Next Article in Journal
Diatomite-like KFeS2 for Use in High-Performance Electrodes for Energy Storage and Oxygen Evolution
Next Article in Special Issue
A Novel Photo-Thermal-Electric Conversion System with an Integrated Support Material
Previous Article in Journal
Electrochemistry of Carbon Materials: Progress in Raman Spectroscopy, Optical Absorption Spectroscopy, and Applications
Previous Article in Special Issue
Characterization of Magnetron Sputtered BiTe-Based Thermoelectric Thin Films
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fabrication of Flexible Poly(m-aminophenol)/Vanadium Pentoxide/Graphene Ternary Nanocomposite Film as a Positive Electrode for Solid-State Asymmetric Supercapacitors

1
Department of Chemical Engineering, College of Engineering, King Faisal University, P.O. Box 380, Al-Ahsa 31982, Saudi Arabia
2
Department of Chemical Engineering, Birla Institute of Technology, Ranchi 835215, India
*
Authors to whom correspondence should be addressed.
Submission received: 17 January 2023 / Revised: 30 January 2023 / Accepted: 1 February 2023 / Published: 6 February 2023
(This article belongs to the Special Issue Nanostructured Materials for Energy Applications)

Abstract

:
In this study, poly(m-aminophenol) (PmAP) has been investigated as a multi-functional conductive supercapacitor binder to replace the conventional non-conductive binder, namely, poly(vinylene difluoride) (PVDF). The kye benefits of using PmAP are that it is easily soluble in common organic solvent and has good film-forming properties, and also its chemical functionalities can be involved in pseudocapacitive reactions to boost the capacitance performance of the electrode. A new ternary nanocomposite film based on vanadium pentoxide (V2O5), amino-functionalized graphene (amino-FG) and PmAP was fabricated via hydrothermal growth of V2O5 nanoparticles on graphene surfaces and then blending with PmAP/DMSO and solution casting. The electrochemical performances of V2O5/amino-FG/PmAP nanocomposite were evaluated in two different electrolytes, such as KCl and Li2SO4, and compared with those of V2O5/amino-FG nanocomposite with PVDF binder. The cyclic voltametric (CV) results of the V2O5/amino-FG/PmAP nanocomposite exhibited strong pseudocapacitive responses from the V2O5 and PmAP phases, while the faradaic redox reactions on the V2O5/amino-FG/PVDF electrode were suppressed by the inferior conductivity of the PVDF. The V2O5/amino-FG/PmAP electrode delivered a 5-fold greater specific capacitance than the V2O5/amino-FG/PVDF electrode. Solid-state asymmetric supercapacitors (ASCs) were assembled with V2O5/amino-FG/PmAP film as a positive electrode, and their electrochemical properties were examined in both KCl and Li2SO4 electrolytes. Although the KCl electrolyte-based ASC has greater specific capacitance, the Li2SO4 electrolyte-based ASC delivers a higher energy density of 51.6 Wh/kg and superior cycling stability.

Graphical Abstract

1. Introduction

Our daily lives are increasingly dependent on energy, resulting in an increase in global energy demand. The over-exploitation of fossil fuels and their irreversible adverse effects on the environment and the climate led to the promotion of clean energy resources. Electrochemical energy conversion and storage (EECS) technologies, including rechargeable batteries, fuel cells, and supercapacitors, can replace coal-based power supplies. As a next-generation energy storage device, supercapacitors have been proven to be the most reliable and effective alternative to conventional batteries since batteries demonstrate slow power delivery and uptake, a short cycle-life, and a less eco-friendly profile, while supercapacitors offer high power density, a fast delivery rate, high safety performance, and a long-life cycle (up to 104) [1,2]. For portable and flexible electronics, such as tablets, laptops, and foldable phones, flexible solid-state supercapacitors can serve as a real-time power backup system and micropower source instead of batteries for these flexible smart electronics [3,4,5,6]. Yet, state-of-the-art supercapacitors offer a limited energy storage density a few orders of magnitude less than conventional batteries, which prevents their widespread use in the energy-storage industries [7,8]. Generally, there are two mechanisms by which supercapacitors store electrical energy: electronic double layer capacitance (EDLC) and pseudocapacitance. For EDLC, the charge storage process occurs through the formation of electrical double-layers by means of the adsorption/desorption of electrolyte ions onto the active electrode surface, while the pseudocapacitive materials store charges via the faradaic charge storage process [5,6].
Generally, EDLC-electrodes are made of porous carbon, including activated carbon, mesoporous carbon, and graphitic nanocarbons [9,10]. Meanwhile, graphene—a thin, 2D, carbon sheet—has received huge interest as an advanced EDLC active material due to its superior electrical conductivity, exceptionally high charge carrier mobility (~15,000 cm2/V/s), large theoretical specific surface area (up to 2675 m2/g), superior theoretical gravimetric capacitance (500 F/g), strong mechanical strength (~1 TPa), and low cost [11,12]. However, the low capacitance of graphene-based electrode materials still remains a fundamental challenge for their widespread commercial use in supercapacitors. The strong agglomeration tendency of graphene sheets significantly reduces their effective surface area and ion-diffusion rate [13]. Several studies have demonstrated that carbon materials combined with pseudocapacitive materials, such as metal oxides [14,15,16], metal hydroxide [17,18], metal sulfides [19,20], metal-organic framework [21,22,23], and conducting polymers [24,25], can improve the capacitance of carbon-based supercapacitors because of their synergistic effects.
Among the several transition-metal-oxides-based pseudocapacitive electrode materials, vanadium pentoxide has gained considerable attention, either as a pseudocapacitive electrode material for supercapacitors or as a high-capacity cathode material for Li-/Na-ion batteries [26,27,28], because of its layered-crystalline structure, rich redox chemistry (V+3, V+4, and V+5), ultra-high theoretical capacitance of 2120 F/g over a voltage window of 1.0 V, and earth-abundant and non-toxic nature [29,30]. In spite of these attractive advantages, the low electrical conductivity and inferior structural stability (in liquid electrolytes) of V2O5 inhibit the charge–discharge rate and hamper the long-term cycling performance [31]. There has been a lot of research done to enhance the ion-diffusion rate of V2O5 and subsequently improve its capacitive performance. The synthesis of hybrid materials, by incorporating highly conductive carbon materials or conducting polymers into V2O5, is the most common approach to overcome those barriers. The specific capacitance of V2O5 was found to enhanced by 257% [32], 520% [33], 129% [34], 28% [35], and 64% [36] by inclusion of exfoliated graphene, reduced graphene oxide, multi-wall carbon nanotube, N-rich mesoporous carbon, and carbon sphere, respectively. Several researchers have successfully improved the specific capacitance of V2O5 by hybridization with polyaniline (PANI) [37,38], polypyrrole (PPy) [39,40], polythiophene (PEDOT) [41], and polydiphenylamine [42].
In recent years, the hybridization of an EDLC material with two different pseudocapacitive materials was found to be a highly effective strategy in producing an electrode with enhanced electrochemical properties for supercapacitors. The combination of carbon material (EDLC) with pseudocapacitive materials can not only enhance the capacitance of carbon material, but also boost the chemical stability of the pseudocapacitive components during electrochemical cycling. This is not only because of the synergistic effects between these electrode materials, but also as a result of the enhanced effective surface area, which in turn reduces the diffusion path length. Many research groups have explored various types of ternary hybrid materials, such as graphene/NiO/PPy [43], graphene oxide/Fe2O3/PPy [44], rGO/MnFe2O4/PPy [45], graphene/SnO2/PEDOT [46], graphene/NiO/poly(aniline-co-m-aminophenol) [47], RGO/ZrO2/PPy [48], and graphene/MoO3/PPy [49], for supercapacitor applications. Most of the conventional synthetic approaches usually produce powdery active materials, and consequently, nonconductive polymeric additives/conductive agents are required to add to the active materials for fabrication of supercapacitor electrodes, which inescapably reduces the content of active materials, as well as affects their electrical conductivity [50,51]. Conducting polymers as an active component could compensate for the lower electrical conductivity of the electrode material, as well as provide additional pseudocapacitance induced by their functional groups in the backbone chain. However, the inferior solubility of the conducting polymers in most of the organic solvents makes it difficult to prepare free-standing film by solution casting, which limits their use as an active material in supercapacitors. PmAP, a derivative of polyaniline, possesses good electrical conductivity (10−5 S/cm), as well as high solubility in organic solvents. The high solubility of PmAP in a common organic medium can allow preparing flexible nanocomposite electrodes via the solution casting approach. In addition, the multiple functional groups in the PmAP backbone chain can act as active redox sites to provide pseudocapacitance and surface wettability, which were observed in our earlier work [23]. These advantages can make PmAP an attractive alternative polymeric binder for commonly used non-conductive binders, namely, PVDV, polytetrafluoroethylene (PTFE), polyvinyl alcohol (PVOH), etc. To the best of our knowledge, the soluble conducting polymers are rarely explored as conductive binders for fabrication of supercapacitor electrodes.
In the present investigation, we have synthesized V2O5/amino-FG/PmAP nanocomposite film by a two-step process: (i) V2O5 nanoparticles were initially grown on the surface of amino-FG nanosheets via the hydrothermal method, and (ii) as-prepared V2O5/amino-FG nanocomposites were dispersed in a PmAP solution and then cast on a glass substrate to form ternary nanocomposite film upon evaporation of the solvent. As far as we know, we are exploring PmAP as binder-cum-pseudocapacitive electrode material for the first time for the possible replacement of commercial PVDF binder. The electrochemical performances of the hydrophilic V2O5/amino-FG/PmAP nanocomposite film were evaluated and compared in different electrolytes, such as KCl and Li2SO4. The rich redox chemistry of V2O5 and the conjugated PmAP backbone can offer pseudocapacitance to boost the specific capacitance of the nanocomposite electrode, while graphene as a conductive component would provide high power density. This bendable nanocomposite film could potentially be used to fabricate flexible supercapacitors for wearable electronic devices. Solid-state ASC cells with different electrolytes were constructed by using ternary nanocomposite film as the positive electrode and activated carbon cloth as the negative electrode.

