Next Article in Journal
Kenaf Fiber and Hemp Fiber Multi-Walled Carbon Nanotube Filler-Reinforced Epoxy-Based Hybrid Composites for Biomedical Applications: Morphological and Mechanical Characterization
Next Article in Special Issue
The Surface Behavior of ZnO Films Prepared at Room Temperature
Previous Article in Journal
Investigation of Solar Air Collectors with Carbon-Nanotube-Based Turbulators and Pin Fin Arrangements
Previous Article in Special Issue
Combination of Cross-Correlation-Based Analysis and Ultrasonic GW Tomography for Barely Visible Impact Damage Detection Preliminary Assessment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nanocomposite Nanofibers of Graphene—Fundamentals and Systematic Developments

1
NPU-NCP Joint International Research Center on Advanced Nanomaterials and Defects Engineering, Northwestern Polytechnical University, Xi’an 710072, China
2
UNESCO-UNISA Africa Chair in Nanosciences/Nanotechnology, iThemba LABS, Somerset West 7129, South Africa
3
NPU-NCP Joint International Research Center on Advanced Nanomaterials and Defects Engineering, National Centre for Physics, Islamabad 44000, Pakistan
4
School of Materials Science & Engineering, Northwestern Polytechnical University, Xi’an 710072, China
5
Department of Physics, College of Science, Imam Mohammad Ibn Saud Islamic University (IMSIU), Riyadh 13318, Saudi Arabia
*
Author to whom correspondence should be addressed.
J. Compos. Sci. 2023, 7(8), 323; https://doi.org/10.3390/jcs7080323
Submission received: 5 July 2023 / Revised: 20 July 2023 / Accepted: 26 July 2023 / Published: 7 August 2023
(This article belongs to the Special Issue Feature Papers in Journal of Composites Science in 2023)

Abstract

:
Research on polymer nanocomposite nanofibers has seen remarkable growth over the past several years. One of the main driving forces for this progress is the increasing applicability of polymer nanocomposite nanofibers for technological applications. This review basically aims to present the current state of manufacturing polymer/graphene nanofiber nanocomposites, using appropriate techniques. Consequently, various conducting and thermoplastic polymers have been processed with graphene nano-reinforcement to fabricate the nanocomposite nanofibers. Moreover, numerous methods have been adopted for the fabrication of polymer/graphene nanocomposites and nanofibers including interfacial polymerization, phase separation, freeze drying, template synthesis, drawing techniques, etc. For the formation of polymer/graphene nanocomposite nanofibers, electrospinning can be preferable due to various advantages such as the need for simple equipment, control over morphology, and superior properties of the obtained material. The techniques such as solution processing, melt spinning, and spin coating have also been used to manufacture nanofibers. Here, the choice of manufacturing techniques and parameters affects the final nanofiber morphology, texture, and properties. The manufactured nanocomposite nanofibers have been examined for exceptional structural, microstructure, thermal, and other physical properties. Moreover, the properties of polymer/graphene nanofiber rely on the graphene content, dispersion, and matrix–nanofiller interactions. The potential of polymer/graphene nanocomposite nanofibers has been investigated for radiation shielding, supercapacitors, membranes, and the biomedical field. Hence, this review explains the literature-driven significance of incorporating graphene in polymeric nanofibers. Conclusively, most of the studies focused on the electrospinning technique to design polymer/graphene nanofibers. Future research in this field may lead to advanced innovations in the design and technical applications of nanocomposite nanofibers. To the best of our knowledge, research reports are available on this topic; however, the stated literature is not in a compiled and updated form. Therefore, field researchers may encounter challenges in achieving future advancements in the area of graphene-based nanocomposite nanofibers without first consulting the recent literature, such as an assembled review, to gain necessary insights, etc. Consequently, this state-of-the-art review explores the manufacturing, properties, and potential of polymer/graphene nanocomposite nanofibers.

Graphical Abstract

1. Introduction

The nanofibers of polymer/nanocarbon nanocomposites have been effectively developed and reported in the literature [1]. Graphene has been considered an important nanocarbon nanomaterial [2,3]. The combination of polymer and graphene has improved the morphological, electrical, thermal, mechanical, and other essential physical properties of the nanocomposites [4]. Improvements in nanocomposite properties have been observed due to the synergistic effect between the polymers and nanofillers [5,6]. Consequently, the polymer/graphene nanocomposite nanofibers were developed. Fine polymer/graphene nanofibers have been manufactured using facile techniques and different polymeric matrices. Electrospinning has been identified as a prominent technique to develop polymer- and nanocomposite-based nanofibers [7,8]. Here, the electrospinning technique revealed various controllable processing parameters. The polymer/graphene nanocomposite nanofibers have been designed and fabricated for radiation shielding, energy, membranes, and biomedical applications [9,10].
This review article comprehends the fundamentals, characteristics, and applications of polymer/graphene nanocomposite nanofibers. The inclusion of graphene nanofiller in polymers and processing using facile manufacturing techniques have resulted in high-performance nanocomposite nanofibers. Enhancement in the physical features has been observed with the increasing nanofiller contents. Applications of the finely manufactured polymer/graphene nanocomposite nanofibers have been observed for supercapacitors, radiation shielding, membranes, and biomedical devices.
Thus, this review is groundbreaking and pioneering to elaborate on the scientific advancements in the field of graphene-based nanofibers. This comprehensive review article states the fundamentals and technical features of polymer/graphene nanocomposite nanofibers. The incorporation of graphene in polymeric matrices using appropriate fiber processing techniques has produced high-performance nanocomposite nanofibers. The characteristics and performance of the nanofibers rely on the compatibility between the matrix–nanofiller, interactions, and consistent nanoparticle dispersal. High-competence nanocomposite nanofibers have been functional in technical fields ranging from energy to biomedical. To the best of our knowledge, this overview is novel and ground-breaking to highlight the field of graphene-derived nanofibers. Research has been reported on polymer/graphene nanofibers; nonetheless, the literature needs to be presented in an updated and compiled form. Hence, these polymer/graphene nanocomposite nanofibers will be helpful for the related scientists for essential future upheavals in graphene-derived nanofiber technologies. This novel article on polymer/graphene nanofibers will definitely be helpful for essential future innovations in nanofiber manufacturing.

2. Polymer and Polymer Nanocomposite Nanofibers

The polymer nanofibers are simply nanostructures with a diameter of a few nanometers and a length of up to several millimeters [11]. Various natural and synthetic polymers have been used to develop the nanofibers. The polymer nanofibers have superior physical properties and technical applications [12]. The polymer nanofiber diameter and characteristics have been dependent on the manufacturing techniques used. Accordingly, polymer nanofibers of different shapes have been manufactured, such as hollow nanofibers, ribbon-like nanofibers, wrinkled nanofibers, smooth nanofibers, and solid filled nanofibers.
Among different types of fibers, solid nanofibers with smooth surfaces have been preferred for various technical uses. In some cases, hollow nanofibers have been preferred according to the desired end application. Synthetic polymers including conducting polymers and thermoplastic polymers have been widely used to form nanofibers [13]. The solid round nanofibers with homogeneous surfaces have a large surface-area-to-volume ratio, optimum porosity, flexibility, toughness, and appreciable mechanical strength for applications such as membrane materials, packaging, and other technical applications. On the other hand, the soft porous and hollow nanofibers of natural polymers have found applications in wound dressing and biomaterials for tissue regeneration. Depending upon the polymer type and nanofiller used, the nanofiber properties such as mechanical properties, tensile strength, toughness, flexibility, electrical properties, optical properties, and chemical properties can be varied.
The polymer nanofibers have been manufactured through a range of techniques including solution drawing, melt extrusion, template technique, template-free process, and spinning techniques [14]. The polymer fibers and nanofibers have been produced through the traditional wet, dry, melt, or gel spinning techniques. The diameter of as-spun nanofibers may vary from nano- to a few micrometers. The traditional dry spinning process involves the use of one or more solvents to dissolve the polymers. This method is termed dry spinning because the solvents are removed through recovery procedures. In the wet spinning process, the fibers are placed in a coagulation bath, and solvent removal or evaporation techniques are not demanded. Due to the involvement of solvent extraction techniques, the dry spinning method has been found costly. The gel spinning technique has also been applied to the polymer fibers. In this method, the polymer gel is used for fiber spinning, and as-sun fibers are usually dried. The effectiveness of these techniques has been analyzed on the basis of surface area, porosity, flexibility, strength, and chemical resistance properties. The fabricated polymer nanofibers have been applied to membrane-based filtration systems, coatings, energy devices, electronics, tissue engineering, and drug delivery [15]. The conducting polymer nanofibers having a diameter of 200 nm and length of ~30 μm have been fabricated by the spinning method [16]. In particular, the polyaniline nanofibers have been processed by the spinning technique [17]. These nanofibers have important applications in electronics, supercapacitors, and sensors. Among thermoplastic polymers, polystyrene, polyethylene, polyamide, poly(vinyl alcohol), poly(ethylene oxide), Nafion, poly(acrylic acid), etc., have been used for the nanofiber manufacturing [18]. The properties of nanofibers have been found to depend on the polymer type and manufacturing method used [19].
Similar to polymers, the polymer nanocomposites have been processed to form nanofibers. Some of the important manufacturing techniques include interfacial polymerization, phase separation, freeze-drying synthesis, template synthesis, drawing method, and spinneret-based tunable engineered parameter (STEP) method.
In the interfacial polymerization of polymer graphene nanocomposites, different monomers are usually dissolved in two different immiscible solvent phases and allowed to react at the interface [20]. Graphene can be introduced in one phase. The monomers are polymerized at the interface of the emulsion [21]. Using interfacial polymerization, the nanofibers can be fabricated due to the homogeneous nucleated growth process. Using this method, different types of polymers and nanocomposites can be produced [22]. In this technique, the concentrations of monomers in the immiscible solvent phases usually affect the molecular weight of the polymer formed and graphene dispersion.
In the phase separation method, the phases usually separate due to physical incompatibility [23]. The solvent phase is extracted in the solution, whereas the other phase remains intact. Another important factor of this method is the polymer dissolution in a solvent at room temperature or elevated temperature. Nanocomposite or nanofiber morphology is usually affected by a controlled gelation step. Here, the polymer concentration and gelation temperature also need to be controlled [24]. However, all polymers do not have a phase separation phenomenon to produce the nanofibers.
Freeze drying is an important method to form polymer/graphene nanocomposites and nanofibers [25]. This technique involves the solid–liquid phase separation and ice segregation-induced self-assembly processes. Initially, the solution is frozen at a low temperature (−70 to −80 °C), allowing ice-crystal growth and nucleation. Then, the drying process and removal of frozen samples through direct sublimation are carried out [26]. Rapid freezing and direct sublimation are important factors to avoid any chemical reactions or side products during the material formation.
The template method usually employs the chemical or electrochemical oxidative polymerization processes [27]. Moreover, the template or mold is used to obtain a desired nanocomposite or nanofiber material. In this technique, the nanofibers are formed by passing polymer solution through the nanopores of the template under water pressure on one side, which causes the extrusion of the polymer and the formation of nanofibers. The fibers solidify upon contact with the solidifying solution [28]. Thus, nanofibers of different diameters can be formed depending on the size of template nanopores. Thus, this technique can be advantageous to control the fiber diameter. However, this method produces nanofibers of only a few micrometers, and cannot produce long nanofibers.
Spinneret-based tunable engineered parameters (STEP) technique has also been used to form the nanocomposite and nanofibers [29]. This technique employs a micropipette spinneret and rotating substrate. The nanofibers of nano- to micrometers in diameter can be formed. Consequently, highly aligned and uniform nanofibers are developed on the substrate [30]. In the STEP methods, numerous factors (polymer type, molecular weight, solvent type, viscosity, etc.) govern the fiber diameter, length, porosity, and defects. Table 1 shows the essential properties and techniques applied for polymer and polymer nanocomposite nanofibers.
Among the non-spinning methods, plasma-induced synthesis has also been used to form nanofibers [49]. The process consists of significant steps such as atomic vapor deposition, expansion of plasma, solution condensation, in situ oxygen reactions, and nanofiber growth. This technique may produce nanofibers of around 15–25 nm in diameter [50]. However, this is a sophisticated technique to form nanofibers and therefore less preferred. As compared to plasma-induced synthesis, electrospinning has been more frequently used due to facile processing advantages.
Thus, for the nanocomposite nanofibers, various carbon and inorganic nanoparticles have been used [51]. Carbon nanotube-reinforced polymeric nanofibers have been developed [52]. Fajardo-Diaz et al. [31] reported the nanofibers of polyamide/carbon nanotube using a spinning process for reverse osmosis membranes. Amer Flayeh et al. [32] fabricated the polystyrene/carbon nanotube nanocomposite nanofibers using the spinning method. The nanofibers have been studied for fiber texture and morphology. Liao et al. [33] manufactured polyaniline/single-walled carbon nanotube nanofibers. The nanofibers revealed enhancement in the electrical conductivity from 10−4 S cm−1 to 102 S cm−1, with increasing nanofiller contents. At high nanofiller contents, the nanofibers revealed a non-homogeneous surface and beaded texture [53]. Inorganic nanoparticles have also been used as an effective reinforcement in polymeric nanofibers [54]. In poly(vinyl alcohol) and cellulose matrices, silica nanoparticles and clay nanoplatelets have been reinforced to form nanocomposite nanofibers [55]. Titania nanoparticles have also been filled in the polymer nanofibers using the sol–gel and spinning methods [56]. The template method has been used for manufacturing polyaniline/titania nanocomposite nanofibers [36]. Patil et al. [34] fabricated the polyaniline/zinc oxide nanofibers through spinning. These nanofibers have uniform surfaces and morphology. The nanofiber diameter was ~200–300 nm. Kayaci et al. [35] manufactured the nylon 6,6/zinc oxide nanofibers through the spinning and atomic layer deposition techniques. The core–shell nanofibers have uniform thickness. Thus, the diameter, properties, and uniformity of the polymer nanocomposite nanofibers rely on the choice of polymer, nanofiller selection, and the fabrication method used.
The choice of manufacturing technique and related parameters definitely affect the final nanofiber material properties:
i.
In the electrospinning technique, the feed rate can influence the polymer solution delivery speed and jet intensity [57]. Consequently, the feed rate has been found to affect the diameter and morphology of the nanofibers. Increasing the polymer solution feed rate may enhance the fiber diameter, whereas slow flow may form thin fibers.
ii.
Distance between the spinneret tip and collector has been found to affect the morphology and fiber diameter [58]. The optimum distance between the tip and collector provides sufficient time to dry the nanofibers and avoid bead formation. Furthermore, the increasing distance results in more round solid fibers.
iii.
The sufficiently high voltage between a needle and metal collector is important to overcome the surface tension holding a drop of liquid at the needle tip [59]. Consequently, a thin fluid jet is projected out. The solvent is easily evaporated during the trajectory between the needle and collector. Hence, fibers with homogeneous surfaces and small diameters have been developed. The low electric field cannot provide enough jet elongation to generate uniform fibers.
iv.
The pumping pressure also influences the flow of polymer solution during the electrospinning process [60]. Very low pumping pressure may increase the nanofiber diameter and bead formation.
The technologies for producing nanofibers have been in progress, since the 20th century [61]. The developments in the 21st century (since the 2000s) have made fiber processes more commercially feasible. Recent developments must focus on enhancing the current manufacturing technologies to produce nanofibers with high product uniformity and process speed. Consequently, a large number of contract manufacturers can trade successfully. For efficient trading and commercial viability of nanofibers for technical applications, fibers having fine diameters must be efficiently produced and employed.
Other spinning-based processing techniques have also been seen in the literature to form nanofibers. The solution blow spinning technique has been adopted to form nanofibers [62,63]. This technique compensates for the essentials of both the electrospinning and melt-blowing methods [64]. This technique may be helpful to form the non-woven micro- and nanofibers. The nanofibers of poly(methyl methacrylate), poly(lactic acid), and poly(vinyl alcohol) have been effectively formed using solution blowing [65]. The choice of polymer and concentration has been found to influence the fiber diameter [66]. The solution blow spinning setup has a syringe pump, concentric nozzle, high-velocity gas flow, and collector [67]. The nanofiber size and diameter have been found comparable to the electrospinning method. Moreover, the fiber production rate was high. Still, electrospinning has been widely used for nanofiber formation.
The centrifugal jet spinning method has been effectively used to form micro- or nanofibers [68]. It is an advantageous method due to the low cost, efficiency, and high throughput to form the fibers. Moreover, this technique has the advantages of the precise handling of centrifugal forces, viscoelastic properties, and mass transfer features of the desired solution used for the nanofibers [69]. The centrifugal jet spinning setup contains a DC motor, a spinning chamber, and multiple fiber collectors [70]. For solution-processed nanofibers, centrifugal jet spinning has been found to be as beneficial as the electrospinning technique [71].
Electrohydrodynamic direct writing (with a mechano-electrospinning process) has been successfully applied to form the micro- or nanofibers [72]. This technique involves the use of a mechano-electrospinning process for the constant/programmable direct writing of the fibers [73]. The method has combined the electrical and mechanical forces to form viscous ink for the large-scale production of the fibers [74]. The main fabrication steps contain the alteration of the electrical field and mechanical drawing force to control the nanofiber size and morphology, changeable nozzle-to-substrate distance, and adjustable applied voltage [75]. However, electrospinning has been more frequently used due to the simple equipment and easily controllable processing conditions for nanofibers.

