.; Sefat, F.; Mozafari, M. Fullerene: Biomedical Engineers Get to Revisit an Old Friend. Mater. Today 2017, 20, 460–480. [Google Scholar] [CrossRef]
  79. Li, X.; Chen, A.; Liu, Y.; Li, L. Preparation and Rectal Administration of Hydroxybutyl Chitosan/Graphene Oxide Composite Thermosensitive Hydrogel. React. Funct. Polym. 2023, 189, 105608. [Google Scholar] [CrossRef]
  80. Qureshi, M.A.U.R.; Arshad, N.; Rasool, A. Graphene Oxide Reinforced Biopolymeric (Chitosan) Hydrogels for Controlled Cephradine Release. Int. J. Biol. Macromol. 2023, 242, 124948. [Google Scholar] [CrossRef] [PubMed]
  81. Rajaei, M.; Rashedi, H.; Yazdian, F.; Navaei-Nigjeh, M.; Rahdar, A.; Díez-Pascual, A.M. Chitosan/Agarose/Graphene Oxide Nanohydrogel as Drug Delivery System of 5-Fluorouracil in Breast Cancer Therapy. J. Drug Deliv. Sci. Technol. 2023, 82, 104307. [Google Scholar] [CrossRef]
  82. Dolatkhah, M.; Hashemzadeh, N.; Barar, J.; Adibkia, K.; Aghanejad, A.; Barzegar-Jalali, M.; Omidi, Y. Graphene-Based Multifunctional Nanosystems for Simultaneous Detection and Treatment of Breast Cancer. Colloids Surf. B Biointerfaces 2020, 193, 111104. [Google Scholar] [CrossRef]
  83. Siegel, R.L.; Miller, K.D.; Jemal, A. Cancer Statistics, 2017. CA. Cancer J. Clin. 2017, 67, 7–30. [Google Scholar] [CrossRef]
  84. Matiyani, M.; Rana, A.; Karki, N.; Garwal, K.; Pal, M.; Sahoo, N.G. Development of Multi-Functionalized Graphene Oxide Based Nanocarrier for the Delivery of Poorly Water Soluble Anticancer Drugs. J. Drug Deliv. Sci. Technol. 2023, 83, 104412. [Google Scholar] [CrossRef]
  85. 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]
  86. Seyyedi Zadeh, E.; Ghanbari, N.; Salehi, Z.; Derakhti, S.; Amoabediny, G.; Akbari, M.; Asadi Tokmedash, M. Smart PH-Responsive Magnetic Graphene Quantum Dots Nanocarriers for Anticancer Drug Delivery of Curcumin. Mater. Chem. Phys. 2023, 297, 127336. [Google Scholar] [CrossRef]
  87. 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] [PubMed]
Figure 1. Molecular structure of (a) graphite, (b) graphene, (c) graphene oxide, (d) reduced graphene oxide (reprinted from [20], with permission from Elsevier, license number: 5595830853941).
Figure 1. Molecular structure of (a) graphite, (b) graphene, (c) graphene oxide, (d) reduced graphene oxide (reprinted from [20], with permission from Elsevier, license number: 5595830853941).
Crystals 13 01413 g001
Figure 2. Electrochemical exfoliation method (reprinted from [32], with permission from Elsevier, license number: 5595791108264).
Figure 2. Electrochemical exfoliation method (reprinted from [32], with permission from Elsevier, license number: 5595791108264).
Crystals 13 01413 g002
Figure 3. Fabrication process flexible glucose oxidase/chitosan/graphene sponge/Prussian blue biosensor for sweat glucose detection. (A) Graphene sponge (GS) mixed with chitosan (CTS) and glucose oxidase (GOx); (B) obtained GOx/CTS/GS composite; (C) Prussian blue (PB), composite and Nafion solution were added to the gold three-electrode; (D) flexible biosensor (reprinted from [43], with permission from Elsevier, license number: 5611890451239).
Figure 3. Fabrication process flexible glucose oxidase/chitosan/graphene sponge/Prussian blue biosensor for sweat glucose detection. (A) Graphene sponge (GS) mixed with chitosan (CTS) and glucose oxidase (GOx); (B) obtained GOx/CTS/GS composite; (C) Prussian blue (PB), composite and Nafion solution were added to the gold three-electrode; (D) flexible biosensor (reprinted from [43], with permission from Elsevier, license number: 5611890451239).
Crystals 13 01413 g003
Figure 4. Pictures present inhibition zones (mm) of materials against S. aureus and E. coli: (a) the pure HAp, GO and silver nanoparticles (AgNPs); (b) HAp/GO, HAp + Ag, GO + Ag and HAp/GO + Ag composites (reprinted from [73], with permission from Elsevier, license number: 5611920074326).
Figure 4. Pictures present inhibition zones (mm) of materials against S. aureus and E. coli: (a) the pure HAp, GO and silver nanoparticles (AgNPs); (b) HAp/GO, HAp + Ag, GO + Ag and HAp/GO + Ag composites (reprinted from [73], with permission from Elsevier, license number: 5611920074326).
Crystals 13 01413 g004
Figure 5. Schematic diagram for swelling of hydrogel due to charge zones at polymeric chains (reprinted from [80], with permission from Elsevier, license number: 5611930020590).
Figure 5. Schematic diagram for swelling of hydrogel due to charge zones at polymeric chains (reprinted from [80], with permission from Elsevier, license number: 5611930020590).
Crystals 13 01413 g005
Table 1. Examples of materials used for tissue engineering [49,51,58].
Table 1. Examples of materials used for tissue engineering [49,51,58].
Natural Materials Synthetic Polymers Inorganic Material
collagens;
hyaluronic acid;
chondroitin sulfate;
chitosan (CS);
silk fibroin
polyethylene glycol (PEG);
poly(lactic-co-glycolic acid) (PLGA);
polycaprolactone (PCL);
ultra-high molecular weight polyethylene (UHMWPE)
hydroxyapatite (HAp);
bioglass;
calcium phosphate cement (CPC)
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

Malisz, K.; Świeczko-Żurek, B. Graphene Production and Biomedical Applications: A Review. Crystals 2023, 13, 1413. https://doi.org/10.3390/cryst13101413

AMA Style

Malisz K, Świeczko-Żurek B. Graphene Production and Biomedical Applications: A Review. Crystals. 2023; 13(10):1413. https://doi.org/10.3390/cryst13101413

Chicago/Turabian Style

Malisz, Klaudia, and Beata Świeczko-Żurek. 2023. "Graphene Production and Biomedical Applications: A Review" Crystals 13, no. 10: 1413. https://doi.org/10.3390/cryst13101413

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