2. Experimental Section

2.1. Materials

The aminated graphene (batch: 900551), m-aminophenol, V2O5 powder, potassium chloride (KCl), lithium sulfate (Li2SO4), nafion solution (5 wt%), ammonium persulfate (APS), and dimethyl sulfoxide (DMSO) were procured from Sigma Aldrich, Mumbai, India. Poly(vinylene difluoride), sulfuric acid (H2SO4), N-methyl pyrrolidone (NMP), ethanol, ammonium hydroxide (NH4OH), ethanol (EtOH), and sodium hydroxide (NaOH) were purchased from Alfa Aesar, Mumbai, India. Activated carbon cloth (Spectracarb 2225) was obtained from Engineered Fibers Technology, LLC, Shelton, CT, USA.

2.2. Synthesis of V2O5/Amino-FG Nanocomposites

The V2O5/amino-FG nanocomposites were synthesized by the hydrothermal approach. The first step in this synthesis was to disperse 250 mg V2O5 in 25 mL of deionized (DI) water by mechanical stirring at 70 °C. In the next step, 10 mL of H2O2 (30%) was added drop-by-drop to the V2O5 solution under constant stirring, while keeping the solution at 70 °C. The yellow color dispersed solution became orange color solution after completion of the exothermic reaction. Separately, 500 mg of amino-FG powder was dispersed in 25 mL EtOH by ultrasonication before being transferred to a Teflon-lined autoclave. The V2O5 sol solution was then added to the graphene solution under constant starring for 2 h and then heated at 120 °C for 24 h. The resulting black precipitate was filtered, repeatedly washed with ethanol and water, and finally dried at 60 °C for 24 h.

2.3. Preparation of V2O5/Amino-FG/Poly(m-aminophenol) Nanocomposite Film

The poly(m-aminophenol) was synthesized through the chemical oxidation polymerization of m-aminophenol monomer by APS in 1 M aqueous NaOH solution at 4 °C [51]. To prepare nanocomposite film, 250 mg PmAP was dissolved in 10 mL DMSO by mechanical stirring at 25 °C. Afterward, 1 g V2O5/amino-FG powder was added to the PmAP solution and the stirred mixture overnight. The homogeneous solution was then poured on a cleaned petri dish and the DMSO evaporated by heating at 70 °C. The solvent-cast nanocomposite film was thoroughly washed with water to ensure the complete removal of residual solvent molecules and dried at 70 °C overnight. The thickness of the as-prepared V2O5/amino-FG/PmAP nanocomposite film was determined to be ~420 μm.

2.4. Characterizations

IR-absorption and Raman spectra of V2O5/amino-FG and V2O5/amino-FG/PmAP nanocomposites were obtained by IRPrestige-21 spectrophotometer and a Renishaw Raman system 3000, respectively. X-ray diffraction (XRD) patterns of as-prepared nanocomposites were recorded on a Rigaku Smart Lab with Cu Kα radiation (40 kV, 40 mA). The elemental compositions of the nanocomposite samples were determined by XPS (MultiLab 2000) and TEM-EDS analyses. The microstructure of the as-prepared nanocomposites were studied by field-emission scanning electron microscopy (FE-SEM, Hitachi, S-4700, Tokyo, Japan) and a transmission electron microscopy (TEM, JEOL JEM 2100-2100F).