3. Graphene

Graphene is a two-dimensional nanostructure consisting of sp2 hybridized carbon atoms [76]. In 2004, Andre Geim and Konstantin Novoselov prepared and reported single-layer graphene [77]. However, theoretically, graphene was reported earlier in 1947 (P. R. Wallace) and then experimentally explored in 1962 [78]. Graphene has been synthesized using various techniques such as mechanical cleavage of graphite, exfoliation of graphite, plasma and laser techniques, chemical vapor deposition, and chemical routes [79]. Graphene revealed excellent structural and physical properties. Graphene is a very thin material and is one atom thick [80]. Graphene has a high electron mobility of ~200,000 cm2 V−1s−1. In addition, graphene has a high thermal conductivity of ~3000–5000 W/mK [81]. Young’s modulus of graphene has been found ~1 TPa, which means that it is 200 times stronger than steel [82]. Due to van der Waals forces, graphene nanosheets may possess wrinkling features [83]. To avoid crumpling problems, graphene has been functionalized through various facile techniques [84].
Graphene has been oxidized to introduce the carbonyl, epoxide, hydroxyl, and carboxylic acid groups on the surface. The oxidized form of graphene is commonly referred to as graphene oxide. Among modified forms of graphene, graphene oxide has a unique structure with hydrophilic functional groups [84]. For synthesis, Brodie’s method has used a mixture of potassium chlorate and nitric acid to form graphene oxide from graphite [85]. Hummers and Offeman methods have been used to form graphene oxide from graphite using sodium nitrate, sulfuric acid, and potassium permanganate [86,87]. Both graphene and graphene oxide have been used as nanofillers in polymeric nanocomposites and nanocomposite nanofibers for high-performance technical applications. Figure 1 depicts the structures of graphene and graphene oxide nanosheets.
Graphene and graphene oxide reveal superior heat stability, electrical conductivity, thermal conductivity, mechanical strength, and chemical stability properties [88]. These features of graphene have been explored for the development of high-performance nanocomposites [89,90]. Graphene and modified graphene nanomaterials have found applications in aerospace/automotive [91], energy devices [92], electronics [93,94], sensors [95], membranes [96], and several other technical fields. Table 2 shows important graphene-based nanocomposite nanofiber systems prepared using varying techniques and related physicochemical properties and parameters.

4. Manufacturing and Properties of Polymer/Graphene Nanocomposite Nanofibers

The polymer/graphene nanocomposite nanofibers have been manufactured through various facile procedures [107]. The commonly adopted techniques include wet spinning [108], melt spinning [109], electrostatic spinning [110], and other solution and chemical methods. Spinning methods have been considered advantageous due to less graphene agglomeration in the resulting nanofibers [111]. Additionally, the spinning processes have numerous controllable fabrication parameters and facile processing [112]. Among spinning methods, electrospinning has been widely used to manufacture nanocomposite nanofibers [113,114]. Accordingly, electrospinning has been effectively applied to form polymer/graphene nanocomposite nanofibers [115]. These nanofibers have been investigated for superior conducting, mechanical, thermal, and other enhanced physio-chemical characteristics [116]. Consequently, polymer/graphene nanocomposite nanofibers have been investigated for industrial-level manufacturing and uses. The electrospun graphene nanofibers have been found promising for advanced sensor technologies [117,118,119]. Electrospinning is considered a simple, inexpensive, and multipurpose technique to manufacture nanofibers having diameters down to nanometers [120]. Figure 2 shows a simple electrospinning setup.
In this technique, electric force has been used to draw charged filaments of polymer solution or polymer melt [121]. Usually, a characteristic electrospinning setup has three gears, including (i) a syringe with a pumping system; (ii) a high-voltage power supply for charging the polymer solution; and (iii) a grounded collector to collect the nanofibers. During this process, the syringe was used to pump the polymer solution at a constant rate. Consequently, the polymer solution was extracted from a syringe needle. Under applied voltage, the polymer solution or melt gets charged and stretched to form the fibers due to electrostatic repulsions and surface tension [122]. The polymer jet is usually elongated through the whipping process. Consequently, the charged polymer jet travels to the grounded collector, and nanofibers get deposited. The thin nanofibers having uniform texture and nanometer-scale diameter have been developed using this technique [123]. Two types of electrospinning processes have been seen in the literature including the vertical and horizontal electrospinning systems. The difference in vertical and horizontal electrospinning systems caused changes in the fiber orientation, fiber surface texture, and fiber morphology. Typically, horizontal electrospinning has resulted in random fiber orientation. On the other hand, vertical electrospinning revealed uniform fiber alignment and smooth texture. Moreover, the vertical electrospinning system has facile parameter optimization and process monitoring. The resulting nanofibers have uniform surfaces and less beaded texture. This set up has been used for lab-scale nanofiber processing. For industrial-level electrospinning, several design challenges need to be overcome. Melt electrospinning has also been used for polymer and nanocomposite nanofiber formation. Similar to solution-based electrospinning, the melt spinning setup consists of a needle, syringe, heating system, tubular collector, high-voltage power supply, and polymer for melt formation. Most fiber processing conditions of melt and solution electrospinning have been found similar. In the case of high molecular weight polymers with less solution solubility, melt spinning has been found useful. However, this technique may result in nanofibers of large diameter. There are some differences between the electrospinning techniques for the polymer melts and polymer solution processing. For example, in the case of melt electrospinning, a continuous heat supply has been required to keep the polymer in molten form for fiber formation. Therefore, this technique uses more power and cost. For efficient melt-based fiber processing, the distance between the needle tip and collector must be reduced. Similar to polymer, the nanocomposite solution/melt can be electrospun to form the nanofibers. The resulting nanocomposite nanofibers have high surface area, controllable surface configuration, porosity, and uniform texture [124,125,126].
One-dimensional nanostructures of polyaniline (a conducting polymer) have been designed with varying shapes and unique characteristics [127,128,129]. These nanostructures include polyaniline nanotubes, nanowhisker, and nanofiber forms. Polyaniline nanostructures are usually manufactured through a variety of procedures including chemical, solution, and spinning processing [130,131]. Graphene-filled polyaniline nanofibers have been fabricated and studied for electrical conductivity and other physical properties [132,133]. Zhou et al. [134] manufactured the pristine polyaniline and polyaniline-wrapped graphene nanofibers through wrapping and electrospinning techniques. After the electrospinning of polyaniline, graphene-wrapped polyaniline nanofibers were formed through the reduction and accumulation of graphene oxide nanosheets on the nanofiber surface (Figure 3) [40,135,136]. A transmission electron microscopy (TEM) image of graphene-wrapped polyaniline nanofiber depicted uniform graphene coating on the surface (Figure 4). Figure 5 demonstrates the specific capacitance of the polyaniline and graphene-wrapped polyaniline nanofibers. The specific capacitance of nanocomposite nanofibers was found to be considerably higher (250 Fg−1) than the neat polyaniline nanofibers (~175 Fg−1). The results indicated uniform nanofiller dispersion and network formation to promote the charge transfer and capacitance of the nanocomposite nanofibers. The capacitance was also found higher than the reported non-electrospun polymer/graphene nanocomposite electrodes [137].
Additionally, the polyaniline/graphene nanocomposite nanofibers have been fabricated through the in situ polymerization of aniline on the graphene surface in acidic conditions [138,139]. In this method, the sodium dodecyl benzenesulfonate surfactant has been used for better graphene dispersion in the polyaniline nanofibers [140,141,142]. Then, the template method was also used to form the polyaniline/graphene nanocomposite nanofibers having high electrical conductivity and specific capacitance (>400 Fg−1) [143].
High-performance nylon/graphene nanocomposites have been reported in the literature [144,145,146]. The nanofibers of nylon/graphene nanocomposites have been developed. Lee et al. [147] manufactured the solution-blown nylon 6 and graphene flakes-derived nanofibers. Maccaferri et al. [148] fabricated the nylon 6,6/graphene nanocomposite nanofibers by electrospinning technique. The effect of optimum electrospinning parameters and graphene suspension on the nanofiber morphology and physical properties were analyzed. The electrospinning processing parameters were optimized using the characteristics given in Table 3. Figure 6 shows TEM micrographs of nanocomposite nanofibers with 5 and 15 wt.% graphene contents. The nanofiber diameter was in the range of 200–300 nm. With increasing nanofiller loading, a protruded nanofiber surface was observed due to nanofiller aggregation in the matrix. The graphene disposition along the nanofiber surface was observed due to nanosheet stacking on the nanofiber surface. A classical approach was applied to analyze the macroscopic width and thickness of the material. The free volume among nanofibers was found to be filled by the bulk material (Figure 7). The porosity in the nanofibrous mat was observed due to the free volume of polymer (~80%), compared with the total volume of the nanomaterial.
The best quality nylon 6,6/graphene nanofibers (homogeneous surface and no bead formation) with a diameter of 260 nm were attained at a high flow rate of 0.70 mL/h. The effect of graphene on the mechanical properties of nanofibers such as stretch, slide, twist, and re-orientation was observed [149]. The nanofibers’ quality was also analyzed after 20 months of aging. The nanocomposite nanofibers did not reveal any structural or texture alteration when compared with pristine nanofibers. Moreover, the mechanical properties of the nanofibers were also maintained after 20 months of aging, thus confirming the polymer/graphene nanofiber stability.
Leyva-Porras et al. [150] manufactured the electrospun nanofibers of pristine nylon 6 and nitroxide-functionalized graphene oxide reinforced nylon 6 nanocomposites. Figure 8 demonstrates the TEM image of neat Nylon 6 nanofiber. The average nanofiber diameter was ~200 nm. The nanofibers have a length of several microns. The Z-contrast image revealed two nanocomposite nanofibers situated perpendicular to one other. In both the nanofibers, fine graphene nanosheet dispersion was observed. TEM analysis revealed the deposition of functional 1–4 graphene oxide layers in nylon nanofibers. Few graphene oxide nanosheets were found fully embedded within the polymer fiber, whereas few nanosheets protruded from the fiber. Consequently, the simple reinforcement mechanism was found to affect the dispersion properties and nanofiber diameter.
Weise et al. [151] applied the melt spinning technique to fabricate polyamide 6 and polyamide 6/graphene nanocomposite nanofibers. The 3 and 5 wt.% graphene contents were filled in the polyamide 6 matrix using the melt compounding method. The effect of melt spinning process parameters was analyzed on the nanofiber properties. Here, the single filament was produced at a draw ratio of 2.5 and a winding speed of >100 m/min. The length-to-diameter ratio of nanofibers was ~2.0. Figure 9 displays the extrusion line and nanofiber winding system. According to differential scanning calorimetric analysis, neat nylon 6 had two crystalline peaks ~200 °C due to gamma and alpha phases. With the inclusion of graphene, the gamma peak disappeared due to a shift toward the alpha peak. The shift in peaks and merging were observed due to the alteration of the nylon 6 structure with increasing graphene contents. Moreover, the crystallization peak of pristine nylon 6 fibers appeared at ~189 °C, which was shifted to >195 °C with increasing nanofiller contents. Here, graphene was found to behave as a nucleating agent in the matrix to improve the crystallization effects. Additionally, the electrical conductivity of as-spun nanocomposite nanofibers was ~10 μ S m−1, suggesting the anti-static textile application of the nylon 6/graphene nanocomposite [152].
The polystyrene/graphene nanomaterials have been described in the literature [153,154,155]. Consequently, the polystyrene/graphene nanocomposite nanofibers have been processed using solution blending and electrospinning techniques [156]. Huang et al. [157] manufactured the electrospun polystyrene/graphene nanocomposite nanofibers. The hydrophobicity properties of the nanofibers were studied. Ponnamma et al. [158] developed the electrospun nanofibers of polystyrene, polystyrene/cobalt oxide, polystyrene/hexagonal boron nitride, and polystyrene/cobalt oxide/hexagonal boron nitride nanocomposites. The nanofibers were irradiated with gamma radiations to enhance the crosslinking density of the matrix nanofiller. This led to enhanced hydrophobicity and oleophilicity properties of the nanofibers. Figure 10 presents the contact angle studies of the nanofibers before and after gamma irradiation. Moreover, irradiation was found to enhance the surface roughness of the nanofibers. Neat polystyrene nanofibers had a contact angle of ~140°, which was enhanced to ~152° due to the superhydrophobicity of the nanocomposite. Due to these properties, the nanofibers have been applied as oil-water separators to hold oil molecules and filter the water molecules. Table 4 demonstrates the tensile strength and Young’s modulus of the nanocomposite fibers. The mechanical properties of the nanofibers were enhanced with the nanofiller loading [159]. In both the irradiated and non-irradiated samples, the inclusion of two types of nanofillers enhanced the mechanical properties due to synergistic effects. Moreover, the gamma irradiation also improved the mechanical properties owing to matrix–nanofiller crosslinking and strengthening effects.
Poly(vinyl alcohol)/graphene nanocomposites have been developed [160,161,162]. Consequently, the poly(vinyl alcohol)/graphene nanocomposite nanofibers have been processed through electrospinning by Abdah et al. [163]. The inclusion of graphene in nanofibers reduced the diameter of nanofibers from 121 nm to 117 nm. The decrease in diameter was due to better interactions and homogeneous crosslinking in the matrix–nanofiller. The supercapacitor electrode based on poly(vinyl alcohol)/poly(3,4-ethylenedioxythiophene)/graphene nanocomposite nanofibers had a specific capacitance of 224.3 Fg−1 and power density of ~304.4 Wkg−1. The specific capacitance was found higher than the non-filled poly(3,4-ethylenedioxythiophene) (167.92 Fg−1) and poly(vinyl alcohol)/poly(3,4-ethylenedioxythiophene) nanofibers (182.73 Fg−1). Enhancement in specific capacitance of nanocomposite nanofibers was observed due to an interconnecting network formation for electron/charge transportation. Barzegar et al. [164] also manufactured the electrospun poly(vinyl alcohol)/graphene nanocomposite nanofibers. The as-spun hollow nanofibers had uniform surfaces and fine graphene dispersion. Neat poly(vinyl alcohol) nanofibers have a diameter of 130–230 nm. The inclusion of graphene reduced the average diameter of nanofibers to ~238–302 nm. It appears that the graphene in polymer solution enhanced the electrical conductivity properties, so influencing the nanofiber diameters under the applied field effect. Moreover, the inclusion of graphene enhanced the thermal stability of the poly(vinyl alcohol) nanofibers. The maximum decomposition temperature of the nanocomposite nanofibers (393–542 °C) was found higher than the neat poly(vinyl alcohol) nanofibers (383–460 °C). The enhancement in the thermal conductivity of the nanocomposite nanofibers was attributed to the stable interlinked polymer–graphene nanostructure. Bao and co-workers [165] prepared the poly(vinyl acetate) and graphene-derived nanofibers via electrospinning. Nanocomposite nanofibers have chemically tunable optoelectronic properties along with enhanced mechanical and thermal properties, and processing advantages. The addition of graphene has enhanced the optical absorption properties of the poly(vinyl acetate)/graphene nanocomposite. The nanocomposite nanofibers were suggested to have applications in ultrafast photonics. For this application, the diameters of the nanofibers should be smaller than the optical wavelength. In this regard, the polymer/graphene nanocomposite nanofibers were attached on the faces of optical fibers to function as a passive mode-locker for ring laser cavity. The polymer-/graphene-based mode-locking performed better than the polymer absorbers in a wide wavelength range of around 10–30 nm.
Natural polymer- and graphene-based nanocomposite nanofibers have also been reported [166,167,168]. The cellulose, chitosan, and chitin-derived nanofibers have been designed and applied for biomedical applications such as tissue engineering, drug delivery, wound dressing, and other uses [169]. Neibolts et al. [170] designed biodegradable nanofibers of nanofibrillated cellulose, poly(butylene succinate), and graphene-based nanofibers. The electrospinning method was used to form the nanofibers with poly(ethylene glycol) as a compatibilizer. The nanocomposite nanofibers were used for the tissue engineering scaffolds. De Faria et al. [171] formed the electrospun chitosan/poly(lactide-co-glycolide) and graphene oxide-based nanocomposite nanofibers. The biodegradable nanofibers have fine antimicrobial properties towards Escherichia coli, Pseudomonas aeruginosa, and Staphylococcus aureus bacteria. Yoon et al. [172] prepared the electrospun poly(d,l-lactic-co-glycolic acid) and graphene oxide-derived nanocomposite nanofibers. The polymer and nanofiller were well-linked through the interfacial interactions. The hydrophilicity properties as well as the biocompatibility of the nanocomposite nanofibers were observed for the neuronal cells [173].
Other than electrospun polymer/graphene nanofibers as reported above, few reports have been observed on the solution blow spinning formed nanofibers so far [174]. Mishra et al. [175] designed the metal-doped polymer/graphene nanofibers using the solution blow spinning technique. The nanofibers have been applied to construct the lithium-ion battery electrodes. Some authors in the literature noticed the polymer/graphene nanofibers produced through the centrifugal jet spinning method [176]. Amir et al. [177] developed the graphene nanoplatelets reinforced polyurethane/phenolic resin nanofibers using the centrifugal jet spinning technique. Here, the nanofiber diameter was greatly affected by the processing parameters such as the polymer concentration, rotation speed, and pressure. The graphene nanoplatelets were found to be finely dispersed and adhered to the nanofibers. Matharu et al. [178] also fabricated the poly(methyl methacrylate)/graphene nanofibers through a centrifugal jet spinning procedure. The graphene nanofiller was loaded around 2–8 wt.% contents. An increasing amount of nanofiller affected the morphology of the nanofibers. Scanning electron microscopy exposed the beaded and porous nature of the nanofibers with an increased amount of graphene contents. The nanofiber diameter was found in the range of 0.75–2.71 mm. Electrohydrodynamic direct writing has been rarely used for polymer graphene nanofibers [179]. For example, Wang et al. [180] designed the homogeneously aligned graphene nanostructures using the direct writing approach. The morphology of nanofibers was controlled by means of the optimized manufacturing parameters. Details of the centrifugal jet spinning, solution blow spinning, and direct writing techniques used for the nanofiber formation and comparison with electrospinning are already stated in Section 2.