2.5. Electrochemical Measurements

Electrochemical measurements with 1M KCl and 1M Li2SO4 were conducted in an Autolab electrochemical workstation. For the CV tests, a three-electrode cell was constituted with a working electrode, a reference electrode (Ag/AgCl), and a counter electrode (platinum wire). The nanocomposite film was initially cut into a circular sample (diameter = 5 mm) and then fixed to the glassy carbon electrode (GCE), using Nafion conductive solution to serve as the working electrode. To use the V2O5/amino-FG nanocomposite as electrode material, V2O5/amino-FG nanocomposite powder (80 wt%), super P (5 wt%), and PVDF binder (20 wt%) were mixed in NMP using stirring for 3 h. The working electrode was prepared by coating the slurry on the surface of GCE, followed by drying at 70 °C. PVDF was selected as a binder because of several advantages, such as better capacitive properties, excellent stability, and lower cost than other polymeric binders like PTFE and PVOH. A two-electrode cell was used to carried out the galvanostatic charge–discharge (GCD), where two square pieces of nanocomposite films (dimension = 1 × 1 cm2) were employed as working electrodes. The total mass of the hybrid electrode is 3.417 mg. EIS measurements were performed in a frequency region of 0.01–100,000 Hz, using an AC voltage of 5 mV. The solid-state ASC cells were assembled by using V2O5/amino-FG/PmAP nanocomposite film as a positive electrode and activated carbon cloth as a negative electrode, and their electrochemical performances were evaluated in KCl and Li2SO4 electrolytes. The polyvinyl alcohol (PVA) + electrolyte membrane was used as electrolyte-cum-separator [52]. The mass ratio of V2O5/amino-FG/PmAP electrode and activated carbon cloth was calculated to be 0.5823. The specific capacitances (Csp) of the nanocomposite electrodes were determined from the equation [5]:
C s p = I × Δ t m × Δ V  
where I, Δt, m and ΔV denote the current, discharge time, mass of the active electrode material, and voltage window, respectively. Similarly, the energy density (ED) and power density (PD) were calculated with the following equations [5]:
E D = C s p × Δ V 2 2 × 3.6  
P D = 3600 × E D Δ t  