5. Applications of Polymer/Graphene Nanocomposite Nanofiber

Natural polymer fibers of cotton, silk, linen, etc. have been studied [181]. A wide variety of synthetic polymer fibers have been produced using polyamide, polyester, polyethylene, polyacrylonitrile, etc. [182,183]. Synthetic polymer nanofibers have low density [184], durability [185], toughness [186], strength [187], abrasion resistance [188], chemical stability [189], and eco-friendliness [190]. The as-prepared polymer fibers have been used to form films, membranes, packings, textiles, and biomedical materials [191,192,193]. Further advancements in the properties of polymer nanofibers have been observed with the nanofiller addition [194]. Polymeric nanocomposite nanofibers have found applications in electronics, sensors, actuators, energy storage materials, and membranes [195]. Electronic devices have been considered a continuous source of environmental pollution [196]. Moreover, electromagnetic radiation may hinder the functioning of electronic systems [197]. Incidentally, nanocomposites have been used as shielding materials against electromagnetic radiation [198,199]. Conducting polymers such as polyaniline-based nanofibers have been developed for electromagnetic interference (EMI) shielding [200,201]. Furthermore, the polyaniline/nanocarbon nanocomposite nanofibers were applied for EMI shielding [202,203,204]. Lyu et al. [205] fabricated the polyaniline/aramid/graphene nanofibers using electrospinning. The nanofibers revealed the high strength of 179 MPa. In addition, the nanocomposite nanofibers had high EMI shielding effectiveness of 30 dB. However, limited attempts have been observed on the polymer/graphene nanofibers for EMI shielding.
Supercapacitors have been researched as promising energy storage devices [206]. Supercapacitors have the capability of efficient charge storage, avoiding any chemical reactions [207]. Nanocarbon-based materials have been successfully used in the supercapacitor electrodes and electrolytes, due to high electrical conductivity, optimum porosity, and chemical stability properties [208]. Consequently, the carbon nanoparticles like graphite, graphene, carbon nanotube, etc. have been focused [209,210,211]. Polymer/nanocarbon nanocomposites have also been used to design supercapacitor electrodes [212]. Rose et al. [213] manufactured the polyaniline/poly(vinyl alcohol)/graphene oxide nanofibers for supercapacitor electrodes. The nanocomposite nanofibers-based electrode had a high surface area and charge carrier mobility, leading to high capacitance performance. The manufacturing of polymer/graphene nanofiber-based electrodes also demands comprehensive studies for further advancements.
An important use of polymeric nanofibers has been observed for membrane fabrication [214,215]. The membranes based on the polymer/graphene and polymer/graphene oxide nanocomposite nanofibers have been reported [216,217,218]. The poly(D, L-lactic-co-glycolic acid)/graphene oxide nanocomposite nanofibrous membranes have superior mechanical properties due to fine graphene dispersion and interfacial interactions in the matrix–nanofiller [172,219,220]. The polyacrylonitrile/graphene and polyacrylonitrile/graphene oxide-derived nanofibrous membranes have been prepared [221,222,223]. The membranes were analyzed for physical properties and water purification performances [224]. The addition of graphene nanofillers facilitated the permeation and filtration properties of the membranes [225].
Essential uses of polymer/graphene nanofibers have been observed in the biomedical field. The nanocomposite nanofibers have great potential for biomedical applications due to biocompatibility and bioresorbability properties (Figure 11) [226]. Various patents have been reported for the application of graphene-based nanofibers in biomedical fields [227,228]. Graphene nanocomposites have been applied for tissue engineering, pharmaceutical, drug delivery, gene delivery, wound healing, and biomedical devices and systems [229,230]. For pharmaceutical applications such as drug transfusion, the nanofibers have fine biocompatibility and anionic-exchange properties.
Systematic studies have been carried out to assess the safe use of graphene nanofibers in pharmaceutical and drug delivery applications [231]. For these applications, the nanofibers must be biocompatible, cytocompatible, bioresorbable, and non-toxic. The potential pharmaceutical use of polymer/graphene nanofibers has been observed for cancer therapy [232]. The electrospun graphene nanocomposite nanofibers have the capability for proliferation, differentiation, cell adhesion, and viability for drug delivery systems. Consequently, nanofibers have found use in regenerative medicine [233]. Gupta et al. [234] prepared graphene nanofibers-based sorbent materials for managing pharmaceutical pollution. Inclusion of 20 wt.% graphene oxide in cellulose nanofibers enhanced the adsorption removal capacity up to 45.04 mg g−1 and 85.30 mg g−1 for ciprofloxacin and ofloxacin, respectively.
Graphene oxide nanofibers have found important uses in wound healing applications [235]. Recent patents revealed the wound-healing capability of graphene nanofibers in the medical sector [236]. The chitosan/graphene nanofibrous membranes were developed for anti-bacterial purposes [237]. The poly(vinyl alcohol))/graphene nanocomposite nanofibers have been manufactured for tissue engineering applications [238,239]. Moreover, poly(D, L-lactic-co-glycolic acid)/graphene oxide nanocomposite nanofibers have been studied as bioengineering scaffolds [240,241].
Applications of graphene-containing nanocomposite nanofibers have been used for medical devices including coatings, sensors, and triboelectric nanogenerators [242,243]. Progress has been observed in the areas of sensing and biosensing [244,245]. Moreover, the polymer/graphene nanofibers were researched for cardiac patches and other medical devices [246].

6. Future Prospects

In this article, the manufacturing and properties of several polymer/graphene nanocomposite nanofibers systems have been reviewed. The factors such as graphene contents, dispersal, and matrix–nanofiller interactions have been found significant to determine the physical features of the polymer/graphene nanocomposite nanofibers such as microstructure, thermal, mechanical, conducting, and other characteristics. Graphene-based polymeric nanocomposites always face major challenges regarding dispersion. In particular, graphene dispersion may be a challenging problem during large-scale production of related nanocomposite materials. Here, the nanofiller aspect ratio, orientation, polymer type, and nanofiller contents directly affect the nanofiller dispersion in the matrix. Due to consistent nanofiller dispersion, there is a large interfacial area between the matrix and graphene reinforcement. The implementation of an efficient synthesis technique, optimum processing conditions, and the choice of the matrix have also been considered important for fine dispersion and enhanced physical properties. Due to fine dispersion, the microstructural properties of the nanocomposite nanofibers have been found to be enhanced. An important effect of graphene dispersion has been observed on the formation of percolation networks in the polymer matrices to enhance the electrical conductivity properties. Owing to fine nanoparticle dispersal and matrix–nanofiller bonding, stress transfer through the nanocomposite has been increased, leading to high mechanical properties. The thermal stability properties of the polymer/graphene nanocomposite nanofibers have also been found to depend upon the graphene dispersion and interface formation for enhancing the heat stability of the materials. Advanced structure–property relationships have been suggested to establish the high efficiency and applications of the nanocomposite nanofibers.
The graphene-derived nanofibers have been fabricated using the in situ solution/chemical approaches, template, method, melt processing, and electrospinning techniques. Among all techniques, electrospinning has been found to be the most efficient to enhance the morphological and physical properties (conducting, thermal, mechanical, etc.) of the polymer/graphene nanocomposite nanofibers. In electrospinning, a polymer solution or melt is initially prepared for better graphene dispersion. Graphene dispersion has been found as an important factor to define the texture and microstructure of the nanofibers. Significant application areas of the polymer/graphene nanocomposite nanofibers include EMI shielding, supercapacitors, membranes, and the biomedical field. In the case of EMI shielding materials, nanofibers need to be developed using a modified graphene nanofiller to enhance the electrical conductivity and EMI shielding effectiveness. Similarly, for the development of supercapacitor electrodes, functional graphene can form an interconnecting network to support electron and charge transportation. Consequently, the fields of EMI shielding and supercapacitors need further research efforts to attain high-performance materials. Few research endeavors have been observed for the polymer/graphene nanocomposite nanofibrous membranes aiming for biomedical applications. Focused future attempts have been required to reveal the true potential of these materials for pharmaceutical and biomedical systems. Additionally, the novel polymer/graphene nanocomposite nanofibers must be manufactured and researched for the aerospace, automobile, coatings, and textile fields. Here, challenges regarding graphene dispersion, modification, ultrahigh conductivity, heat stability, and strength properties need to be overcome, as discussed above.
For effective future utilization of polymer/graphene nanocomposite nanofibers in EMI shielding, supercapacitors, membranes, biomedical, and other fields, certain challenges need to be overcome. Most importantly, the electrospun polymer/graphene nanocomposites have attained progress for EMI shielding. For the future application of polymer/graphene-based nanocomposite nanofibers in the EMI shielding materials, in addition to the fine nanofiller dispersion, the thickness and weight of these shields must be decreased to enhance the radiation shielding impact. In this concern, the optimization of polymer nanocomposite nanofiber properties during fabrication has been required for better EMI shielding. The nanofiller functionalization for the nanocomposite nanofibers definitely enhances the dispersion, interactions, conductive, and EMI shielding properties. For practical applications, the new hybrid systems have been designed for EMI shielding to efficiently function in the range of 8.2–12.4 GHz with an effectiveness of >30–50 dB [247]. Moreover, the design of graphene-derived layered nanocomposite shields with 8 to 16 layers can better reduce the impedance mismatch between the adjacent layers to enhance the EMI shielding performance. Hence, future high-performance shields can be developed by adjusting the number of layers to meet the commercial criterion. Additionally, there exist challenges to the commercialization of these nanofibrous EMI shields [248]. Presently, EMI shielding application demands the use of nanocomposite materials in harsh environments of high or low temperatures, and strong acids or strong bases medium. The surface of electrospun nanofibers must be modified with inorganic dopants to enhance their resistance properties. Moreover, the polymer/graphene nanofibers must have strong interfacial interactions to overcome the loss of mechanical performance during EMI shielding. Furthermore, the advanced EMI shields for radar technology must be developed with a wide-band frequency range of 8–26.5 GHz [249]. Undoubtedly, the electrospun polymer/graphene nanofibers have been advantageously used for supercapacitors; however, key issues have been faced during their use [250]. Generally, polymer/nanofibers have a large specific surface area, high electrical conductivity, and high electrochemical activity toward high-performance supercapacitor electrodes. In addition, the microporous nanofibrous structure may enhance the specific surface area to facilitate the ion transport and capacitance properties at low current densities. On the other hand, strong mechanical robustness has been desirable for the mesoporous electrodes while keeping flexibility. Here, it has been a challenge to integrate the porous functionalities with strength and flexibility together in the polymer/graphene nanofiber electrodes. Moreover, the mass production of these electrospun electrodes at low cost for large-scale applications has been the key issue. Graphene derived micro-supercapacitors designed on paper substrates can be a good addition to the future energy storage devices. The micro-supercapacitors have been focused according to the future demands of simple, flexible, and reasonable energy storage devices. To improve the overall device performance, the redox-active species may lead to the volumetric capacitance of 29.6 mF cm−3 [251]. The micro-supercapacitor devices can also be produced using printing techniques. The application of electrospun polymer/graphene nanofibers in membrane application demands the large-scale formation of these materials, which has been found to be difficult while maintaining a low cost [252]. Here, the foremost research analyzed for overcoming the challenges in transforming electrospinning from a lab-scale technology to an industrial-scale for talented applications. In addition, limited designs have been developed and studied for the separation membranes so far. Therefore, more research has been demanded in this area to overcome the related challenges. Additionally, the potential formation of new polymer/graphene nanofibers along with the post-treatment strategies have been desirable and challenging to control the pore formation on the nanofiber surface because, in membrane separation applications, surface area and porosity of nanofibers have been considered as key factors for efficient molecular transportation. Hence, the challenges related to the better separation applications can be overcome through the post-treatment techniques to endorse the transportation mechanism to the microporous nanofibers. The pore clogging issue of polymer/graphene membranes has been observed and solved by varying the membrane designs [253]. The antifouling polymer/graphene membranes have been developed using the functional graphene reinforced polymeric membranes. Here, polymer/graphene oxide-derived nanocomposite nanofiber membranes revealed the least microfouling phenomenon. Another advantage of polymer/graphene-derived nanofibrous membranes have been observed for the formation of freestanding, thin, and strengthened membranes for the water treatment systems. However, the gathering of pollutants (organic/inorganic contaminants like dyes, metal particles, bacteria/viruses, etc.) on the membrane surface can still prevent the rapid passage of clean water [254]. Additionally, the main challenges to using polymer/graphene nanofibers in tissue engineering and drug delivery applications involve attaining better biocompatibility and non-toxicity properties [255]. Challenges in using polymer/nanofiber nanomaterials in tissue engineering involve improving the function of damaged organs by forming precise biological substitutes [256]. In this regard, sufficient related research attempts have been desirable on the advanced polymer/graphene nanofibers. Moreover, the engineered polymer/graphene nanofiber tissues must have the capability to mimic the physiological environment perfectly by upholding their structural, topographical, and mechanical properties. Biocompatible graphene-based 3D scaffolds have been developed having mechanical and organ-biomimicking properties [257]. Owing to the large specific surface area and surface chemistry, the graphene-based 3D scaffolds can better interact with proteins/peptides to form strong interactions. Such nanocomposites have strong capability to direct stem cell differentiation to specific tissues including the cardiac cells, nerve, bone, etc. The porous graphene nanostructures also allow nutrient diffusion/waste discharge during tissue regeneration. Due to high electrical conductivity, graphene oxide-derived scaffolds have been applied for the cardiac and nerve tissue cultures. Further studies have been desirable to explicate these interactions and mechanisms towards organ-specific scaffolds for tissue engineering. Additionally, the challenges in drug delivery of polymer/graphene nanofiber can be overcome by sufficient research attempts in this field [237]. Current graphene research has focuses on the graphene nanomaterials with large surface area for loading biomolecules [258]. Toxicity of graphene-based drug carriers have also been researched for better future designs. Here, major encounters have been found related to the controlled loading and release of drugs in various mediums. New design combinations of polymer and graphene or graphene oxide-derived nanofibers may enhance the effective drug transport capacity for facile uptake by cells, so efficient drug delivery carriers can be designed.

7. Conclusions

This comprehensive review article presents an overview of polymer/graphene-based nanocomposite nanofibers. The design, manufacturing, structural, morphological, and physical characteristics of polymer/graphene nanocomposite nanofibers have been the focus. The polymer and nanocomposite nanofibers have been processed using various manufacturing techniques such as freeze-drying synthesis, interfacial polymerization, phase separation, template synthesis, drawing method, STEP method, electrospinning, and other spinning techniques. For graphene-based nanocomposite nanofibers, electrospinning has been found as a widely used and effective method. For graphene-based nanofibers, the choice of manufacturing technique, type of polymer, graphene form, solvent used, concentration, speed, flow rate, applied field, and other process parameters, and physicochemical properties, influence the quality, morphology, diameter, and applications of the as-prepared nanofibers. The effect of manufacturing techniques and processing parameters on varying the nanofiber diameter and properties have been surveyed. Consequently, the manufactured graphene-based nanofiber properties including flexibility, modulus, tensile strength, toughness, electrical properties, thermal properties, capacitance, hydrophobicity, optical properties, chemical features, and other physical properties have been investigated. High-performance nanocomposite nanofibers have revealed potential for EMI shielding materials, supercapacitors, membranes, pharmaceutical, and biomedical-related devices and systems. Future advancements in the field of polymer/graphene nanocomposite nanofibers and related application areas depend upon the new design possibilities and invention of innovative manufacturing methodologies for these nanofibers, to overcome the related challenges.

Author Contributions

Conceptualization, A.K.; data curation, A.K.; writing of original draft preparation, A.K.; review and editing, A.K., I.A., T.Z., O.A., K.H.I. and M.H.E. All authors have read and agreed to the published version of the manuscript.

Funding

The authors extend their appreciation to the Deputyship for Research & Innovation, Ministry of Education in Saudi Arabia for funding this research through the project number IFP-IMSIU-2023001. The authors also appreciate the Deanship of Scientific Research at Imam Mohammad Ibn Saud Islamic University (IMSIU) for supporting and supervising this project.

Acknowledgments

The authors extend their appreciation to the Deputyship for Research & Innovation, Ministry of Education in Saudi Arabia for funding this research. The authors also appreciate the Deanship of Scientific Research at Imam Mohammad Ibn Saud Islamic University (IMSIU) for supporting and supervising this project.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lee, J.K.Y.; Chen, N.; Peng, S.; Li, L.; Tian, L.; Thakor, N.; Ramakrishna, S. Polymer-based composites by electrospinning: Preparation & functionalization with nanocarbons. Prog. Polym. Sci. 2018, 86, 40–84. [Google Scholar]
  2. Ichakpa, M.; Goodyear, M.; Duthie, J.; Duthie, M.; Wisely, R.; MacPherson, A.; Keyte, J.; Pancholi, K.; Njuguna, J. Investigation on Mechanical and Thermal Properties of 3D-Printed Polyamide 6, Graphene Oxide and Glass-Fibre-Reinforced Composites under Dry, Wet and High Temperature Conditions. J. Compos. Sci. 2023, 7, 227. [Google Scholar] [CrossRef]
  3. Kausar, A.; Ahmad, I.; Zhao, T.; Eisa, M.; Aldaghri, O.; Gupta, M.; Bocchetta, P. Green-Synthesized Graphene for Supercapacitors—Modern Perspectives. J. Compos. Sci. 2023, 7, 108. [Google Scholar] [CrossRef]
  4. Alegaonkar, A.P.; Alegaonkar, P.S. Nanocarbons: Preparation, Assessments, and Applications; CRC Press: Boca Raton, FL, USA, 2023. [Google Scholar]
  5. Sokolov, M.; Nugmanova, A.; Shkolin, A.; Zvyagina, A.; Senchikhin, I.; Kalinina, M. Ion-Mediated Self-Assembly of Graphene Oxide and Functionalized Perylene Diimides into Hybrid Materials with Photocatalytic Properties. J. Compos. Sci. 2023, 7, 14. [Google Scholar] [CrossRef]
  6. Kausar, A.; Ahmad, I.; Bocchetta, P. High-performance corrosion-resistant polymer/graphene nanomaterials for biomedical relevance. J. Compos. Sci. 2022, 6, 362. [Google Scholar] [CrossRef]
  7. Al-Dhahebi, A.M.; Ling, J.; Krishnan, S.G.; Yousefzadeh, M.; Elumalai, N.K.; Saheed, M.S.M.; Ramakrishna, S.; Jose, R. Electrospinning research and products: The road and the way forward. Appl. Phys. Rev. 2022, 9, 011319. [Google Scholar] [CrossRef]
  8. Zhang, X.; Ru, Z.; Sun, Y.; Zhang, M.; Wang, J.; Ge, M.; Liu, H.; Wu, S.; Cao, C.; Ren, X. Recent advances in applications for air pollutants purification and perspectives of electrospun nanofibers. J. Clean. Prod. 2022, 378, 134567. [Google Scholar] [CrossRef]
  9. Bulut, U.; Sayin, V.O.; Altin, Y.; Cevher, S.C.; Cirpan, A.; Bedeloglu, A.C.; Soylemez, S. A flexible carbon nanofiber and conjugated polymer-based electrode for glucose sensing. Microchem. J. 2023, 184, 108148. [Google Scholar] [CrossRef]
  10. Nair, A.B.; Shamsudeen, S.P.; Joys, M.; Varghese, N. Future Perspectives of Polymer Supercapacitors for Advanced Energy Storage Applications. In Polymer Nanocomposites in Supercapacitors; CRC Press: Boca Raton, FL, USA, 2023; pp. 237–257. [Google Scholar]
  11. Xu, X.; Lv, H.; Zhang, M.; Wang, M.; Zhou, Y.; Liu, Y.; Yu, D.-G. Recent progress in electrospun nanofibers and their applications in heavy metal wastewater treatment. Front. Chem. Sci. Eng. 2023, 17, 249–275. [Google Scholar] [CrossRef]
  12. Maliszewska, I.; Czapka, T. Electrospun Polymer Nanofibers with Antimicrobial Activity. Polymers 2022, 14, 1661. [Google Scholar] [CrossRef]
  13. Shinde, S.S.; Kher, J.A. A review on polyaniline and its noble metal composites. Int. J. Innov. Res. Sci. Eng. Technol. 2014, 3, 16570–16576. [Google Scholar] [CrossRef]
  14. Zhao, G.; Shi, L.; Yang, G.; Zhuang, X.; Cheng, B. 3D fibrous aerogels from 1D polymer nanofibers for energy and environmental applications. J. Mater. Chem. A 2023, 11, 512–547. [Google Scholar] [CrossRef]
  15. Nemati, S.; Kim, S.-j.; Shin, Y.M.; Shin, H. Current progress in application of polymeric nanofibers to tissue engineering. Nano Converg. 2019, 6, 36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Chaudhari, S.; Sharma, Y.; Archana, P.S.; Jose, R.; Ramakrishna, S.; Mhaisalkar, S.; Srinivasan, M. Electrospun polyaniline nanofibers web electrodes for supercapacitors. J. Appl. Polym. Sci. 2013, 129, 1660–1668. [Google Scholar] [CrossRef]
  17. Aliheidari, N.; Aliahmad, N.; Agarwal, M.; Dalir, H. Electrospun Nanofibers for Label-Free Sensor Applications. Sensors 2019, 19, 3587. [Google Scholar] [CrossRef] [Green Version]
  18. Ibrahim, H.M.; Klingner, A. A review on electrospun polymeric nanofibers: Production parameters and potential applications. Polym. Test. 2020, 90, 106647. [Google Scholar] [CrossRef]
  19. Wang, C.; Wang, J.; Zeng, L.; Qiao, Z.; Liu, X.; Liu, H.; Zhang, J.; Ding, J. Fabrication of electrospun polymer nanofibers with diverse morphologies. Molecules 2019, 24, 834. [Google Scholar] [CrossRef] [Green Version]
  20. Zhang, F.; Fan, J.b.; Wang, S. Interfacial polymerization: From chemistry to functional materials. Angew. Chem. Int. Ed. 2020, 59, 21840–21856. [Google Scholar] [CrossRef]
  21. Raaijmakers, M.J.; Benes, N.E. Current trends in interfacial polymerization chemistry. Prog. Polym. Sci. 2016, 63, 86–142. [Google Scholar] [CrossRef]
  22. Beachley, V.; Wen, X. Polymer nanofibrous structures: Fabrication, biofunctionalization, and cell interactions. Prog. Polym. Sci. 2010, 35, 868–892. [Google Scholar] [CrossRef] [Green Version]
  23. Zhang, X.; Lu, Y. Centrifugal spinning: An alternative approach to fabricate nanofibers at high speed and low cost. Polym. Rev. 2014, 54, 677–701. [Google Scholar] [CrossRef]
  24. Huang, Y.; Song, J.; Yang, C.; Long, Y.; Wu, H. Scalable manufacturing and applications of nanofibers. Mater. Today 2019, 28, 98–113. [Google Scholar] [CrossRef]
  25. Liapis, A.I.; Bruttini, R. Freeze drying. In Handbook of Industrial Drying; CRC Press: Boca Raton, FL, USA, 2020; pp. 309–343. [Google Scholar]
  26. Stojanovska, E.; Canbay, E.; Pampal, E.S.; Calisir, M.D.; Agma, O.; Polat, Y.; Simsek, R.; Gundogdu, N.S.; Akgul, Y.; Kilic, A. A review on non-electro nanofibre spinning techniques. RSC Adv. 2016, 6, 83783–83801. [Google Scholar] [CrossRef]
  27. Malgras, V.; Ji, Q.; Kamachi, Y.; Mori, T.; Shieh, F.-K.; Wu, K.C.-W.; Ariga, K.; Yamauchi, Y. Templated synthesis for nanoarchitectured porous materials. Bull. Chem. Soc. Jpn. 2015, 88, 1171–1200. [Google Scholar] [CrossRef]
  28. Zhang, Z.-M.; Duan, Y.-S.; Xu, Q.; Zhang, B. A review on nanofiber fabrication with the effect of high-speed centrifugal force field. J. Eng. Fibers Fabr. 2019, 14, 1558925019867517. [Google Scholar] [CrossRef] [Green Version]
  29. Wang, J.; Nain, A.S. Suspended micro/nanofiber hierarchical biological scaffolds fabricated using non-electrospinning STEP technique. Langmuir 2014, 30, 13641–13649. [Google Scholar] [CrossRef]
  30. Jin, L.; Xu, Q.; Li, C.; Huang, J.; Zhang, Y.; Wu, D.; Wang, Z. Engineering 3D aligned nanofibers for regulation of cell growth behavior. Macromol. Mater. Eng. 2017, 302, 1600448. [Google Scholar] [CrossRef]
  31. Fajardo-Diaz, J.L.; Morelos-Gomez, A.; Cruz-Silva, R.; Ishii, K.; Yasuike, T.; Kawakatsu, T.; Yamanaka, A.; Tejima, S.; Izu, K.; Saito, S. Low-pressure reverse osmosis membrane made of cellulose nanofiber and carbon nanotube polyamide nano-nanocomposite for high purity water production. Chem. Eng. J. 2022, 448, 137359. [Google Scholar] [CrossRef]
  32. Amer Flayeh, A.; Jawad Kadhim, H. Enhancing the physical properties of polystyrene nanofibers by adding multiwall carbon nanotubes and natural dye. Fuller. Nanotub. Carbon Nanostruct. 2022, 30, 1090–1096. [Google Scholar] [CrossRef]
  33. Liao, Y.; Zhang, C.; Zhang, Y.; Strong, V.; Tang, J.; Li, X.-G.; Kalantar-Zadeh, K.; Hoek, E.M.; Wang, K.L.; Kaner, R.B. Carbon nanotube/polyaniline composite nanofibers: Facile synthesis and chemosensors. Nano Lett. 2011, 11, 954–959. [Google Scholar] [CrossRef]
  34. Patil, P.T.; Anwane, R.S.; Kondawar, S.B. Development of electrospun polyaniline/ZnO composite nanofibers for LPG sensing. Procedia Mater. Sci. 2015, 10, 195–204. [Google Scholar] [CrossRef] [Green Version]
  35. Kayaci, F.; Ozgit-Akgun, C.; Donmez, I.; Biyikli, N.; Uyar, T. Polymer-inorganic core-shell nanofibers by electrospinning and atomic layer deposition: Flexible nylon-ZnO core-shell nanofiber mats and their photocatalytic activity. ACS Appl. Mater. Interfaces 2012, 4, 6185–6194. [Google Scholar] [CrossRef]
  36. Kellenberger, A.; Plesu, N.; Mihali, M.T.-L.; Vaszilcsin, N. Synthesis of polyaniline nanostructures by electrochemical deposition on niobium. Polymer 2013, 54, 3166–3174. [Google Scholar] [CrossRef]
  37. Snari, R.M.; Bayazeed, A.; Ibarhiam, S.F.; Alnoman, R.B.; Attar, R.; Abumelha, H.M.; El-Metwaly, N.M. Solution blowing spinning of polylactate/polyvinyl alcohol/ZnO nanocomposite toward green and sustainable preparation of wound dressing nanofibrous films. Microsc. Res. Tech. 2022, 85, 3860–3870. [Google Scholar] [CrossRef] [PubMed]
  38. Costa, R.G.; Brichi, G.S.; Ribeiro, C.; Mattoso, L.H. Nanocomposite fibers of poly(lactic acid)/titanium dioxide prepared by solution blow spinning. Polym. Bull. 2016, 73, 2973–2985. [Google Scholar] [CrossRef]
  39. Ikegame, M.; Tajima, K.; Aida, T. Template synthesis of polypyrrole nanofibers insulated within one-dimensional silicate channels: Hexagonal versus lamellar for recombination of polarons into bipolarons. Angew. Chem. Int. Ed. 2003, 42, 2154–2157. [Google Scholar] [CrossRef] [PubMed]
  40. Huang, J. Syntheses and applications of conducting polymer polyaniline nanofibers. Pure Appl. Chem. 2006, 78, 15–27. [Google Scholar] [CrossRef]
  41. Shah, A.A.; Cho, Y.H.; Nam, S.-E.; Park, A.; Park, Y.-I.; Park, H. High performance thin-film nanocomposite forward osmosis membrane based on PVDF/bentonite nanofiber support. J. Ind. Eng. Chem. 2020, 86, 90–99. [Google Scholar] [CrossRef]
  42. Li, Y.; Li, J.; Soria, R.B.; Volodine, A.; Van der Bruggen, B. Aramid nanofiber and modified ZIF-8 constructed porous nanocomposite membrane for organic solvent nanofiltration. J. Membr. Sci. 2020, 603, 118002. [Google Scholar] [CrossRef]
  43. Huang, Y.; Zhou, T.; He, S.; Xiao, H.; Dai, H.; Yuan, B.; Chen, X.; Yang, X. Flame-retardant polyvinyl alcohol/cellulose nanofibers hybrid carbon aerogel by freeze drying with ultra-low phosphorus. Appl. Surf. Sci. 2019, 497, 143775. [Google Scholar] [CrossRef]
  44. Mueller, S.; Sapkota, J.; Nicharat, A.; Zimmermann, T.; Tingaut, P.; Weder, C.; Foster, E.J. Influence of the nanofiber dimensions on the properties of nanocellulose/poly(vinyl alcohol) aerogels. J. Appl. Polym. Sci. 2015, 132, 41740. [Google Scholar] [CrossRef]
  45. Qian, L.; Zhang, H. Controlled freezing and freeze drying: A versatile route for porous and micro-/nano-structured materials. J. Chem. Technol. Biotechnol. 2011, 86, 172–184. [Google Scholar] [CrossRef]
  46. Abdal-hay, A.; Hamdy Makhlouf, A.S.; Khalil, K.A. Novel, facile, single-step technique of polymer/TiO2 nanofiber composites membrane for photodegradation of methylene blue. ACS Appl. Mater. Interfaces 2015, 7, 13329–13341. [Google Scholar] [CrossRef] [PubMed]
  47. Pan, Z.; Zhai, J.; Shen, B. Multilayer hierarchical interfaces with high energy density in polymer nanocomposites composed of BaTiO3@TiO2@Al2O3 nanofibers. J. Mater. Chem. A 2017, 5, 15217–15226. [Google Scholar] [CrossRef]
  48. Zhang, Y.; Venugopal, J.R.; El-Turki, A.; Ramakrishna, S.; Su, B.; Lim, C.T. Electrospun biomimetic nanocomposite nanofibers of hydroxyapatite/chitosan for bone tissue engineering. Biomaterials 2008, 29, 4314–4322. [Google Scholar] [CrossRef] [PubMed]
  49. Chang, W.-M.; Wang, C.-C.; Chen, C.-Y. Plasma-induced polyaniline grafted on carbon nanotube-embedded carbon nanofibers for high-performance supercapacitors. Electrochim. Acta 2016, 212, 130–140. [Google Scholar] [CrossRef]
  50. Rezaei, F.; Gorbanev, Y.; Chys, M.; Nikiforov, A.; Van Hulle, S.W.; Cos, P.; Bogaerts, A.; De Geyter, N. Investigation of plasma-induced chemistry in organic solutions for enhanced electrospun PLA nanofibers. Plasma Process. Polym. 2018, 15, 1700226. [Google Scholar] [CrossRef]
  51. Mittal, V. Synthesis Techniques for Polymer Nanocomposites; John Wiley & Sons: Hoboken, NJ, USA, 2015. [Google Scholar]
  52. Arjmandi, S.K.; Khademzadeh Yeganeh, J.; Zare, Y.; Rhee, K.Y. Development of Kovacs model for electrical conductivity of carbon nanofiber-polymer systems. Sci. Rep. 2023, 13, 7. [Google Scholar] [CrossRef] [PubMed]
  53. Sundaray, B.; Choi, A.; Park, Y.W. Highly conducting electrospun polyaniline-polyethylene oxide nanofibrous membranes filled with single-walled carbon nanotubes. Synth. Met. 2010, 160, 984–988. [Google Scholar] [CrossRef]
  54. Kausar, A. Thermally conducting polymer/nanocarbon and polymer/inorganic nanoparticle nanocomposite: A review. Polym. -Plast. Technol. Mater. 2020, 59, 895–909. [Google Scholar] [CrossRef]
  55. Takeno, H.; Shikano, R.; Kikuchi, R. Mechanical Performance of Corn Starch/Poly(Vinyl Alcohol) Composite Hydrogels Reinforced by Inorganic Nanoparticles and Cellulose Nanofibers. Gels 2022, 8, 514. [Google Scholar] [CrossRef]
  56. Shin, Y.J.; Kameoka, J. Amperometric cholesterol biosensor using layer-by-layer adsorption technique onto electrospun polyaniline nanofibers. J. Ind. Eng. Chem. 2012, 18, 193–197. [Google Scholar] [CrossRef]
  57. Haider, A.; Haider, S.; Kang, I.-K. A comprehensive review summarizing the effect of electrospinning parameters and potential applications of nanofibers in biomedical and biotechnology. Arab. J. Chem. 2018, 11, 1165–1188. [Google Scholar] [CrossRef]
  58. Meyer, M. Processing of collagen based biomaterials and the resulting materials properties. Biomed. Eng. Online 2019, 18, 24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Ding, J.; Zhang, J.; Li, J.; Li, D.; Xiao, C.; Xiao, H.; Yang, H.; Zhuang, X.; Chen, X. Electrospun polymer biomaterials. Prog. Polym. Sci. 2019, 90, 1–34. [Google Scholar] [CrossRef]
  60. Bhardwaj, N.; Kundu, S.C. Electrospinning: A fascinating fiber fabrication technique. Biotechnol. Adv. 2010, 28, 325–347. [Google Scholar] [CrossRef]
  61. Cooley, J.F. Improved Methods of and Apparatus for Electrically Separating the Relatively Volatile Liquid Component from the Component of Relatively Fixed Substances of Composite Fluids. UK Patent GB 06385, 1900. [Google Scholar]