3. Results and Discussion

The FTIR spectra of binary V2O5/amino-FG nanocomposite and ternary V2O5/amino-FG/PmAP nanocomposite film are shown in Figure 1a. The FTIR spectrum of V2O5/amino-FG nanocomposite displays few characteristic absorption peaks at 1616, 1008, 771, and 484 cm−1, which refer to the bending vibration of the crystalline water molecule, stretching modes of terminal V=O, V-O-V asymmetrical stretching, and stretching vibration of the O-(V)3 bond, respectively [53]. The peak at 1222 cm−1 is related to the vibrational modes of the C-O group in the aminated graphene. The broad band in the 3600–2700 cm−1 regime denotes O-H stretching. In contrast, the ternary V2O5/amino-FG/PmAP nanocomposite film displayed some additional IR peaks, along with the absorption bands related to V2O5. A couple of characteristic peaks at 2920 and 2343 cm−1 correspond to stretching of C=C bonds in benzene rings and stretching modes of C=C=N and C=C=O bonds [54]. The C=C stretching mode of the quinoid form appeared at 1611 cm−1, while that of the benzenoid form is observed at 1577 cm−1. IR spectrum reveals two absorption peaks at 1611 and 1577 cm−1 for the quinoid and benzenoid ring stretching vibrations, respectively [55]. The peak at 1144 cm−1 refers to C-O-C linkages in the backbone chain [56]. Figure 1b presents the Raman spectra of V2O5/amino-FG nanocomposite and ternary V2O5/amino-FG/PmAP nanocomposite films. For the V2O5/amino-FG nanocomposite, the characteristic peaks in the range frequency range of 200–1000 cm−1 are associated with the crystalline V2O5 phase. The G- and D-band are located at 1571 and 1355 cm−1, which refer to the vibration of the sp2-carbon lattice in the graphite plane and disordered carbon lattice (sp3-carbon), respectively. The peaks at 991, 688, and 526 cm−1 are related to stretching modes of terminal V=O, V2-O, and V3-O bonds, respectively [33]. The peak located at 404 cm−1 is due to the V=O bending vibration. As shown in the Raman spectra of V2O5/amino-FG/PmAP film, in addition to D and G bands, five characteristic bands related to the PmAP phase are observed at 1614, 1486, 1264, 1228, and 1163 cm−1, which refer to the stretching absorption of the C=C bond, the stretching mode of the C=N bond, the stretching of the C-N bond in quinoid/polaron forms, the in-plane bending of the C-H bond in benzenoid, and in quinoid units, respectively. The XRD patterns of as-synthesized nanocomposites are shown in Figure 1c. The V2O5/amino-FG nanocomposite demonstrates characteristic diffraction peaks at 2θ = 6.8°, 26.3°, 31.7°, 34.6°, 46.7°, 50.5°, and 61.4°, which are associated with the (001), (110), (220), (210), (310), (006), and (601) crystal planes of orthorhombic V2O5 phase (JCPDS file No. 40-1296), respectively. The d-spacing of V2O5.xH2O was calculated to 0.65 nm from the diffraction angle of the (001) plane, using the Bragg equation. In contrast to V2O5/amino-FG nanocomposite, the diffraction pattern of V2O5/amino-FG/PmAP nanocomposite shows additional peaks at 2θ = 9.1°, 15.2°, 17.5°, 20.4°, and 23.2° corresponding to the (001), (111), (201), (020), and (200) crystal planes of the PmAP phase, respectively, along with diffraction peaks from the V2O5 phase. The diffraction peak of the (001) plane of V2O5 for V2O5/amino-FG/PmAP nanocomposite was shifted to the lower 2θ value of 6.39° compared to that of V2O5/amino-FG nanocomposite. The d-spacing of V2O5.xH2O was calculated to 0.691 nm from the diffraction angle of the (001) plane. In the ternary nanocomposite, the enhanced d-spacing of V2O5.xH2O might be attributed the intercalation of PmAP chains between V2O5 layers, which results in a widening of interplanar spacing of the (001) plane of orthorhombic V2O5 phase [57].
The surface chemical compositions of V2O5/amino-FG and V2O5/amino-FG/PmAP nanocomposite film was evaluated by XPS. The survey spectrum of V2O5/amino-FG/PmAP nanocomposite film displays four distinct signals for C, N, O, and V elements, indicating their presence in the nanocomposite, while the XPS spectrum of V2O5/amino-FG powder indicates the existence of C, O, and V elements (Figure 2a). The high-resolution V2p spectrum exhibits V2p3/2 and V2p1/2 peaks at 513.8 eV and 521.3 eV, respectively (Figure 2b). The O1s spectrum could be deconvoluted into four peaks by Gaussian curve fitting. The peaks at 529.9 eV and 530.7 eV are assigned to the V-O-V and V-O-H bonds in the V2O5 lattice (Figure 2c) [58]. The small broad peak centered at 533.2 eV is associated with crystal water in V2O5, indicating the hydrous nature of the vanadium oxide on graphene sheets. The XPS peak at 534.3 eV is attributed to the phenolic oxygen present in PmAP backbone [59,60]. The N1s spectrum can be deconvoluted into two peaks at 399.1 eV and 401.7 eV (Figure 2d), which are related to C-N and C=N+- groups in the PmAP backbone.
The microstructures of V2O5/amino-FG and V2O5/amino-FG/PmAP nanocomposite films were studied by FE-SEM analysis. The SEM images of V2O5/amino-FG clearly exhibit that the graphene nanosheets are well wrapped in or covered with tiny V2O5 nanoparticles. The high-resolution SEM image (Figure 3c) indicates the in situ hydrothermal growth of highly dispersed V2O5 nanoparticles on the surface of graphene nanosheets. It can be believed that the V2O5 nanoparticles were sandwiched between graphene sheets. SEM images of the fracture surface of the V2O5/amino-FG/PmAP nanocomposite film is demonstrated in Figure 3d–f. The microstructure of the fracture surface reveals the structural compactness and the good dispersion of V2O5-coated graphene nanosheets in the PmAP matrix. Numerous protuberances on the graphene surfaces were observed in high-resolution SEM images, indicating the presence of V2O5 nanoparticles. This self-assembled interconnected morphology of the nanocomposite would be beneficial for charge transfer between them. The TEM image of the V2O5/amino-FG/PmAP nanocomposite further demonstrates a consistent dispersion of V2O5 nanoparticles (average size 10 nm) on graphene surfaces (Figure 3g). The individual V2O5 nanoparticle clearly exhibits crystal lattice with a lattice spacing of 0.61 nm, corresponding to the d-spacing of (001) lattice planes of the orthorhombic V2O5 phase. The SAED pattern of the nanocomposite displays well-defined rings with a clear diffraction spot array, indicating the crystalline characteristics of the V2O5. The diffraction spots in the SEAD pattern can be indexed to the (001), (110), (220), (210), and (310) crystal planes in orthorhombic V2O5, which is well in agreement with the XRD results. The EDX elemental mapping (Figure 3f) clearly indicates the uniform distribution of carbon, nitrogen, oxygen and vanadium, indicating homogeneous dispersion of V2O5 nanoparticles throughout the nanocomposite.
The electrochemical properties of as-prepared V2O5/amino-FG/PmAP nanocomposite films were evaluated in two different electrolytes and compared with those of V2O5/amino-FG/PVDF nanocomposite. The CV profiles of V2O5/amino-FG/PmAP nanocomposite-based electrode in 1 M KCl and 1 M Li2SO4 electrolytes over a potential range of 0–1.0 V at 25 mV/s are demonstrated in Figure 4a. The CV curves of the V2O5/amino-FG/PmAP electrode obtained in both the electrolytes consist of two pairs of broad oxidation-reduction peaks and are quasi-rectangular in shape. These results suggest that the EDLC from graphene nanosheets combined with the pseudocapacitance from V2O5 and PmAP makes up the overall capacitance of the nanocomposite electrode. In contrast to the V2O5/amino-FG/PmAP electrode, the V2O5/amino-FG/PVDF electrode displays quasi-rectangular CV profiles without redox peaks in either electrolyte, which might be associated with the low conductivity of PVDF and high charge transfer resistance of the V2O5/amino-FG electrode with PVDF binder compared to V2O5/amino-FG/PmAP (Figure 5f). The CV profiles of V2O5/amino-FG/PmAP electrode in KCl electrolyte at different scan rates are displayed in Figure 4b. In the CV curves, two anodic peaks are observed at 0.33 V and 0.72 V, as well as two cathodic peaks at 0.61 V and 0.21 V. When the scan rate is increased from 5 to 100 mV/s, there is a gradual increase in redox peak currents, as well as an increase in the integrated CV area. This could be ascribed to the increased interfacial ionic mobility at higher scan rates. The large number of functional redox-active sites in V2O5/amino-FG/PmAP nanocomposite makes it suitable for fast reversible faradaic reactions. When the scan rate was increased, there were noticeable shifts in the cathodic and anodic peak potentials towards more positive and negative values, respectively. At higher scan rates, there might be inadequate diffusion of electrolyte ions into nanocomposite matrix as well as restricted charge transfer interactions between electrolytes and electrode surface. Figure 4c illustrates the plots of the redox peak currents against scan rate in logarithmic scale. The slopes of the anodic and cathodic fitting curves are 0.722 and 0.843, respectively. Theoretically, slope value of 0.5 and 1.0 indicates faradaic diffusion process and surface-controlled (pseudocapacitance) process, respectively. Since the slope values of V2O5/amino-FG/PmAP electrode lie between 0.5 and 1.0, both diffusion and pseudocapacitive behavior affected the electrochemical properties of nanocomposite electrodes. However, their capacitance behavior is dominated by pseudocapacitance process rather than diffusion process, probably because of high ionic mobility of the K+ ions (0.68 × 10−3 cm2/s/V). The electrochemical insertion/desertion of K+ ions into the hybrid electrode occurs via the possible reaction: V2O5 + nK+ + ne ↔ KnV2O5, where n is mole fraction of K+ ions. The GCD curves of V2O5/amino-FG/PmAP electrode in 1 M KCl at different current densities are illustrated in Figure 4d. The non-linearity of the GCD profiles could be attributed to the occurrence of faradaic redox reactions, where V2O5 and PmAP participated as electroactive materials. It can be noticed that charge–discharge rate remains uncharged when the current density is increased. The coulombic efficiency (η) of the V2O5/amino-FG/PmAP electrode was calculated to be around 92%, indicating its excellent cycling stability and rate capability. The specific capacitances of V2O5/amino-FG/PmAP and V2O5/amino-FG/PVDF electrodes as function of current density are illustrated in Figure 4e. The V2O5/amino-FG/PmAP electrode exhibited a remarkable Csp of 1438.4 F/g at 0.5 A/g, which is more than four times as low as that of V2O5/amino-FG/PVDF electrode (292.9 F/g). The greater Csp values of the V2O5/amino-FG/PmAP electrode could be ascribed to the higher level of pseudocapacitive contribution from V2O5 and PmAP components, which comprise of numerous redox active functionalities to enhance electron-transport ability of the nanocomposite electrode. In contrast, the non-conductive PVDF binder significantly suppressed electron exchange between active material and electrolyte during cycling. As the current density raised to 5 F/g, the Csp value decreased from 1438.4 F/g to 760.3 F/g. The V2O5/amino-FG/PmAP electrode can retain 76% and 53% of its capacitance after 4-fold and 10-fold increases in current density, respectively, indicating its good rate capability. The V2O5/amino-FG/PmAP electrode can deliver a maximum areal capacitance of 719.2 F/g and a minimum areal capacitance of 380.1 F/g at current densities of 0.5 and 5 A/g, respectively. In Figure 4f, a gradual decrease in the Csp of V2O5/amino-FG/PmAP electrode with increasing the cycle number is observed. However, the V2O5/amino-FG/PmAP electrode can retain its specific capacitance of 92% after 7000 cycles, which implies its impressive cycling life in KCl electrolyte.