  62. Tan, N.P.; Cabatingan, L.K.; Lim, K.J.A. Synthesis of TiO2 nanofiber by solution blow spinning (SBS) method. Key Eng. Mater. 2020, 858, 122–128. [Google Scholar] [CrossRef]
  63. Sinha-Ray, S.; Sinha-Ray, S.; Yarin, A.L.; Pourdeyhimi, B. Theoretical and experimental investigation of physical mechanisms responsible for polymer nanofiber formation in solution blowing. Polymer 2015, 56, 452–463. [Google Scholar] [CrossRef]
  64. Medeiros, E.S.; Glenn, G.M.; Klamczynski, A.P.; Orts, W.J.; Mattoso, L.H. Solution blow spinning: A new method to produce micro-and nanofibers from polymer solutions. J. Appl. Polym. Sci. 2009, 113, 2322–2330. [Google Scholar] [CrossRef]
  65. Liu, R.; Xu, X.; Zhuang, X.; Cheng, B. Solution blowing of chitosan/PVA hydrogel nanofiber mats. Carbohydr. Polym. 2014, 101, 1116–1121. [Google Scholar] [CrossRef]
  66. Oliveira, J.E.; Mattoso, L.H.; Orts, W.J.; Medeiros, E.S. Structural and morphological characterization of micro and nanofibers produced by electrospinning and solution blow spinning: A comparative study. Adv. Mater. Sci. Eng. 2013, 2013, 409572. [Google Scholar] [CrossRef] [Green Version]
  67. Wu, X.F.; Yarin, A.L. Recent progress in interfacial toughening and damage self-healing of polymer composites based on electrospun and solution-blown nanofibers: An overview. J. Appl. Polym. Sci. 2013, 130, 2225–2237. [Google Scholar] [CrossRef]
  68. Xu, H.; Yagi, S.; Ashour, S.; Du, L.; Hoque, M.E.; Tan, L. A Review on Current Nanofiber Technologies: Electrospinning, Centrifugal Spinning, and Electro-Centrifugal Spinning. Macromol. Mater. Eng. 2023, 308, 2200502. [Google Scholar] [CrossRef]
  69. Ren, L.; Kotha, S.P. Centrifugal jet spinning for highly efficient and large-scale fabrication of barium titanate nanofibers. Mater. Lett. 2014, 117, 153–157. [Google Scholar] [CrossRef] [Green Version]
  70. Marjuban, S.M.H.; Rahman, M.; Duza, S.S.; Ahmed, M.B.; Patel, D.K.; Rahman, M.S.; Lozano, K. Recent Advances in Centrifugal Spinning and Their Applications in Tissue Engineering. Polymers 2023, 15, 1253. [Google Scholar] [CrossRef]
  71. Alsharif, A.M. Power law liquid jets’ trajectories and instability during centrifugal spinning. Alex. Eng. J. 2023, 68, 301–314. [Google Scholar] [CrossRef]
  72. Huang, Y.; Duan, Y.; Ding, Y.; Bu, N.; Pan, Y.; Lu, N.; Yin, Z. Versatile, kinetically controlled, high precision electrohydrodynamic writing of micro/nanofibers. Sci. Rep. 2014, 4, 5949. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Duan, Y.; Ding, Y.; Xu, Z.; Huang, Y.; Yin, Z. Helix electrohydrodynamic printing of highly aligned serpentine micro/nanofibers. Polymers 2017, 9, 434. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Zhang, Z.; He, H.; Fu, W.; Ji, D.; Ramakrishna, S. Electro-hydrodynamic direct-writing technology toward patterned ultra-thin fibers: Advances, materials and applications. Nano Today 2020, 35, 100942. [Google Scholar] [CrossRef]
  75. Yin, Z.; Huang, Y.; Duan, Y.; Zhang, H. Electrohydrodynamic Direct-Writing for Flexible Electronic Manufacturing; Springer: Berlin/Heidelberg, Germany, 2018. [Google Scholar]
  76. Gao, Y.; Zhang, Y.; Chen, P.; Li, Y.; Liu, M.; Gao, T.; Ma, D.; Chen, Y.; Cheng, Z.; Qiu, X. Toward single-layer uniform hexagonal boron nitride-graphene patchworks with zigzag linking edges. Nano Lett. 2013, 13, 3439–3443. [Google Scholar] [CrossRef]
  77. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.-e.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric field effect in atomically thin carbon films. Science 2004, 306, 666–669. [Google Scholar] [CrossRef] [Green Version]
  78. Barkan, T. Graphene: The hype versus commercial reality. Nat. Nanotechnol. 2019, 14, 904–906. [Google Scholar] [CrossRef] [PubMed]
  79. Wei, C.; Negishi, R.; Ogawa, Y.; Akabori, M.; Taniyasu, Y.; Kobayashi, Y. Turbostratic multilayer graphene synthesis on CVD graphene template toward improving electrical performance. Jpn. J. Appl. Phys. 2019, 58, SIIB04. [Google Scholar] [CrossRef]
  80. Narayanam, P.K.; Botcha, V.D.; Ghosh, M.; Major, S.S. Growth and Photocatalytic Behaviour of Transparent Reduced GO-ZnO Nanocomposite Sheets. Nanotechnology 2019, 30, 485601. [Google Scholar] [CrossRef]
  81. Shen, X.J.; Zeng, X.L.; Dang, C.Y. Graphene Composites. Handb. Graphene 2019, 1, 1–25. [Google Scholar]
  82. Zandiatashbar, A.; Lee, G.-H.; An, S.J.; Lee, S.; Mathew, N.; Terrones, M.; Hayashi, T.; Picu, C.R.; Hone, J.; Koratkar, N. Effect of defects on the intrinsic strength and stiffness of graphene. Nat. Commun. 2014, 5, 3186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Zhou, Q.; Xia, G.; Du, M.; Lu, Y.; Xu, H. Scotch-tape-like exfoliation effect of graphene quantum dots for efficient preparation of graphene nanosheets in water. Appl. Surf. Sci. 2019, 483, 52–59. [Google Scholar] [CrossRef]
  84. Pei, S.; Cheng, H.-M. The reduction of graphene oxide. Carbon 2012, 50, 3210–3228. [Google Scholar] [CrossRef]
  85. Brodie, B.C. XIII. On the atomic weight of graphite. Philos. Trans. R. Soc. Lond. 1859, 149, 249–259. [Google Scholar]
  86. Feicht, P.; Biskupek, J.; Gorelik, T.E.; Renner, J.; Halbig, C.E.; Maranska, M.; Puchtler, F.; Kaiser, U.; Eigler, S. Brodie's or Hummers’ method: Oxidation conditions determine the structure of graphene oxide. Chem. A Eur. J. 2019, 25, 8955–8959. [Google Scholar] [CrossRef] [PubMed]
  87. Hummers Jr, W.S.; Offeman, R.E. Preparation of graphitic oxide. J. Am. Chem. Soc. 1958, 80, 1339. [Google Scholar] [CrossRef]
  88. Lee, H.; Lee, K.S. Interlayer Distance Controlled Graphene, Supercapacitor and Method of Producing the Same. U.S. Patent 10,214,422, 26 February 2019. [Google Scholar]
  89. Bikiaris, N.D.; Koumentakou, I.; Samiotaki, C.; Meimaroglou, D.; Varytimidou, D.; Karatza, A.; Kalantzis, Z.; Roussou, M.; Bikiaris, R.D.; Papageorgiou, G.Z. Recent Advances in the Investigation of Poly(lactic acid) (PLA) Nanocomposites: Incorporation of Various Nanofillers and their Properties and Applications. Polymers 2023, 15, 1196. [Google Scholar] [CrossRef] [PubMed]
  90. Huang, Z.; Li, L.; Wang, Y.; Zhang, C.; Liu, T. Polyaniline/graphene nanocomposites towards high-performance supercapacitors: A review. Compos. Commun. 2018, 8, 83–91. [Google Scholar] [CrossRef]
  91. Wazalwar, R.; Sahu, M. Novel applications of graphene in the aerospace industry. In Novel Applications of Carbon Based Nano-materials; CRC Press: Boca Raton, FL, USA, 2022; pp. 180–198. [Google Scholar]
  92. Tang, C.; Titirici, M.-M.; Zhang, Q. A review of nanocarbons in energy electrocatalysis: Multifunctional substrates and highly active sites. J. Energy Chem. 2017, 26, 1077–1093. [Google Scholar] [CrossRef] [Green Version]
  93. Han, J.T.; Jang, J.I.; Cho, J.Y.; Hwang, J.Y.; Woo, J.S.; Jeong, H.J.; Jeong, S.Y.; Seo, S.H.; Lee, G.-W. Synthesis of nanobelt-like 1-dimensional silver/nanocarbon hybrid materials for flexible and wearable electronics. Sci. Rep. 2017, 7, 4931. [Google Scholar] [CrossRef] [Green Version]
  94. Marconcini, P.; Macucci, M. Transport Simulation of Graphene Devices with a Generic Potential in the Presence of an Orthogonal Magnetic Field. Nanomaterials 2022, 12, 1087. [Google Scholar] [CrossRef]
  95. Panwar, N.; Soehartono, A.M.; Chan, K.K.; Zeng, S.; Xu, G.; Qu, J.; Coquet, P.; Yong, K.-T.; Chen, X. Nanocarbons for biology and medicine: Sensing, imaging, and drug delivery. Chem. Rev. 2019, 119, 9559–9656. [Google Scholar] [CrossRef]
  96. Kausar, A. Applications of polymer/graphene nanocomposite membranes: A review. Mater. Res. Innov. 2019, 23, 276–287. [Google Scholar] [CrossRef]
  97. Abdali, H.; Ajji, A. Preparation of electrospun nanocomposite nanofibers of polyaniline/poly(methyl methacrylate) with amino-functionalized graphene. Polymers 2017, 9, 453. [Google Scholar] [CrossRef]
  98. Ramazani, S.; Karimi, M. Study the molecular structure of poly(ε-caprolactone)/graphene oxide and graphene nanocomposite nanofibers. J. Mech. Behav. Biomed. Mater. 2016, 61, 484–492. [Google Scholar] [CrossRef]
  99. Bagheri, M.; Mahmoodzadeh, A. Polycaprolactone/graphene nanocomposites: Synthesis, characterization and mechanical properties of electrospun nanofibers. J. Inorg. Organomet. Polym. Mater. 2020, 30, 1566–1577. [Google Scholar] [CrossRef]
  100. Ramazani, S.; Karimi, M. Aligned poly(ε-caprolactone)/graphene oxide and reduced graphene oxide nanocomposite nanofibers: Morphological, mechanical and structural properties. Mater. Sci. Eng. C 2015, 56, 325–334. [Google Scholar] [CrossRef] [PubMed]
  101. Li, B.; Yuan, H.; Zhang, Y. Transparent PMMA-based nanocomposite using electrospun graphene-incorporated PA-6 nanofibers as the reinforcement. Compos. Sci. Technol. 2013, 89, 134–141. [Google Scholar] [CrossRef]
  102. Pan, X.; Shen, L.; Schenning, A.P.; Bastiaansen, C.W. Transparent, high-thermal-conductivity ultradrawn polyethylene/graphene nanocomposite films. Adv. Mater. 2019, 31, 1904348. [Google Scholar] [CrossRef] [PubMed]
  103. Sahoo, S.; Dhibar, S.; Hatui, G.; Bhattacharya, P.; Das, C.K. Graphene/polypyrrole nanofiber nanocomposite as electrode material for electrochemical supercapacitor. Polymer 2013, 54, 1033–1042. [Google Scholar] [CrossRef]
  104. Mo, M.; Chen, C.; Gao, H.; Chen, M.; Li, D. Wet-spinning assembly of cellulose nanofibers reinforced graphene/polypyrrole microfibers for high performance fiber-shaped supercapacitors. Electrochim. Acta 2018, 269, 11–20. [Google Scholar] [CrossRef]
  105. Zhang, X.; Wang, A.; Zhou, X.; Chen, F.; Fu, Q. Fabrication of aramid nanofiber-wrapped graphene fibers by coaxial spinning. Carbon 2020, 165, 340–348. [Google Scholar] [CrossRef]
  106. León-Boigues, L.; Flores, A.; Gómez-Fatou, M.A.; Vega, J.F.; Ellis, G.J.; Salavagione, H.J. PET/Graphene Nanocomposite Fibers Obtained by Dry-Jet Wet-Spinning for Conductive Textiles. Polymers 2023, 15, 1245. [Google Scholar] [CrossRef]
  107. Ganguly, S. Preparation/processing of polymer-graphene composites by different techniques. In Polymer Nanocomposites Containing Graphene; Elsevier: Amsterdam, The Netherlands, 2022; pp. 45–74. [Google Scholar]
  108. Jeong, K.; Kim, D.H.; Chung, Y.S.; Hwang, S.K.; Hwang, H.Y.; Kim, S.S. Effect of processing parameters of the continuous wet spinning system on the crystal phase of PVDF fibers. J. Appl. Polym. Sci. 2018, 135, 45712. [Google Scholar] [CrossRef]
  109. Huang, T.; Marshall, L.R.; Armantrout, J.E.; Yembrick, S.; Dunn, W.H.; Oconnor, J.M.; Mueller, T.; Avgousti, M.; Wetzel, M.D. Production of Nanofibers by Melt Spinning. U.S. Patent 20080242171A1, 2 October 2008. [Google Scholar]
  110. Jirsak, O.; Sanetrnik, F.; Lukas, D.; Kotek, V.; Martinova, L.; Chaloupek, J. Method of Nanofibres Production from a Polymer Solution Using Electrostatic Spinning and a Device for Carrying Out the Method. U.S. Patent 7,585,437, 8 September 2009. [Google Scholar]
  111. Li, X.; Cao, M.; Zhang, H.; Zhou, L.; Cheng, S.; Yao, J.-L.; Fan, L.-J. Surface-enhanced Raman scattering-active substrates of electrospun polyvinyl alcohol/gold–silver nanofibers. J. Colloid Interface Sci. 2012, 382, 28–35. [Google Scholar] [CrossRef]
  112. Ismagilov, Z.R.; Shalagina, A.E.; Podyacheva, O.Y.; Ischenko, A.V.; Kibis, L.S.; Boronin, A.I.; Chesalov, Y.A.; Kochubey, D.I.; Romanenko, A.I.; Anikeeva, O.B. Structure and electrical conductivity of nitrogen-doped carbon nanofibers. Carbon 2009, 47, 1922–1929. [Google Scholar] [CrossRef]
  113. Liu, Y.-L.; Li, Y.; Xu, J.-T.; Fan, Z.-Q. Cooperative effect of electrospinning and nanoclay on formation of polar crystalline phases in poly(vinylidene fluoride). ACS Appl. Mater. Interfaces 2010, 2, 1759–1768. [Google Scholar] [CrossRef] [PubMed]
  114. Scarlet, R.; Manea, L.R.; Sandu, I.; Cramariuc, B.; Sandu, A.V. The Influence of the Needle-Collector Distance Upon the Characteristics of the Polyetherimide Nanofibres Obtained by Electrospinning. Rev. Chim. 2012, 63, 777–782. [Google Scholar]
  115. Mahmoudi, N.; Simchi, A. On the biological performance of graphene oxide-modified chitosan/polyvinyl pyrrolidone nanocomposite membranes: In vitro and in vivo effects of graphene oxide. Mater. Sci. Eng. C 2017, 70, 121–131. [Google Scholar] [CrossRef] [PubMed]
  116. Park, S.; Park, K.; Yoon, H.; Son, J.; Min, T.; Kim, G. Apparatus for preparing electrospun nanofibers: Designing an electrospinning process for nanofiber fabrication. Polym. Int. 2007, 56, 1361–1366. [Google Scholar] [CrossRef]
  117. Li, Y.; Zhu, J.; Cheng, H.; Li, G.; Cho, H.; Jiang, M.; Gao, Q.; Zhang, X. Developments of advanced electrospinning techniques: A critical review. Adv. Mater. Technol. 2021, 6, 2100410. [Google Scholar] [CrossRef]
  118. Li, X.; Chen, W.; Qian, Q.; Huang, H.; Chen, Y.; Wang, Z.; Chen, Q.; Yang, J.; Li, J.; Mai, Y.W. Electrospinning-based strategies for battery materials. Adv. Energy Mater. 2021, 11, 2000845. [Google Scholar] [CrossRef]
  119. Luraghi, A.; Peri, F.; Moroni, L. Electrospinning for drug delivery applications: A review. J. Control. Release 2021, 334, 463–484. [Google Scholar] [CrossRef]
  120. Yadav, T.C.; Srivastava, A.K.; Mishra, P.; Singh, D.; Raghuwanshi, N.; Singh, N.K.; Singh, A.K.; Tiwari, S.K.; Prasad, R.; Pruthi, V. Electrospinning: An efficient biopolymer-based micro-and nanofibers fabrication technique. In Next Generation Biomanufacturing Technologies; ACS Publications: Columbus, OH, USA, 2019; pp. 209–241. [Google Scholar]
  121. Yu, D.-G.; Li, Q.; Song, W.; Xu, L.; Zhang, K.; Zhou, T. Advanced technique-based combination of innovation education and safety education in higher education. J. Chem. Educ. 2023, 100, 507–516. [Google Scholar] [CrossRef]
  122. Hosseini Ravandi, S.A.; Sadrjahani, M.; Valipouri, A.; Dabirian, F.; Ko, F.K. Recently developed electrospinning methods: A review. Text. Res. J. 2022, 92, 5130–5145. [Google Scholar] [CrossRef]
  123. Liu, Y.; Hao, M.; Zhou, C.; Yang, B.; Jiang, S.; Huang, J.; Chen, Z.; Liu, Y.; Ramakrishna, S. Scale-up strategies for electrospun nanofiber production. In Electrospun and Nanofibrous Membranes; Elsevier: Amsterdam, The Netherlands, 2023; pp. 205–266. [Google Scholar]
  124. Jha, C.B.; Santhosh, S.; Singh, C.; Bose, S.; Manna, K.; Varshney, R.; Mathur, R. Electrospun nanofiber a smart drug carriers: Production methods, problems, solutions, and applications. In Fiber and Textile Engineering in Drug Delivery Systems, Elsevier: Amsterdam, The Netherlands, 2023; pp. 285–306.