The electrochemical performance of the V2O5/amino-FG/PmAP electrode was evaluated in Li2SO4 electrolyte and compared with KOH electrolyte. Figure 5a illustrates the CV profiles of V2O5/amino-FG/PmAP electrode in 1 M Li2SO4 at different scan rates. In the CV curves, a pair of anodic peaks can be found at 0.27 V and 0.38 V with a pair of cathodic peaks at 0.16 V and 0.31 V in the reverse scan. It is believed that the pseudocapacitive properties are caused by the redox interaction between V2O5/PmAP and Li2SO4. The current level of the CV curves gradually rises as scan rate increases. The nanocomposite electrode in 1 M Li2SO4 exhibits a lower current value and a smaller CV area than those in 1 M KCl. This might be due to the higher ionic radii of the hydrated SO42− ions (3.79 Å) and subsequently its lower ionic mobility than those of the hydrated Cl ions (3.32 Å). Hence, the insertion and desertion of SO42− ions into and out of the electrode matrix are much more difficult during cycling even though the ionic conductivity of sulfate ions is superior to that of chloride ions [58]. It can be concluded that the rate of diffusion of electrolyte ions plays an influential role in achieving better performance of the supercapacitors in addition to their ionic conductivity. Figure 5b shows the plots of faradaic peak currents against the scan rate at the logarithmic scale. The anodic and cathodic slope values are determined to be 0.725 and 0.683, respectively, which indicates that the capacitance performance is controlled by both diffusion and pseudocapacitance processes. The GCD profiles of V2O5/amino-FG/PmAP in 1 M Li2SO4 are demonstrated in Figure 5c. With increasing current density, discharge time gets shorter. The specific capacitances of V2O5/amino-FG/PmAP and V2O5/amino-FG/PVDF electrodes as a function of current density are presented in Figure 5d. The nanocomposites containing PmAP binder delivered significantly higher specific capacitance compared to those containing PVDF binder. The V2O5/amino-FG/PmAP electrode achieved a highest specific capacitance of 1191.7 F/g at 0.5 A/g in 1 M Li2SO4, while the capacitance value decreased to 510.3 F/g at 5 A/g. The V2O5/amino-FG/PmAP electrode showed 42% capacitance retention after 10-fold increases in current density, which is lower than that observed in KCl electrolyte (53% retention). The lower specific capacitance in Li2SO4 electrolyte could be ascribed to the higher volume of hydrated SO42− ions than Cl, which sluggish the transport rate of SO42− ions and consequently enhance the internal resistance. The V2O5/amino-FG/PmAP electrode displayed 97% capacitance retention over 7000 GCD cycles in 1 M Li2SO4. The superior cycling stability in Li2SO4 is probably due to the strong solvation of Li+ and SO42− ions in the aqueous medium due to their high charge density. The Nyquist plots of V2O5/amino-FG/PmAP and V2O5/amino-FG/PVDF electrodes in different electrolytes are illustrated in Figure 5f. The impedance curves of V2O5/amino-FG/PmAP displayed a semi-circular arc in a high frequency regime together with a Warburg line in a low frequency regime. The V2O5/amino-FG/PmAP electrode reveals a smaller semicircle arc in KCl electrolyte, indicating its lower charge transfer resistance (Rct) than in Li2SO4 electrolyte. The Rct values of V2O5/amino-FG/PmAP electrode in KCl and Li2SO4 electrolytes are 46.2 Ω and 67.5 Ω, respectively. In contrast to Li2SO4, the smaller volume of hydrated K+/Cl ions is responsible for the lower Rct value in KCl. Furthermore, in aqueous 1 M KCl and 1 M Li2SO4, the nanocomposite electrode displays a solution resistance of 1.35 and 4.34, respectively. These results correlated well with the charge–discharge performances of the V2O5/amino-FG/PmAP electrode in these electrolytes. Furthermore, a steeper Warburg slope in KCl medium implies easier ionic movement across the aqueous KCl solution as well as the electrode–electrolyte interface. In contrast to the V2O5/amino-FG/PmAP electrode, the V2O5/amino-FG/PmAP electrodes possesses a smaller semicircle arc, which indicates their higher charge transfer resistance because PVDF is a poor electrical conductor than PmAP. The Rct values of V2O5/amino-FG/PmAP and V2O5/amino-FG/PVDF electrodes in 1 M KCl are 46.2 Ω and 257.9 Ω, respectively, which are well in agreement with their electrochemical performances.
Figure 6a demonstrates the CV profiles of solid-state V2O5/amino-FG/PmAP//activated carbon cloth asymmetric cell in 1 M KCl over various operating voltage windows at 25 mV/s. The CV results indicate that the as-assembled ASC cell is capable of stable operation up to as high as 1.4 V, which is further confirmed by the charge–discharge results (Figure 6c). As a result of extending the working voltage window, the integral CV area increases without changing their shape significantly, demonstrating a stable cyclic performance with a reasonable rate capacity of the ASC device when applied to a higher voltage window. The CV profiles are almost rectangular and have broad faradaic peaks, indicating that the ASC cell stored energy via both EDLC and pseudocapacitance processes. The CV curves of the ASC at different scan rates over a cell potential limit of 0–1.4 V are demonstrated in Figure 6b. With an increasing scan rate, the integral CV area increased gradually, indicating good rate capability. Figure 6d shows the GCD curves of as-assembled ASC cell at varying current densities, while Figure 6e illustrates the variation of specific capacitances as a function of current density. The maximum Csp value for the ASC could reach to 251.4 F/g at 0.5 A/g in KCl electrolyte. The ASC in KCl electrolyte also displayed reasonable capacitance retention, i.e., 54%, when current density was increased to 5 A/g. The ASC retained 93.2% capacitance after 4000 cycles in KCl electrolyte, implying its good cycling performance. The electrochemical performances of ASC cell in Li2SO4 electrolyte were evaluated and compared with those obtained with KCl electrolyte. Figure 7a illustrates the CV curves of ASC cells in Li2SO4 electrolyte at varying voltage ranges at 25 mV/s. As shown in Figure 7a, the ASC cell in Li2SO4 electrolyte can stably operate up to a working potential window of 0–1.7 V. Moreover, the CV results corroborated well by GCD results (Figure 7c). The excellent electrochemical stability in neutral Li2SO4 electrolyte over a wide potential window can be attributed to the low parasitic side reactions [5]. The asymmetrical rectangular shape of the CV curves indicates that both faradaic and non-faradaic mechanisms are involved in the charge storage process. The CV profiles of Li2SO4-based ASC device at different scan rates are shown in Figure 7b. There is no change in the shape of the CV curves as the scan rate increases, but the total CV area gradually increases. It can be noticed that at any given scan rate, the current level of the Li2SO4-based ASC is relatively lower than that found for KCl-based ASC, indicating better capacitive response of the ASC cell in KCl electrolyte. The GCD curves of ASC cell in Li2SO4 at different current densities are illustrated in Figure 7d. The asymmetric feature of the GCD profiles is further implies the occurrence of faradaic redox reaction during charge–discharge process. The specific capacitances of ASC cell in Li2SO4 electrolyte at different current densities are presented in Figure 7e. A maximum specific capacitance of 216.4 F/g at a current density of 0.5 A/g was achieved for the ASC cell in Li2SO4 electrolyte. The capacitance was reduced to 125.3 F/g at a current density of 5 A/g, which correspond to a 58% capacitance retention. The ASC with Li2SO4 electrolyte possesses outstanding cycling stability with 96.7% capacity retention after 4000 cycles compared to 93.2% retention with KCl electrolyte. The coulombic efficiency (η) of ASC in Li2SO4 electrolyte was calculated to be 83%, which indicates its excellent rate capability and electrochemical stability.
The impedance analyses of the ASC cell in both KCl and Li2SO4 electrolytes were conducted, and the resulting Nyquist plots are demonstrated in Figure 8a. The Nyquist plots of ASC cells exhibit a semicircular arc with Warburg straight line in the high and low frequency regions, respectively. The Rct values of ASC cells in KCl and Li2SO4 electrolytes are 51.7 Ω and 90.1 Ω, respectively. In contrast to Li2SO4 based ASC cell, the ASC cell based on KOH electrolyte exhibit steeper Warburg line, indicating faster ion transport. This could be attributed to the smaller volume of hydrated K+ and Cl ions than those of Li+ and SO42− ions. The Ragone plots of as-assembled ASC devices are demonstrated in Figure 8b. ASC device with Li2SO4 electrolyte delivered a highest energy density of 51.6 Wh/kg at a power density of 500 W kg−1, which significantly higher than those of KCl-based ASC cell (48.1 Wh/kg). The higher energy density of Li2SO4-based ASC cell is associated with the wider potential window (0–1.7 V). Furthermore, the ASC cell with Li2SO4 electrolyte retained the energy density of 29.6 Wh/ kg even at high power density of 6655 W/kg, while in contrast, the KCl-based ASC device can deliver the energy density of 26 Wh/kg at a power density of 4607 W/kg. Table 1 and Figure 8b illustrate a comparison of energy density at corresponding power density between present V2O5/amino-FG/PmAP//activated carbon ASC cell and previously reported ASC cells consisting of various ternary metal oxide/graphene/conducting polymer nanocomposites, which indicates better electrochemical performances of the present ASC cell compared to reported devices. Inset of Figure 8b displays the lightening of LED bulb using KCl-based ASC cell. The CV analysis of the as-assembled ASC cell in KCl electrolyte was carried out at different bending angles to evaluate its bending stability. Figure 8c illustrates that both shape and current level of the CV curves remain unchanged after bending the ASC cell from 0° to 135°, indicating its excellent bending stability.

4. Conclusions

In this study, we have demonstrated that PmAP can be potentially used as a conductive binder-cum-pseudocapacitive material instead of PVDF binder to fabricate self-standing supercapacitor electrodes with better capacitance performances. The electrochemical properties of as-prepared V2O5/amino-FG/PmAP nanocomposites were compared with those of V2O5/amino-FG nanocomposites with PVDF binder. The V2O5/amino-FG electrode with PmAP binder had a 5-fold higher specific capacitance in 1 M KCl electrolyte than the electrode with PVDF binder, which could be attributed to the higher electrical conductivity and pseudocapacitive properties of PmAP. EIS results revealed that the V2O5/amino-FG/PmAP electrode has much lower charge transfer resistance (46.2 Ω) than the electrode using PVDF as a binder (257.9 Ω). The type of electrolyte plays a crucial role in improving the capacitance of the V2O5/amino-FG/PmAP electrode. The specific capacitance is as high as 1438.4 F/g in 1 M KCl electrolyte and 1191.7 F/g in 1 M Li2SO4 electrolyte at 0.5 A/g. An as-assembled solid-state V2O5/amino-FG/PmAP//activated carbon cloth asymmetric cell exhibited higher energy density and better cycling stability with KCl electrolyte than with Li2SO4 electrolyte. The Li2SO4-based ASC cell can deliver maximum energy and power densities of 51.6 Wh/kg and 6655 W/kg, respectively, which are substantially higher than those of ASC devices constructed with various previously reported ternary composites as the positive electrode. We believe that PmAP could be used as a cheap binder-cum-pseudocapacitive material for next-generation supercapacitors with high energy density, compared with the PVDF binder currently used in commercial supercapacitors.

Author Contributions

Conceptualization, S.S.H. and A.C.; methodology, B.D.; validation, S.S.H., A.C., B.D. and S.S.A.; formal analysis, A.C. and S.S.H.; investigation, B.D. and S.S.A.; resources, S.S.H.; data curation, B.D. and S.S.A.; writing—original draft preparation, A.C.; writing—review and editing, S.S.H. and A.C.; project administration, S.S.H. and A.C.; funding acquisition, S.S.H. and S.S.A. 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, Saudi Arabia, for funding this research work through project number INST016.

Data Availability Statement

Not applicable.