  125. Sun, L.; Miyagi, D.; Cai, Y.; Ullah, A.; Haider, M.K.; Zhu, C.; Gopiraman, M.; Kim, I.S. Rational construction of hierarchical nanocomposites by growing dense polyaniline nanoarrays on carbon black-functionalized carbon nanofiber backbone for freestanding supercapacitor electrodes. J. Energy Storage 2023, 61, 106738. [Google Scholar] [CrossRef]
  126. Jiang, J.; Liu, Y.; Chen, J.; Wang, X.; Yu, Z.; Li, W.; Zheng, G. In-situ molding of micro three-dimensional columnar structure by electric-field-focused electrospinning. Mater. Today Commun. 2023, 35, 105589. [Google Scholar] [CrossRef]
  127. Sumitha, M.; Xavier, T. Synthesis and electrochemical characterization of electrospun biocompatible Poly(ε-caprolactone) and Polyaniline nanofiber composite electrode materials for various biosensing applications. Mater. Today: Proc. 2023, 80, 1297–1305. [Google Scholar] [CrossRef]
  128. Wang, R.; Jing, Y. The effect of inorganic salt on the morphology and nucleation of polyaniline nanofibers synthesized via self-assembly. Des. Monomers Polym. 2023, 26, 45–53. [Google Scholar] [CrossRef] [PubMed]
  129. Guo, Y.; Wei, S.; Chen, Y.; Ye, H.; Xue, S.; Niu, Q.J. Sulfonated polyaniline interlayer with controllable doping conditions for high-performance nanofiltration. J. Membr. Sci. 2023, 672, 121478. [Google Scholar] [CrossRef]
  130. Mostafaei, A.; Nasirpouri, F. Epoxy/polyaniline-ZnO nanorods hybrid nanocomposite coatings: Synthesis, characterization and corrosion protection performance of conducting paints. Prog. Org. Coat. 2014, 77, 146–159. [Google Scholar] [CrossRef]
  131. Menzel, V.C.; Tudela, I. Additive manufacturing of polyaniline-based materials: An opportunity for new designs and applications in energy and biotechnology. Curr. Opin. Chem. Eng. 2022, 35, 100742. [Google Scholar] [CrossRef]
  132. Nakagaito, A.N.; Nogi, M.; Yano, H. Displays from transparent films of natural nanofibers. MRS Bull. 2010, 35, 214–218. [Google Scholar] [CrossRef]
  133. Andrady, A.L. Science and Technology of Polymer Nanofibers; John Wiley & Sons: Hoboken, NJ, USA, 2008. [Google Scholar]
  134. Zhou, S.; Zhang, H.; Zhao, Q.; Wang, X.; Li, J.; Wang, F. Graphene-wrapped polyaniline nanofibers as electrode materials for organic supercapacitors. Carbon 2013, 52, 440–450. [Google Scholar] [CrossRef]
  135. Huang, J.; Kaner, R.B. The intrinsic nanofibrillar morphology of polyaniline. Chem. Commun. 2006, 4, 367–376. [Google Scholar] [CrossRef]
  136. Sapurina, I.; Stejskal, J. The mechanism of the oxidative polymerization of aniline and the formation of supramolecular polyaniline structures. Polym. Int. 2008, 57, 1295–1325. [Google Scholar] [CrossRef]
  137. Ehsani, A.; Heidari, A.A.; Shiri, H.M. Electrochemical pseudocapacitors based on ternary nanocomposite of conductive polymer/graphene/metal oxide: An introduction and review to it in recent studies. Chem. Rec. 2019, 19, 908–926. [Google Scholar] [CrossRef] [PubMed]
  138. Zhang, K.; Zhang, L.L.; Zhao, X.; Wu, J. Graphene/polyaniline nanofiber composites as supercapacitor electrodes. Chem. Mater. 2010, 22, 1392–1401. [Google Scholar] [CrossRef]
  139. Mao, L.; Zhang, K.; Chan, H.S.O.; Wu, J. Surfactant-stabilized graphene/polyaniline nanofiber composites for high performance supercapacitor electrode. J. Mater. Chem. 2012, 22, 80–85. [Google Scholar] [CrossRef]
  140. Noh, Y.J.; Joh, H.-I.; Yu, J.; Hwang, S.H.; Lee, S.; Lee, C.H.; Kim, S.Y.; Youn, J.R. Ultra-high dispersion of graphene in polymer composite via solvent free fabrication and functionalization. Sci. Rep. 2015, 5, 9141. [Google Scholar] [CrossRef]
  141. Wei, T.; Luo, G.; Fan, Z.; Zheng, C.; Yan, J.; Yao, C.; Li, W.; Zhang, C. Preparation of graphene nanosheet/polymer composites using in situ reduction–extractive dispersion. Carbon 2009, 47, 2296–2299. [Google Scholar] [CrossRef]
  142. Liu, S.; Tian, J.; Wang, L.; Li, H.; Zhang, Y.; Sun, X. Stable aqueous dispersion of graphene nanosheets: Noncovalent functionalization by a polymeric reducing agent and their subsequent decoration with Ag nanoparticles for enzymeless hydrogen peroxide detection. Macromolecules 2010, 43, 10078–10083. [Google Scholar] [CrossRef]
  143. Xu, D.; Xu, Q.; Wang, K.; Chen, J.; Chen, Z. Fabrication of free-standing hierarchical carbon nanofiber/graphene oxide/polyaniline films for supercapacitors. ACS Appl. Mater. Interfaces 2014, 6, 200–209. [Google Scholar] [CrossRef] [PubMed]
  144. Jin, J.; Rafiq, R.; Gill, Y.Q.; Song, M. Preparation and characterization of high performance of graphene/nylon nanocomposites. Eur. Polym. J. 2013, 49, 2617–2626. [Google Scholar] [CrossRef]
  145. Zhuang, Y.; Cao, X.; Zhang, J.; Ma, Y.; Shang, X.; Lu, J.; Yang, S.; Zheng, K.; Ma, Y. Monomer casting nylon/graphene nanocomposite with both improved thermal conductivity and mechanical performance. Compos. Part A Appl. Sci. Manuf. 2019, 120, 49–55. [Google Scholar] [CrossRef]
  146. Pan, B.; Zhang, S.; Li, W.; Zhao, J.; Liu, J.; Zhang, Y.; Zhang, Y. Tribological and mechanical investigation of MC nylon reinforced by modified graphene oxide. Wear 2012, 294, 395–401. [Google Scholar] [CrossRef]
  147. Lee, J.-G.; Kim, D.-Y.; Mali, M.G.; Al-Deyab, S.S.; Swihart, M.T.; Yoon, S.S. Supersonically blown nylon-6 nanofibers entangled with graphene flakes for water purification. Nanoscale 2015, 7, 19027–19035. [Google Scholar] [CrossRef] [PubMed]
  148. Maccaferri, E.; Mazzocchetti, L.; Benelli, T.; Zucchelli, A.; Giorgini, L. Morphology, thermal, mechanical properties and ageing of nylon 6,6/graphene nanofibers as Nano2 materials. Compos. Part B Eng. 2019, 166, 120–129. [Google Scholar] [CrossRef]
  149. Xu, Z.; Liu, Y.; Zhao, X.; Peng, L.; Sun, H.; Xu, Y.; Ren, X.; Jin, C.; Xu, P.; Wang, M. Ultrastiff and strong graphene fibers via full-scale synergetic defect engineering. Adv. Mater. 2016, 28, 6449–6456. [Google Scholar] [CrossRef]
  150. Leyva-Porras, C.; Ornelas-Gutiérrez, C.; Miki-Yoshida, M.; Avila-Vega, Y.I.; Macossay, J.; Bonilla-Cruz, J. EELS analysis of Nylon 6 nanofibers reinforced with nitroxide-functionalized graphene oxide. Carbon 2014, 70, 164–172. [Google Scholar] [CrossRef] [Green Version]
  151. Weise, B.A.; Wirth, K.G.; Völkel, L.; Morgenstern, M.; Seide, G. Pilot-scale fabrication and analysis of graphene-nanocomposite fibers. Carbon 2019, 144, 351–361. [Google Scholar] [CrossRef]
  152. Li, B.; Wu, N.; Yang, Y.; Pan, F.; Wang, C.; Wang, G.; Xiao, L.; Liu, W.; Liu, J.; Zeng, Z. Graphene Oxide-Assisted Multiple Cross-Linking of MXene for Large-Area, High-Strength, Oxidation-Resistant, and Multifunctional Films. Adv. Funct. Mater. 2023, 33, 2213357. [Google Scholar] [CrossRef]
  153. Jia, Y.; Chen, L.; Yu, H.; Zhang, Y.; Dong, F. Graphene oxide/polystyrene composite nanofibers on quartz crystal microbalance electrode for the ammonia detection. RSC Adv. 2015, 5, 40620–40627. [Google Scholar] [CrossRef]
  154. Yu, Y.-H.; Lin, Y.-Y.; Lin, C.-H.; Chan, C.-C.; Huang, Y.-C. High-performance polystyrene/graphene-based nanocomposites with excellent anti-corrosion properties. Polym. Chem. 2014, 5, 535–550. [Google Scholar] [CrossRef]
  155. Chen, Y.; Wang, Y.; Zhang, H.-B.; Li, X.; Gui, C.-X.; Yu, Z.-Z. Enhanced electromagnetic interference shielding efficiency of polystyrene/graphene composites with magnetic Fe3O4 nanoparticles. Carbon 2015, 82, 67–76. [Google Scholar] [CrossRef]
  156. Ruan, K.; Guo, Y.; Tang, Y.; Zhang, Y.; Zhang, J.; He, M.; Kong, J.; Gu, J. Improved thermal conductivities in polystyrene nanocomposites by incorporating thermal reduced graphene oxide via electrospinning-hot press technique. Compos. Commun. 2018, 10, 68–72. [Google Scholar] [CrossRef]
  157. Huang, J.; Deng, H.; Song, D.; Xu, H. Electrospun polystyrene/graphene nanofiber film as a novel adsorbent of thin film microextraction for extraction of aldehydes in human exhaled breath condensates. Anal. Chim. Acta 2015, 878, 102–108. [Google Scholar] [CrossRef] [PubMed]
  158. Ponnamma, D.; Nair, S.S.; Parangusan, H.; Hassan, M.K.; Adham, S.; Karim, A.; Al Ali Al-Maadeed, M. White graphene-cobalt oxide hybrid filler reinforced polystyrene nanofibers for selective oil absorption. Polymers 2019, 12, 4. [Google Scholar] [CrossRef] [Green Version]
  159. Wu, N.; Yang, Y.; Wang, C.; Wu, Q.; Pan, F.; Zhang, R.; Liu, J.; Zeng, Z. Ultrathin cellulose nanofiber assisted ambient-pressure-dried, ultralight, mechanically robust, multifunctional MXene aerogels. Adv. Mater. 2023, 35, 2207969. [Google Scholar] [CrossRef] [PubMed]
  160. Afzal, H.M.; Shehzad, F.; Zubair, M.; Bakather, O.Y.; Al-Harthi, M.A. Influence of microwave irradiation on thermal properties of PVA and PVA/graphene nanocomposites. J. Therm. Anal. Calorim. 2020, 139, 353–365. [Google Scholar] [CrossRef]
  161. Li, C.; Vongsvivut, J.; She, X.; Li, Y.; She, F.; Kong, L. New insight into non-isothermal crystallization of PVA–graphene composites. Phys. Chem. Chem. Phys. 2014, 16, 22145–22158. [Google Scholar] [CrossRef] [PubMed]
  162. Liang, J.; Huang, Y.; Zhang, L.; Wang, Y.; Ma, Y.; Guo, T.; Chen, Y. Molecular-level dispersion of graphene into poly(vinyl alcohol) and effective reinforcement of their nanocomposites. Adv. Funct. Mater. 2009, 19, 2297–2302. [Google Scholar] [CrossRef]
  163. Abdah, M.A.A.M.; Zubair, N.A.; Azman, N.H.N.; Sulaiman, Y. Fabrication of PEDOT coated PVA-GO nanofiber for supercapacitor. Mater. Chem. Phys. 2017, 192, 161–169. [Google Scholar] [CrossRef]
  164. Barzegar, F.; Bello, A.; Fabiane, M.; Khamlich, S.; Momodu, D.; Taghizadeh, F.; Dangbegnon, J.; Manyala, N. Preparation and characterization of poly(vinyl alcohol)/graphene nanofibers synthesized by electrospinning. J. Phys. Chem. Solids 2015, 77, 139–145. [Google Scholar] [CrossRef] [Green Version]
  165. Bao, Q.; Zhang, H.; Yang, J.x.; Wang, S.; Tang, D.Y.; Jose, R.; Ramakrishna, S.; Lim, C.T.; Loh, K.P. Graphene-polymer nanofiber membrane for ultrafast photonics. Adv. Funct. Mater. 2010, 20, 782–791. [Google Scholar] [CrossRef]
  166. Verma, R.; Kumar Gupta, S.; Lamba, N.P.; Singh, B.K.; Singh, S.; Bahadur, V.; Chauhan, M.S. Graphene and Graphene Based Nanocomposites for Bio-Medical and Bio-safety Applications. ChemistrySelect 2023, 8, e202204337. [Google Scholar] [CrossRef]
  167. Anwer, A.H.; Ahtesham, A.; Shoeb, M.; Mashkoor, F.; Ansari, M.Z.; Zhu, S.; Jeong, C. State-of-the-art advances in nanocomposite and bio-nanocomposite polymeric materials: A comprehensive review. Adv. Colloid Interface Sci. 2023, 318, 102955. [Google Scholar] [CrossRef]
  168. Bhattacharya, A. A Feasibility Study of the Bioinspired Green Manufacturing of Nanocomposite Materials. In Bioinspired and Green Synthesis of Nanostructures: A Sustainable Approach; Sen, M., Mukherjee, M., Eds.; Wiley, Scrivener Publishing LLC.: Hoboken, NJ, USA, 2023; pp. 231–261. [Google Scholar]
  169. Ibrahim, M.A.; Alhalafi, M.H.; Emam, E.-A.M.; Ibrahim, H.; Mosaad, R.M. A Review of Chitosan and Chitosan Nanofiber: Preparation, Characterization, and Its Potential Applications. Polymers 2023, 15, 2820. [Google Scholar] [CrossRef] [PubMed]
  170. Neibolts, N.; Platnieks, O.; Gaidukovs, S.; Barkane, A.; Thakur, V.; Filipova, I.; Mihai, G.; Zelca, Z.; Yamaguchi, K.; Enachescu, M. Needle-free electrospinning of nanofibrillated cellulose and graphene nanoplatelets based sustainable poly(butylene succinate) nanofibers. Mater. Today Chem. 2020, 17, 100301. [Google Scholar] [CrossRef]
  171. De Faria, A.F.; Perreault, F.; Shaulsky, E.; Arias Chavez, L.H.; Elimelech, M. Antimicrobial electrospun biopolymer nanofiber mats functionalized with graphene oxide-silver nanocomposites. ACS Appl. Mater. Interfaces 2015, 7, 12751–12759. [Google Scholar] [CrossRef] [PubMed]
  172. Yoon, O.J.; Jung, C.Y.; Sohn, I.Y.; Kim, H.J.; Hong, B.; Jhon, M.S.; Lee, N.-E. Nanocomposite nanofibers of poly(D, L-lactic-co-glycolic acid) and graphene oxide nanosheets. Compos. Part A Appl. Sci. Manuf. 2011, 42, 1978–1984. [Google Scholar] [CrossRef]
  173. Jayakumar, A.; Mathew, S.; Radoor, S.; Kim, J.T.; Rhim, J.-W.; Siengchin, S. Recent advances in two-dimensional nanomaterials: Properties, antimicrobial, and drug delivery application of nanocomposites. Mater. Today Chem. 2023, 30, 101492. [Google Scholar] [CrossRef]
  174. Salavagione, H.J.; Gómez-Fatou, M.A.; Shuttleworth, P.S.; Ellis, G.J. New perspectives on graphene/polymer fibers and fabrics for smart textiles: The relevance of the polymer/graphene interphase. Front. Mater. 2018, 5, 18. [Google Scholar] [CrossRef] [Green Version]
  175. Mishra, S.; Singh, S.K.; Singh, A. Polymer Graphene-Based Nanofibers and Their Application for Batteries. In Electrospinning of Graphene, Springer: Berlin/Heidelberg, Germany, 2021; pp. 119–148.
  176. Ahmed, J.; Tabish, T.A.; Zhang, S.; Edirisinghe, M. Porous graphene composite polymer fibres. Polymers 2020, 13, 76. [Google Scholar] [CrossRef]
  177. Amir, A.; Mahalingam, S.; Wu, X.; Porwal, H.; Colombo, P.; Reece, M.; Edirisinghe, M. Graphene nanoplatelets loaded polyurethane and phenolic resin fibres by combination of pressure and gyration. Compos. Sci. Technol. 2016, 129, 173–182. [Google Scholar] [CrossRef]
  178. Matharu, R.K.; Porwal, H.; Ciric, L.; Edirisinghe, M. The effect of graphene-poly(methyl methacrylate) fibres on microbial growth. Interface Focus 2018, 8, 20170058. [Google Scholar] [CrossRef] [Green Version]
  179. Solís Pinargote, N.W.; Smirnov, A.; Peretyagin, N.; Seleznev, A.; Peretyagin, P. Direct ink writing technology (3d printing) of graphene-based ceramic nanocomposites: A review. Nanomaterials 2020, 10, 1300. [Google Scholar] [CrossRef] [PubMed]
  180. Wang, Y.; Zhou, W.; Cao, K.; Hu, X.; Gao, L.; Lu, Y. Architectured graphene and its composites: Manufacturing and structural applications. Compos. Part A Appl. Sci. Manuf. 2021, 140, 106177. [Google Scholar] [CrossRef]
  181. Alarifi, I.M. A Review on Factors Affecting Machinability and Properties of Fiber-Reinforced Polymer Composites. J. Nat. Fibers 2023, 20, 2154304. [Google Scholar] [CrossRef]