Acknowledgments

We are thankful to the operators of FE-SEM (A. K. Singh), BET (S. K. Pradhan), FTIR (S. K. Saw), and XRD (N. Yadav) at the Central Instrumental Facility, Birla, Institute of Technology, Mesra, Ranchi, India, for their technical assistance executing the experimental analysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Vimuna, V.M.; Athira, A.R.; Dinesh Babu, K.V.; Xavier, T.S. Simultaneous stirring and microwave assisted synthesis of nanoflakes MnO2/rGO composite electrode material for symmetric supercapacitor with enhanced electrochemical performance. Diam. Relat. Mater. 2020, 110, 108129. [Google Scholar] [CrossRef]
  2. Ahmad, M.W.; Anand, S.; Dey, B.; Fatima, A.; Yang, D.J.; Choudhury, A. N/P/O/S Heteroatom-Doped Porous Carbon Nanofiber Mats Derived from a Polyacrylonitrile/l-Cysteine/P2O5 Precursor for Flexible Electrochemical Supercapacitors. ACS Appl. Energy Mater. 2021, 4, 12177–12190. [Google Scholar] [CrossRef]
  3. Zan, G.; Wu, T.; Zhu, F.; He, P.; Cheng, Y.; Chai, S.; Wang, Y.; Huang, X.; Zhang, W.; Wan, Y.; et al. A biomimetic conductive super-foldable material. Matter 2021, 4, 3232–3247. [Google Scholar] [CrossRef]
  4. Zan, G.; Wu, T.; Dong, W.; Zhou, J.; Tu, T.; Xu, R.; Chen, Y.; Wang, Y.; Wu, Q. Two-level biomimetic designs enable intelligent stress dispersion for super-foldable C/NiS nanofiber free-standing electrode. Adv. Fiber Mater. 2022, 4, 1177–1190. [Google Scholar] [CrossRef]
  5. Ahmad, M.W.; Choudhury, A.; Dey, B.; Anand, S.; Saidi, A.K.A.A.; Lee, G.H.; Yang, D.J. Three-dimensional core-shell niobium-metal organic framework@ carbon nanofiber mat as a binder-free positive electrode for asymmetric supercapacitor. J. Energy Storage 2022, 55, 105484. [Google Scholar] [CrossRef]
  6. Ahmad, M.W.; Anand, S.; Dey, B.; Yang, D.J.; Choudhury, A. Asymmetric supercapacitors based on porous MnMoS4 nanosheets-anchored carbon nanofiber and N, S-doped carbon nanofiber electrodes. J. Alloys Compd. 2022, 906, 164271. [Google Scholar] [CrossRef]
  7. Wang, H.; Yi, H.; Chen, X.; Wang, X. Asymmetric supercapacitors based on nano-architectured nickel oxide/graphene foam and hierarchical porous nitrogen-doped carbon nanotubes with ultrahigh-rate performance. J. Mater. Chem. A 2014, 2, 3223. [Google Scholar] [CrossRef]
  8. Paul, A.; Ghosh, S.; Kolya, H.; Kang, C.-W.; Murmu, N.C.; Kuila, T. High performance asymmetric supercapacitor device based on lanthanum doped nickel-tin oxide/reduced graphene oxide composite. J. Energy Storage 2022, 55, 105526. [Google Scholar] [CrossRef]
  9. Wei, L.; Deng, W.; Li, S.; Wu, Z.; Cai, J.; Luo, J. Sandwich-like chitosan porous carbon Spheres/MXene composite with high specific capacitance and rate performance for supercapacitors. J. Biores. Bioprod. 2022, 7, 63–72. [Google Scholar] [CrossRef]
  10. Zheng, S.; Zhang, J.; Deng, H.; Du, Y.; Shi, X. Chitin derived nitrogen-doped porous carbons with ultrahigh specific surface area and tailored hierarchical porosity for high performance supercapacitors. J. Biores. Bioprod. 2021, 6, 142–151. [Google Scholar] [CrossRef]
  11. Bokhari, S.W.; Siddique, A.H.; Sherrell, P.C.; Yue, X.; Karumbaiah, K.M.; Wei, S.; Ellis, A.V.; Gao, W. Advances in graphene-based supercapacitor electrodes. Energy Rep. 2020, 6, 2768–2784. [Google Scholar] [CrossRef]
  12. Horn, M.; Gupta, B.; MacLeod, J.; Liu, J.; Motta, N. Graphene-based supercapacitor electrodes: Addressing challenges in mechanisms and materials. Cur. Opin. Green Sustain. Chem. 2019, 17, 42–48. [Google Scholar] [CrossRef]
  13. Yin, S.; Niu, Z.; Chen, X. Assembly of graphene sheets into 3D macroscopic structures. Small 2012, 8, 2458–2463. [Google Scholar] [CrossRef] [PubMed]
  14. Ke, Q.; Wang, J. Graphene-based materials for supercapacitor electrodes–A review. J. Materiomics. 2016, 2, 37–54. [Google Scholar] [CrossRef]
  15. Velasco, A.; Ryu, Y.K.; Boscá, A.; Ladrón-de-Guevara, A.; Hunt, E.; Zuo, J.; Pedrós, J.; Calle, F.; Martinez, J. Recent trends in graphene supercapacitors: From large area to microsupercapacitors. Sust. Energy Fuels 2021, 5, 1235–1254. [Google Scholar] [CrossRef]
  16. Ding, X.; Liu, R.; Zhao, J.; Hu, J.; Wu, J.; Zhang, C.; Lin, J. In situ formation of Co3O4 nanocrystals embedded in laser-induced graphene foam for high-energy flexible micro-supercapacitors. Dalton Trans. 2022, 51, 2846–2854. [Google Scholar] [CrossRef]
  17. Low, W.H.; Khiew, P.S.; Lim, S.S.; Siong, C.W.; Ezeigwe, E.R. Recent development of mixed transition metal oxide and graphene/mixed transition metal oxide-based hybrid nanostructures for advanced supercapacitors. J. Alloys Compd. 2019, 775, 1324–1356. [Google Scholar] [CrossRef]
  18. Zhou, Y.; Liu, M.; Yang, H.; Liu, Q.; Li, W.; Yu, C.-M. Porous graphene oxide nanosheets warped by Ni(OH)2 platelets as an efficient binder-free electrode material for supercapacitors. Synt. Met. 2020, 267, 116452. [Google Scholar] [CrossRef]
  19. Lo, H.-J.; Huang, M.-C.; Lai, Y.-H.; Chen, H.-Y. Towards bi-functional all-solid-state supercapacitor based on nickel hydroxide-reduced graphene oxide composite electrodes. Mater. Chem. Phys. 2021, 262, 124306. [Google Scholar] [CrossRef]
  20. Balu, R.; Dakshanamoorthy, A. One pot preparation of tin sulfide decorated graphene nanocomposite for high performance supercapacitor applications. Inorg. Chem. Commun. 2022, 136, 109148. [Google Scholar] [CrossRef]
  21. Marand, N.A.; Masoudpanah, S.M.; Alamolhoda, S.; Bafghi, M.S. Solution combustion synthesis of nickel sulfide/reduced graphene oxide composite powders as electrode materials for high-performance supercapacitors. J. Energy Storage 2021, 39, 102637. [Google Scholar] [CrossRef]
  22. Barua, A.; Mehra, P.; Paul, A. Understanding Integrated Graphene–MOF Nanostructures as Binder-and Additive-Free High-Performance Supercapacitors at Commercial Scale Mass Loading. ACS Appl. Energy Mater. 2021, 4, 14249–14259. [Google Scholar] [CrossRef]
  23. Ngo, T.V.; Moussa, M.; Tung, T.T.; Coghlan, C.; Losic, D. Hybridization of MOFs and graphene: A new strategy for the synthesis of porous 3D carbon composites for high performing supercapacitors. Electrochim. Acta 2020, 329, 135104. [Google Scholar]
  24. Choudhury, A.; Dey, B.; Mahapatra, S.S.; Kim, D.-W.; Yang, K.-S.; Yang, D.-J. Flexible and freestanding supercapacitor based on nanostructured poly(m-aminophenol)/carbon nanofiber hybrid mats with high energy and power densities. Nanotechnology 2018, 29, 165401. [Google Scholar] [CrossRef] [PubMed]