  182. Yuan, M.; Teng, Z.; Wang, S.; Xu, Y.; Wu, P.; Zhu, Y.; Wang, C.; Wang, G. Polymeric carbon nitride modified polyacrylonitrile fabrics with efficient self-cleaning and water disinfection under visible light. Chem. Eng. J. 2020, 391, 123506. [Google Scholar] [CrossRef]
  183. Wu, K.; Wang, J.; Liu, D.; Lei, C.; Liu, D.; Lei, W.; Fu, Q. Highly Thermoconductive, Thermostable, and Super-Flexible Film by Engineering 1D Rigid Rod-Like Aramid Nanofiber/2D Boron Nitride Nanosheets. Adv. Mater. 2020, 32, 1906939. [Google Scholar] [CrossRef]
  184. Cheng, J.; Pu, H.; Du, J. A processing method with high efficiency for low density polyethylene nanofibers reinforced by aligned carbon nanotubes via nanolayer coextrusion. Polymer 2017, 111, 222–228. [Google Scholar] [CrossRef]
  185. Qiu, Q.; Chen, S.; Li, Y.; Yang, Y.; Zhang, H.; Quan, Z.; Qin, X.; Wang, R.; Yu, J. Functional nanofibers embedded into textiles for durable antibacterial properties. Chem. Eng. J. 2020, 384, 123241. [Google Scholar] [CrossRef]
  186. Barzoki, P.K.; Latifi, M.; Rezadoust, A. The outstanding effect of nanomat geometry on the interlaminar fracture toughness behavior out of autoclave made glass/phenolic composites under mode-I loading. Eng. Fract. Mech. 2019, 205, 108–119. [Google Scholar] [CrossRef]
  187. Pillai, R.R.; Thomas, V. Plasma Surface Engineering of Natural and Sustainable Polymeric Derivatives and Their Potential Applications. Polymers 2023, 15, 400. [Google Scholar] [CrossRef]
  188. Jung, J.; Sodano, H.A. Aramid nanofiber reinforced rubber compounds for the application of tire tread with high abrasion resistance and fuel saving efficiency. ACS Appl. Polym. Mater. 2020, 2, 4874–4884. [Google Scholar] [CrossRef]
  189. Chung, J.; Kwak, S.-Y. Solvent-assisted heat treatment for enhanced chemical stability and mechanical strength of meta-aramid nanofibers. Eur. Polym. J. 2018, 107, 46–53. [Google Scholar] [CrossRef]
  190. Shi, W.; Hu, B.; Zhang, H.; Li, J.; Yang, J.; Liu, J. Carbon-Encapsulated Iron Oxide Nanoparticles in Self-Supporting Carbon Nanofiber for High-Performance Supercapacitor in Acid Electrolyte with Superior Stability. ACS Appl. Energy Mater. 2020, 3, 12652–12661. [Google Scholar] [CrossRef]
  191. Kumar, T.S.M.; Kumar, K.S.; Rajini, N.; Siengchin, S.; Ayrilmis, N.; Rajulu, A.V. A comprehensive review of electrospun nanofibers: Food and packaging perspective. Compos. Part B Eng. 2019, 175, 107074. [Google Scholar] [CrossRef]
  192. Akampumuza, O.; Gao, H.; Zhang, H.; Wu, D.; Qin, X.H. Raising nanofiber output: The progress, mechanisms, challenges, and reasons for the pursuit. Macromol. Mater. Eng. 2018, 303, 1700269. [Google Scholar] [CrossRef]
  193. Navaratnam, S.; Selvaranjan, K.; Jayasooriya, D.; Rajeev, P.; Sanjayan, J. Applications of natural and synthetic fiber reinforced polymer in infrastructure: A suitability assessment. J. Build. Eng. 2023, 66, 105835. [Google Scholar] [CrossRef]
  194. Adapa, S.K. Prospects of Natural Fiber-Reinforced Polymer Composites for Additive Manufacturing Applications: A Review. JOM 2023, 75, 920–940. [Google Scholar] [CrossRef]
  195. Shin, M.K.; Lee, B.; Kim, S.H.; Lee, J.A.; Spinks, G.M.; Gambhir, S.; Wallace, G.G.; Kozlov, M.E.; Baughman, R.H.; Kim, S.J. Synergistic toughening of composite fibres by self-alignment of reduced graphene oxide and carbon nanotubes. Nat. Commun. 2012, 3, 650. [Google Scholar] [CrossRef] [Green Version]
  196. Kaczor-Urbanowicz, K.E.; Martín Carreras-Presas, C.; Kaczor, T.; Tu, M.; Wei, F.; Garcia-Godoy, F.; Wong, D.T. Emerging technologies for salivaomics in cancer detection. J. Cell. Mol. Med. 2017, 21, 640–647. [Google Scholar] [CrossRef]
  197. Glyva, V.; Kovalenko, V.; Levchenko, L.; Tykhenko, O. Research into protective properties of electromagnetic screens based on the metal-containing nanostructures. Вoстoчнo-Еврoпейский Журнал Передoвых Технoлoгий 2017, 3, 50–56. [Google Scholar] [CrossRef] [Green Version]
  198. Lu, T.; Gu, H.; Hu, Y.; Zhao, T.; Zhu, P.; Sun, R.; Wong, C.-P. Three Dimensional Copper Foam-Filled Elastic Conductive Composites with Simultaneously Enhanced Mechanical, Electrical, Thermal and Electromagnetic Interference (EMI) Shielding Properties. In Proceedings of the 2019 IEEE 69th Electronic Components and Technology Conference (ECTC), Las Vegas, NV, USA, 28–31 May 2019; IEEE: Singapore, 2019; pp. 1916–1920. [Google Scholar]
  199. Qin, R.; Hu, M.; Zhang, N.; Guo, Z.; Yan, Z.; Li, J.; Liu, J.; Shan, G.; Yang, J. Flexible Fabrication of Flexible Electronics: A General Laser Ablation Strategy for Robust Large-Area Copper-Based Electronics. Adv. Electron. Mater. 2019, 5, 1900365. [Google Scholar] [CrossRef]
  200. Shahzad, F.; Alhabeb, M.; Hatter, C.B.; Anasori, B.; Hong, S.M.; Koo, C.M.; Gogotsi, Y. Electromagnetic interference shielding with 2D transition metal carbides (MXenes). Science 2016, 353, 1137–1140. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  201. Ji, H.; Zhao, R.; Zhang, N.; Jin, C.; Lu, X.; Wang, C. Lightweight and flexible electrospun polymer nanofiber/metal nanoparticle hybrid membrane for high-performance electromagnetic interference shielding. NPG Asia Mater. 2018, 10, 749–760. [Google Scholar] [CrossRef]
  202. Kumar, S.; Purohit, R.; Malik, M. Properties and applications of polymer matrix nano composite materials. Mater. Today Proc. 2015, 2, 3704–3711. [Google Scholar] [CrossRef]
  203. Liu, F.; Dong, S.; Zhang, Z.; Dai, X.; Xin, Y.; Wang, X.; Liu, K.; Yuan, Z.; Zhang, J.; Chen, M. Polyaniline/MWCNT Nanocomposite as Sensor for Electroanalytical Determination of Phenol in Oil Field Wastewater. Int. J. Electrochem. Sci. 2019, 14, 9122–9131. [Google Scholar] [CrossRef]
  204. Joseph, N.; Varghese, J.; Sebastian, M.T. A facile formulation and excellent electromagnetic absorption of room temperature curable polyaniline nanofiber based inks. J. Mater. Chem. C 2016, 4, 999–1008. [Google Scholar] [CrossRef]
  205. Lyu, J.; Zhao, X.; Hou, X.; Zhang, Y.; Li, T.; Yan, Y. Electromagnetic interference shielding based on a high strength polyaniline-aramid nanocomposite. Compos. Sci. Technol. 2017, 149, 159–165. [Google Scholar] [CrossRef]
  206. Larcher, D.; Tarascon, J.-M. Towards greener and more sustainable batteries for electrical energy storage. Nat. Chem. 2015, 7, 19. [Google Scholar] [CrossRef]
  207. Li, X.; Wei, B. Supercapacitors based on nanostructured carbon. Nano Energy 2013, 2, 159–173. [Google Scholar] [CrossRef]
  208. Szubzda, B.; Szmaja, A.; Ozimek, M.; Mazurkiewicz, S. Polymer membranes as separators for supercapacitors. Appl. Phys. A 2014, 117, 1801–1809. [Google Scholar] [CrossRef]
  209. Wang, C.; Murugadoss, V.; Kong, J.; He, Z.; Mai, X.; Shao, Q.; Chen, Y.; Guo, L.; Liu, C.; Angaiah, S. Overview of carbon nanostructures and nanocomposites for electromagnetic wave shielding. Carbon 2018, 140, 696–733. [Google Scholar] [CrossRef]
  210. Yang, Z.; Hao, X.; Chen, S.; Ma, Z.; Wang, W.; Wang, C.; Yue, L.; Sun, H.; Shao, Q.; Murugadoss, V. Long-term antibacterial stable reduced graphene oxide nanocomposites loaded with cuprous oxide nanoparticles. J. Colloid Interface Sci. 2019, 533, 13–23. [Google Scholar] [CrossRef] [PubMed]
  211. Jiang, D.; Murugadoss, V.; Wang, Y.; Lin, J.; Ding, T.; Wang, Z.; Shao, Q.; Wang, C.; Liu, H.; Lu, N. Electromagnetic interference shielding polymers and nanocomposites-a review. Polym. Rev. 2019, 59, 280–337. [Google Scholar] [CrossRef]
  212. Shearer, C.J.; Cherevan, A.; Eder, D. Application and future challenges of functional nanocarbon hybrids. Adv. Mater. 2014, 26, 2295–2318. [Google Scholar] [CrossRef] [PubMed]
  213. Rose, A.; Prasad, K.G.; Sakthivel, T.; Gunasekaran, V.; Maiyalagan, T.; Vijayakumar, T. Electrochemical analysis of graphene oxide/polyaniline/polyvinyl alcohol composite nanofibers for supercapacitor applications. Appl. Surf. Sci. 2018, 449, 551–557. [Google Scholar] [CrossRef]
  214. Hofmeister, W.H.; Terekhov, A.Y.; Da Costa, J.L.V. Semipermeable Ultrathin Polymer Membranes. U.S. Patent 16/943,956, 19 November 2020. [Google Scholar]
  215. Mora-Boza, A.; López-Ruiz, E.; López-Donaire, M.L.; Jiménez, G.; Aguilar, M.R.; Marchal, J.A.; Pedraz, J.L.; Vázquez-Lasa, B.; Román, J.S.; Gálvez-Martín, P. Evaluation of Glycerylphytate Crosslinked Semi-and Interpenetrated Polymer Membranes of Hyaluronic Acid and Chitosan for Tissue Engineering. Polymers 2020, 12, 2661. [Google Scholar] [CrossRef]
  216. Morales-Zamudio, L.; Lozano, T.; Caballero-Briones, F.; Zamudio, M.A.; Angeles-San Martin, M.E.; de Lira-Gomez, P.; Martinez-Colunga, G.; Rodriguez-Gonzalez, F.; Neira, G.; Sanchez-Valdes, S. Structure and Mechanical Properties of Graphene Oxide-Reinforced Polycarbonate. Mater. Chem. Phys. 2020, 261, 124180. [Google Scholar] [CrossRef]
  217. Bonetti, L.; Fiorati, A.; Serafini, A.; Masotti, G.; Tana, F.; D'Agostino, A.; Draghi, L.; Altomare, L.; Chiesa, R.; Farè, S. Graphene nanoplatelets composite membranes for thermal comfort enhancement in performance textiles. J. Appl. Polym. Sci. 2020, 138, 49645. [Google Scholar] [CrossRef]
  218. Sierra-Solache, R.; Muro, C.; Maciel, A.; Illescas, J.; Díaz, M.; Carbajal-Franco, G.; Hernández, O. Water recovery from textile wastewater treatment by encapsulated cells of Phanerochaete chrysosporium and ultrafiltration system. Biologia 2020, 75, 1717–1729. [Google Scholar] [CrossRef]
  219. Wang, H.; Zhou, J.; Sun, J.; Wang, Y.; Ma, Y.; Bai, Z.; Zhao, Y.; Zhang, W. Hierarchically Flower-Like Cobalt Oxide@Doped-Sn Carbon Nanofiber with Core-Shell structure as Anodes for Lithium Ion Battery. Int. J. Electrochem. Sci 2020, 15, 9849–9863. [Google Scholar] [CrossRef]
  220. Liang, C.; Zhao, P.; Xie, N.; Wang, S.; Huang, Y.; Lu, L.; Cheng, X. Enhanced comprehensive performance of polymer-CSA cement coating with graphene oxide. Constr. Build. Mater. 2023, 363, 129885. [Google Scholar] [CrossRef]
  221. Zou, L.; Li, Y.; Cao, S.; Ye, B. Gold nanoparticles/polyaniline Langmuir-Blodgett Film modified glassy carbon electrode as voltammetric sensor for detection of epinephrine and uric acid. Talanta 2013, 117, 333–337. [Google Scholar] [CrossRef] [PubMed]
  222. Mo, Y.; Yang, M.; Lu, Z.; Huang, F. Preparation and tribological performance of chemically-modified reduced graphene oxide/polyacrylonitrile composites. Compos. Part A Appl. Sci. Manuf. 2013, 54, 153–158. [Google Scholar] [CrossRef]
  223. Cai, C.; Zhang, Y.; Li, M.; Chen, Y.; Zhang, R.; Wang, X.; Wu, Q.; Chen, T.; Sun, P. Multiple-responsive shape memory polyacrylonitrile/graphene nanocomposites with rapid self-healing and recycling properties. RSC Adv. 2018, 8, 1225–1231. [Google Scholar] [CrossRef]
  224. Armentano, I.; Dottori, M.; Puglia, D.; Kenny, J.M. Effects of carbon nanotubes (CNTs) on the processing and in-vitro degradation of poly(DL-lactide-co-glycolide)/CNT films. J. Mater. Sci. Mater. Med. 2008, 19, 2377–2387. [Google Scholar] [CrossRef]
  225. Wang, M.; Cai, L.; Jin, Q.; Zhang, H.; Fang, S.; Qu, X.; Zhang, Z.; Zhang, Q. One-pot composite synthesis of three-dimensional graphene oxide/poly(vinyl alcohol)/TiO2 microspheres for organic dye removal. Sep. Purif. Technol. 2017, 172, 217–226. [Google Scholar] [CrossRef]
  226. Grant, J.J.; Pillai, S.C.; Hehir, S.; McAfee, M.; Breen, A. Biomedical applications of electrospun graphene oxide. ACS Biomater. Sci. Eng. 2021, 7, 1278–1301. [Google Scholar] [CrossRef]
  227. Lai, L.; Shen, Z.; Lin, J. Conducting Polymer/Graphene-Based Material Composites, and Methods for Preparing the Composites. U.S. Patent 9,734,954, 15 August 2017. [Google Scholar]
  228. Barzegar, F. Synthesis and Characterization of Polymer/Graphene Electrospun Nanofibers; University of Pretoria: Pretoria, South Africa, 2013. [Google Scholar]
  229. Shemshaki, N.S.; Laurencin, C. Graphene-Based Nanofibers for Skeletal Muscle Tissue Regeneration. U.S. Patent 17/352,189, 30 December 2021. [Google Scholar]
  230. Clauss, A.D.; Pan, G.; Wietfeldt, N.R.; Hall, M.C.; Taft, D.D. Polymer-Graphene Nanocomposites. U.S. Patent 9790334B2, 17 October 2017. [Google Scholar]
  231. Chen, L.; Li, J.; Chen, Z.; Gu, Z.; Yan, L.; Zhao, F.; Zhang, A. Toxicological evaluation of graphene-family nanomaterials. J. Nanosci. Nanotechnol. 2020, 20, 1993–2006. [Google Scholar] [CrossRef] [PubMed]
  232. Chowdhury, S.M.; Lalwani, G.; Zhang, K.; Yang, J.Y.; Neville, K.; Sitharaman, B. Cell specific cytotoxicity and uptake of graphene nanoribbons. Biomaterials 2013, 34, 283–293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Jain, V.P.; Chaudhary, S.; Sharma, D.; Dabas, N.; Lalji, R.S.K.; Singh, B.K.; Jaiswar, G. Advanced functionalized nanographene oxide as a biomedical agent for drug delivery and anti-cancerous therapy: A review. Eur. Polym. J. 2021, 142, 110124. [Google Scholar] [CrossRef]
  234. Gupta, K.; Kaushik, A.; Singhal, S. Amelioration of adsorptive efficacy by synergistic assemblage of functionalized graphene oxide with esterified cellulose nanofibers for mitigation of pharmaceutical waste. J. Hazard. Mater. 2022, 424, 127541. [Google Scholar]
  235. Ahmed, M.; Mansour, S.; Al-Wafi, R.; Menazea, A. Composition and design of nanofibrous scaffolds of Mg/Se-hydroxyapatite/graphene oxide@ε-polycaprolactone for wound healing applications. J. Mater. Res. Technol. 2020, 9, 7472–7485. [Google Scholar] [CrossRef]
  236. Jiang, S. Biodegradable Graphene Oxide Biocomposite Fibrous Membrane, Preparation Method and Uses Thereof. U.S. Patent 16/387,071, 22 October 2020. [Google Scholar]
  237. Goenka, S.; Sant, V.; Sant, S. Graphene-based nanomaterials for drug delivery and tissue engineering. J. Control. Release 2014, 173, 75–88. [Google Scholar] [CrossRef] [PubMed]
  238. Othman, F.E.C.; Yusof, N.; Ismail, A.F. Activated-Carbon Nanofibers/Graphene Nanocomposites and Their Adsorption Performance Towards Carbon Dioxide. Chem. Eng. Technol. 2020, 43, 2023–2030. [Google Scholar] [CrossRef]
  239. Boroojeni, F.R.; Mashayekhan, S.; Abbaszadeh, H.-A.; Ansarizadeh, M.; Khoramgah, M.-S.; Movaghar, V.R. Bioinspired Nanofiber Scaffold for Differentiating Bone Marrow-Derived Neural Stem Cells to Oligodendrocyte-Like Cells: Design, Fabrication, and Characterization. Int. J. Nanomed. 2020, 15, 3903–3920. [Google Scholar] [CrossRef]
  240. Pinto, A.M.; Cabral, J.; Tanaka, D.A.P.; Mendes, A.M.; Magalhães, F.D. Effect of incorporation of graphene oxide and graphene nanoplatelets on mechanical and gas permeability properties of poly(lactic acid) films. Polym. Int. 2013, 62, 33–40. [Google Scholar] [CrossRef]
  241. Ferrari, A.C.; Bonaccorso, F.; Fal’Ko, V.; Novoselov, K.S.; Roche, S.; Bøggild, P.; Borini, S.; Koppens, F.H.; Palermo, V.; Pugno, N. Science and technology roadmap for graphene, related two-dimensional crystals, and hybrid systems. Nanoscale 2015, 7, 4598–4810. [Google Scholar] [CrossRef] [Green Version]
  242. Parandeh, S.; Kharaziha, M.; Karimzadeh, F.; Hosseinabadi, F. Triboelectric nanogenerators based on graphene oxide coated nanocomposite fibers for biomedical applications. Nanotechnology 2020, 31, 385402. [Google Scholar] [CrossRef]
  243. Parandeh, S.; Kharaziha, M.; Karimzadeh, F. An eco-friendly triboelectric hybrid nanogenerators based on graphene oxide incorporated polycaprolactone fibers and cellulose paper. Nano Energy 2019, 59, 412–421. [Google Scholar] [CrossRef]
  244. Baek, S.H.; Roh, J.; Park, C.Y.; Kim, M.W.; Shi, R.; Kailasa, S.K.; Park, T.J. Cu-nanoflower decorated gold nanoparticles-graphene oxide nanofiber as electrochemical biosensor for glucose detection. Mater. Sci. Eng. C 2020, 107, 110273. [Google Scholar] [CrossRef]
  245. Wang, Y.; Hao, J.; Huang, Z.; Zheng, G.; Dai, K.; Liu, C.; Shen, C. Flexible electrically resistive-type strain sensors based on reduced graphene oxide-decorated electrospun polymer fibrous mats for human motion monitoring. Carbon 2018, 126, 360–371. [Google Scholar] [CrossRef]
  246. Bahrami, S.; Solouk, A.; Mirzadeh, H.; Seifalian, A.M. Electroconductive polyurethane/graphene nanocomposite for biomedical applications. Compos. Part B Eng. 2019, 168, 421–431. [Google Scholar] [CrossRef]
  247. Orasugh, J.T.; Ray, S.S. Graphene-Based Electrospun Fibrous Materials with Enhanced EMI Shielding: Recent Developments and Future Perspectives. ACS Omega 2022, 7, 33699–33718. [Google Scholar] [CrossRef] [PubMed]
  248. Huang, L.; Li, J.; Li, Y.; He, X.; Yuan, Y. Lightweight and flexible hybrid film based on delicate design of electrospun nanofibers for high-performance electromagnetic interference shielding. Nanoscale 2019, 11, 8616–8625. [Google Scholar] [CrossRef]
  249. Guo, H.; Chen, Y.; Li, Y.; Zhou, W.; Xu, W.; Pang, L.; Fan, X.; Jiang, S. Electrospun fibrous materials and their applications for electromagnetic interference shielding: A review. Compos. Part A Appl. Sci. Manuf. 2021, 143, 106309. [Google Scholar] [CrossRef]
  250. Li, X.-Y.; Yan, Y.; Zhang, B.; Bai, T.-J.; Wang, Z.-Z.; He, T.-S. PAN-derived electrospun nanofibers for supercapacitor applications: Ongoing approaches and challenges. J. Mater. Sci. 2021, 56, 10745–10781. [Google Scholar] [CrossRef]
  251. Ullah, K.; Khan, B.M.; Rashid, A.U.; Oh, W.C. Perspective Chapter: Graphene Based Nanocomposites for Supercapacitor Electrodes. In Updates on Supercapacitors; IntechOpen: London, UK, 2022. [Google Scholar]
  252. Ahmed, F.E.; Lalia, B.S.; Hashaikeh, R. A review on electrospinning for membrane fabrication: Challenges and applications. Desalination 2015, 356, 15–30. [Google Scholar] [CrossRef]
  253. Liu, T.; Lyv, J.; Xu, Y.; Zheng, C.; Liu, Y.; Fu, R.; Liang, L.; Wu, J.; Zhang, Z. Graphene-based woven filter membrane with excellent strength and efficiency for water desalination. Desalination 2022, 533, 115775. [Google Scholar] [CrossRef]
  254. Fang, Q.; Zhou, X.; Deng, W.; Zheng, Z.; Liu, Z. Freestanding bacterial cellulose-graphene oxide composite membranes with high mechanical strength for selective ion permeation. Sci. Rep. 2016, 6, 33185. [Google Scholar] [CrossRef] [Green Version]
  255. Banerjee, A.N. Graphene and its derivatives as biomedical materials: Future prospects and challenges. Interface Focus 2018, 8, 20170056. [Google Scholar] [CrossRef]
  256. Shin, S.R.; Li, Y.-C.; Jang, H.L.; Khoshakhlagh, P.; Akbari, M.; Nasajpour, A.; Zhang, Y.S.; Tamayol, A.; Khademhosseini, A. Graphene-based materials for tissue engineering. Adv. Drug Deliv. Rev. 2016, 105, 255–274. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Biru, E.I.; Necolau, M.I.; Zainea, A.; Iovu, H. Graphene Oxide-Protein-Based Scaffolds for Tissue Engineering: Recent Advances and Applications. Polymers 2022, 14, 1032. [Google Scholar] [CrossRef] [PubMed]
  258. Kaur, H.; Garg, R.; Singh, S.; Jana, A.; Bathula, C.; Kim, H.-S.; Kumbar, S.G.; Mittal, M. Progress and challenges of graphene and its congeners for biomedical applications: Drug delivery, gene delivery, biosensing, bioimaging, and tissue engineering. J. Mol. Liq. 2022, 368, 120703. [Google Scholar] [CrossRef]
Figure 1. Graphene and graphene oxide.
Figure 1. Graphene and graphene oxide.
Jcs 07 00323 g001
Figure 2. Electrospinning setup.
Figure 2. Electrospinning setup.
Jcs 07 00323 g002
Figure 3. Formation of graphene-wrapped polyaniline (PANI) nanofiber [134]. Reproduced with permission from Elsevier.
Figure 3. Formation of graphene-wrapped polyaniline (PANI) nanofiber [134]. Reproduced with permission from Elsevier.
Jcs 07 00323 g003
Figure 4. Transmission electron microscopy (TEM) image of graphene-wrapped polyaniline nanofibers [134]. Reproduced with permission from Elsevier.
Figure 4. Transmission electron microscopy (TEM) image of graphene-wrapped polyaniline nanofibers [134]. Reproduced with permission from Elsevier.
Jcs 07 00323 g004
Figure 5. Specific capacitance of polyaniline (PANI) and graphene-wrapped polyaniline nanofibers (Gra100-wrap-PANI-NF) vs. varying current densities [134]. Reproduced with permission from Elsevier.
Figure 5. Specific capacitance of polyaniline (PANI) and graphene-wrapped polyaniline nanofibers (Gra100-wrap-PANI-NF) vs. varying current densities [134]. Reproduced with permission from Elsevier.
Jcs 07 00323 g005
Figure 6. TEM micrographs of nanofibrous mats with (A) 5 wt.% graphene and (B) 15 wt.% graphene [148]. Reproduced with permission from Elsevier.
Figure 6. TEM micrographs of nanofibrous mats with (A) 5 wt.% graphene and (B) 15 wt.% graphene [148]. Reproduced with permission from Elsevier.
Jcs 07 00323 g006
Figure 7. Schematic representation of bulk versus nanofibrous material and cross-sectional area of nanofibrous material [148]. Reproduced with permission from Elsevier.
Figure 7. Schematic representation of bulk versus nanofibrous material and cross-sectional area of nanofibrous material [148]. Reproduced with permission from Elsevier.
Jcs 07 00323 g007
Figure 8. (A) TEM bright field image of nylon 6 nanofiber; (B) STEM Z-contrast image showing two nanocomposite nanofibers containing GOFT layers [150]. STEM = scanning transmission electron microscopy; GOFT = nitroxide-functionalized graphene oxide layers. Reproduced with permission from Elsevier.
Figure 8. (A) TEM bright field image of nylon 6 nanofiber; (B) STEM Z-contrast image showing two nanocomposite nanofibers containing GOFT layers [150]. STEM = scanning transmission electron microscopy; GOFT = nitroxide-functionalized graphene oxide layers. Reproduced with permission from Elsevier.
Jcs 07 00323 g008
Figure 9. (A) The pilot-scale extrusion line and (B) winding system used to produce graphene-modified polyamide 6 yarns [151]. Reproduced with permission from Elsevier.
Figure 9. (A) The pilot-scale extrusion line and (B) winding system used to produce graphene-modified polyamide 6 yarns [151]. Reproduced with permission from Elsevier.
Jcs 07 00323 g009
Figure 10. Contact angle of nanocomposite nanofibers before and after gamma irradiation [158]. PS = polystyrene; PS/Co-O = PS/Co3O4 at 1 wt.%; PS/hB = polystyrene/hexagonal boron nitride at 1 wt.%; PS/hBCo = PS/hexagonal boron nitride/Co3O4 at 1 wt.%. Reproduced with permission from MDPI.
Figure 10. Contact angle of nanocomposite nanofibers before and after gamma irradiation [158]. PS = polystyrene; PS/Co-O = PS/Co3O4 at 1 wt.%; PS/hB = polystyrene/hexagonal boron nitride at 1 wt.%; PS/hBCo = PS/hexagonal boron nitride/Co3O4 at 1 wt.%. Reproduced with permission from MDPI.
Jcs 07 00323 g010
Figure 11. Biomedical applications of graphene-based nanofibers [226]. Reproduced with permission from ACS.
Figure 11. Biomedical applications of graphene-based nanofibers [226]. Reproduced with permission from ACS.
Jcs 07 00323 g011
Table 1. Techniques, process parameters, and properties of polymer and polymer nanocomposite nanofibers.
Table 1. Techniques, process parameters, and properties of polymer and polymer nanocomposite nanofibers.
MethodsElectrospinningSolution BlowingTemplate SynthesisPhase Inversion Freeze DryingSTEP
Techniques
Nanofiber diameter range40 nm to 2 μm40 nm to several
μm
40 nm to a few hundred nanometers50 nm to
1 μm
50 nm–
1 μm
50 nm to several
μm
Production rate (injection rate)5 μL/min20 μL/min---15–100 μL/min
Influencing parametersVoltages,
viscosity,
solution feeding rate, distance
Nozzle geometry, viscosity, solution feeding rate, gas pressureTemplate shape, template pore sizeSolvent properties,
polymer concentration
Freezing rate, solvent features,
polymer concentration
Polymer type, molecular weight,
solvent
properties, substrate,
speed
Voltage requirement10–40 kVNO~30 V
(for electropolymerization)
NoNoNo
IndustrializationYesYesNoNoNoNo
Possibility of producing aligned nanofibersYesYesYesYesYesYes
Possibility for melt spinningYesYesYesNoNoNo
Possibility for spinning from highly concentrated polymer
solutions
YesYesNoNoNoYes
Production of core/shell nanofibersYesYesHollowYesYesNo
Production of polymer/composite fiberPolyamide; polystyrene; polyaniline;
polyamide/carbon nanotube;
polystyrene/carbon nanotube; polyaniline/single-walled carbon nanotube nanofibers; polyaniline/zinc oxide; the nylon 6,6/zinc oxide; polyaniline/titania
Poly(vinyl alcohol); poly (lactic acid);
poly(vinyl alcohol)/zinc oxide;
poly(lactic acid)/titanium dioxide
Polypyrrole; polyaniline;
polypyrrole/silica;
Poly(vinyl fluoride); aramid;
poly(vinyl fluoride)/bentonite; aramid/zeolite
Poly(vinyl alcohol); poly(vinyl alcohol)/cellulose;Polymer/titania;
polymer/titania/alumina; hydroxyapatite/chitosan
Refs.[31,32,33,34,35,36][37,38][39,40][41,42][43,44,45][46,47,48]
Table 2. Properties and fabrication techniques for polymer/graphene nanocomposite nanofibers.
Table 2. Properties and fabrication techniques for polymer/graphene nanocomposite nanofibers.
Polymer/
Nanocomposite
TechniqueDiameter/
Size
Solvent/
Concentration
Flow Rate Voltage RequirementPhysical PropertiesRef.
Polyaniline/poly (methyl methacrylate)/amino-functionalized graphene.Electrospinning 35–133 nmDimethyl formamide0.3 mL/h18–20 kVThermal stability[97]
Poly(ε-caprolactone)/
graphene oxide
Electrospinning201–264 nmGlacial acetic acid;
1.5 w/v%
1 mL/h12 kVIncrement in tensile stress by 189%[98]
Poly(ε-caprolactone)/
graphene
Electrospinning121–154 nmDichloromethane/methanol; 10–12 wt.%0.8–1 mL/h15–17 kVYoung’s modulus tensile strength of 3771 MPa and 56.08 MPa, respectively[99]
Poly(ε-caprolactone)/
reduced graphene oxide
Electrospinning100–130 nm Glacial acetic acid; 1.5 w/v%1 mL/h12 kVTensile strength increase by 304 % [100]
Polyamide/grapheneElectrospinning76–338 nmHexafluoroisopropanol;
0.005–0.01 wt.%
0.05 mL/h8–10 kVIncrease in tensile strength,
Young’s modulus fracture, toughness by 56%, 113%, and 250%, respectively
[101]
Polyethylene/grapheneDrawing-Xylene; 0.1 wt.%Draw ratio 30–70-Thermal conductivity
≈ 75 W m−1 K−1 ρ−1
[102]
Polypyrrole/grapheneTemplate method 80–100 nmNaOH and acidic solutions--Specific capacitance
466 Fg−1;
energy density
165.7 Wh/Kg
[103]
Polypyrrole/grapheneWet spinning method40 µmAcidic solution; ethanol: water;
~10 mg/ml
--Tensile strength 364.3 MPa;
specific capacitance
334 mF cm−2
[104]
Aramid/grapheneCo-axial spinning~8 nmDimethyl sulfoxide2, 4, 6 mL/h-Increase in ultimate tensile stress by 700%[105]
Polyethylene terephthalate/grapheneDry-jet wet spinning -Methanol;
3 mg/mL
Air gap
~3 cm; pressure 25.0 psi
-Percolation threshold
0.2 S/cm
[106]
Table 3. Electrospinning parameters used for nanofiber production [148]. Reproduced with permission from Elsevier.
Table 3. Electrospinning parameters used for nanofiber production [148]. Reproduced with permission from Elsevier.
Electrospun SolutionGraphene Content
ppm (wt.%)
Flow Rate/mL/hElectric Potential/kVDistance/cmHumidity (%)
NY-0G00.3220.017.038–40
NY-0.05G500 (0.05)0.2320.818.043–45
NY-0.1G1000 (0.1)0.2521.018.054–45
NY-1.5G15,000 (1.5)0.1718.018.039–31
NY-2G20,000 (2.0)0.5016.715.033–35
NY-5G50,000 (5.0)0.7017.315.028–30
NY-8G80,000 (8.0)0.3015.120.024–26
NY-15G150,000 (15.0)0.5020.015.031–33
Table 4. Mechanical properties of the nanocomposites. [158]. PS = polystyrene; PS/Co-O = PS/Co3O4 at 1 wt.%; PS/hB = polystyrene/hexagonal boron nitride at 1 wt.%; PS/hBCo = PS/hexagonal boron nitride/Co3O4 at 1 wt.%. Reproduced with permission from MDPI.
Table 4. Mechanical properties of the nanocomposites. [158]. PS = polystyrene; PS/Co-O = PS/Co3O4 at 1 wt.%; PS/hB = polystyrene/hexagonal boron nitride at 1 wt.%; PS/hBCo = PS/hexagonal boron nitride/Co3O4 at 1 wt.%. Reproduced with permission from MDPI.
SampleTensile Strength (MPa)Young’s Modulus (MPa)
Non-irradiatedPS
PS/Co-O
PS/hBN PS/hBCo-O
28.54 ± 1.1
30.47 ± 2.3
35.44 ± 2.5
48.24 ± 2.6
72.50 ± 6.44
76.33 ± 2.21
77.65 ± 4.05
98.15 ± 4.79
IrradiatedPS
PS/Co-O
PS/hBN PS/hBCo-O
30.87 ± 1.3
31.05 ± 2.2
38.25 ± 1.7
52.54 ± 2.1
73.30 ± 9.01
77.88 ± 8.75
82.03 ± 2.45
110.35 ± 4.55
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

Kausar, A.; Ahmad, I.; Zhao, T.; Aldaghri, O.; Ibnaouf, K.H.; Eisa, M.H. Nanocomposite Nanofibers of Graphene—Fundamentals and Systematic Developments. J. Compos. Sci. 2023, 7, 323. https://doi.org/10.3390/jcs7080323

AMA Style

Kausar A, Ahmad I, Zhao T, Aldaghri O, Ibnaouf KH, Eisa MH. Nanocomposite Nanofibers of Graphene—Fundamentals and Systematic Developments. Journal of Composites Science. 2023; 7(8):323. https://doi.org/10.3390/jcs7080323

Chicago/Turabian Style

Kausar, Ayesha, Ishaq Ahmad, Tingkai Zhao, Osamah Aldaghri, Khalid H. Ibnaouf, and M. H. Eisa. 2023. "Nanocomposite Nanofibers of Graphene—Fundamentals and Systematic Developments" Journal of Composites Science 7, no. 8: 323. https://doi.org/10.3390/jcs7080323

Article Metrics

Back to TopTop