  25. Liu, H.; Wang, Y.; Gou, X.; Qi, T.; Yang, J.; Ding, Y. Three-dimensional graphene/polyaniline composite material for high-performance supercapacitor applications. Mater. Sci. Eng. B 2013, 178, 293–298. [Google Scholar] [CrossRef]
  26. Zhang, R.; Pang, H. Application of graphene-metal/conductive polymer-based composites in supercapacitors. J. Energy Storage 2021, 33, 102037. [Google Scholar] [CrossRef]
  27. Leistenschneider, D.; Abedi, Z.; Ivey, D.G.; Chen, W. Coating of Low-Cost Asphaltenes-Derived Carbon Fibers with V2O5 for Supercapacitor Application. Energy Fuels 2022, 36, 3328–3338. [Google Scholar] [CrossRef]
  28. Panigrahi, K.; Howli, P.; Chattopadhyay, K.K. 3D network of V2O5 for flexible symmetric supercapacitor. Electrochim. Acta 2020, 337, 135701. [Google Scholar] [CrossRef]
  29. Zhu, Y.; Yang, M.; Huang, Q.; Wang, D.; Yu, R.; Wang, J.; Zheng, Z.; Wang, D. V2O5 Textile Cathodes with High Capacity and Stability for Flexible Lithium-Ion Batteries. Adv. Mater. 2020, 32, 1906205. [Google Scholar] [CrossRef] [PubMed]
  30. Qian, A.; Pang, Y.; Wang, G.; Hao, Y.; Liu, Y.; Shi, H.; Chung, C.H.; Du, Z.; Cheng, F. Pseudocapacitive Charge Storage in MXene–V2O5 for Asymmetric Flexible Energy Storage Devices. ACS Appl. Mater. Interf. 2020, 12, 54791–54797. [Google Scholar] [CrossRef]
  31. Chen, D.; Li, J.; Wu, Q. Review of V2O5-based nanomaterials as electrode for supercapacitor. J. Nanopart. Res. 2019, 21, 201. [Google Scholar] [CrossRef]
  32. Xu, X.; Xiong, F.; Meng, J.; Wang, X.; Niu, C.; An, Q.; Mai, L. Vanadium-based nanomaterials: A promising family for emerging metal-ion batteries. Adv. Funct. Mater. 2020, 30, 1904398. [Google Scholar] [CrossRef]
  33. Choudhury, A.; Bonso, J.S.; Wunch, M.K.; Yang, S.; Ferraris, J.P.; Yang, D.J. In-situ synthesis of vanadium pentoxide nanofibre/exfoliated graphene nanohybrid and its supercapacitor applications. J. Power Source 2015, 287, 283–290. [Google Scholar] [CrossRef]
  34. Raj, T.N.V.; Hoskeri, P.A.; Hamzad, S.; Anantha, M.S.; Joseph, C.M.; Muralidhara, H.B.; Kumar, K.Y.; Alharti, F.A.; Jeong, B.-H.; Raghu, M.S. Moringa Oleifera leaf extract mediated synthesis of reduced graphene oxide-vanadium pentoxide nanocomposite for enhanced specific capacitance in supercapacitors. Inorg. Chem. Comm. 2022, 142, 109648. [Google Scholar]
  35. Wang, Q.; Zou, Y.; Xiang, C.; Chu, H.; Zhang, H.; Xu, F.; Sun, L.; Tang, C. High-performance supercapacitor based on V2O5/carbon nanotubes-super activated carbon ternary composite. Ceram. Int. 2016, 42, 12129–12135. [Google Scholar] [CrossRef]
  36. Saravanakumar, B.; Purushothaman, K.K.; Muralidharan, G. V2O5/nitrogen enriched mesoporous carbon spheres nanocomposite as supercapacitor electrode. Microporous Mesoporous Mater. 2018, 258, 83–94. [Google Scholar] [CrossRef]
  37. Dinesh, A.; Ramadas, A.; Anantha, M.S.; Umesh, M.K.; Venkatesh, K.; Kundu, M.; Muralidhara, H.B.; Kumar, K.Y. Synergistic behavior of vanadium pentoxide-carbon sphere electrocatalyst towards iron-based redox flow battery and supercapacitor applications. J. Energy Storage 2022, 55, 105487. [Google Scholar] [CrossRef]
  38. Roy, A.; Ray, A.; Sadhukhan, P.; Saha, S.; Das, S. Morphological behaviour, electronic bond formation and electrochemical performance study of V2O5-polyaniline composite and its application in asymmetric supercapacitor. Mater. Res. Bull. 2018, 107, 379–390. [Google Scholar] [CrossRef]
  39. Bai, M.-H.; Liu, T.-Y.; Luan, F.; Li, Y.; Liu, X.-X. Electrodeposition of vanadium oxide–polyaniline composite nanowire electrodes for high energy density supercapacitors. J. Mater. Chem. A 2014, 2, 10882–10888. [Google Scholar] [CrossRef]
  40. Sun, X.; Li, Q.; Mao, Y. Understanding the influence of polypyrrole coating over V2O5 nanofibers on electrochemical properties. Electrochim. Acta 2015, 174, 563–573. [Google Scholar] [CrossRef]
  41. Qu, Q.; Zhu, Y.; Gao, X.; Wu, Y. Core–Shell Structure of Polypyrrole Grown on V2O5 Nanoribbon as High-Performance Anode Material for Supercapacitors. Adv. Energy Mater. 2012, 2, 950–955. [Google Scholar] [CrossRef]
  42. Bi, W.; Wu, Y.; Liu, C.; Wang, J.; Du, Y.; Gao, G.; Wu, G.; Cao, G. Gradient Oxygen Vacancies in V2O5/PEDOT Nanocables for High-Performance Supercapacitors. ACS Appl. Energy Mater. 2019, 2, 668–677. [Google Scholar] [CrossRef]
  43. Khatoon, H.; Iqbal, S.; Ahmad, S. Influence of medium on structure, morphology and electrochemical properties of polydiphenylamine/vanadium pentoxide composite. SN Appl. Sci. 2019, 1, 261. [Google Scholar] [CrossRef]
  44. Golkhatmi, S.Z.; Sedghi, A.; Miankushki, H.N.; Khalaj, M. Structural properties and supercapacitive performance evaluation of the nickel oxide/graphene/polypyrrole hybrid ternary nanocomposite in aqueous and organic electrolytes. Energy 2021, 214, 118950. [Google Scholar] [CrossRef]
  45. Vigneshwaran, J.; Abraham, S.; Muniyandi, B.; Prasankumar, T.; Li, J.-T.; Jose, S. Fe2O3 decorated graphene oxide/polypyrrole matrix for high energy density flexible supercapacitor. Surf. Interf. 2021, 27, 101572. [Google Scholar] [CrossRef]
  46. Ishaq, S.; Moussa, M.; Kanwal, F.; Ayub, R.; Van, T.N.; Azhar, U.; Losic, D. One step strategy for reduced graphene oxide/cobalt-iron oxide/polypyrrole nanocomposite preparation for high performance supercapacitor electrodes. Electrochim. Acta 2022, 427, 140883. [Google Scholar] [CrossRef]
  47. Wang, W.; Lei, W.; Yao, T.; Xia, X.; Huang, W.; Hao, Q.; Wang, X. One-pot synthesis of graphene/SnO2/PEDOT ternary electrode material for supercapacitors. Electrochim. Acta 2013, 108, 118–126. [Google Scholar] [CrossRef]
  48. Ji, X.; Xu, Q.; Zhou, T.; Wang, X.; Xu, H.; Yao, X.; Lan, W.; Kong, Y. Synthesis of poly (aniline-co-m-aminophenol)/graphene/NiO nanocomposite and its application in supercapacitors. Synt. Met. 2016, 211, 14–18. [Google Scholar] [CrossRef]
  49. Alves, A.P.P.; Koizumi, R.; Samanta, A.; Machado, L.D.; Singh, A.K.; Galvao, D.S.; Silva, G.G.; Tiwary, C.S.; Ajayan, P.M. One-step electrodeposited 3D-ternary composite of zirconia nanoparticles, rGO and polypyrrole with enhanced supercapacitor performance. Nano Energy 2017, 31, 225–232. [Google Scholar] [CrossRef]
  50. Hsu, F.-H.; Wu, T.-M. Poypyrrole/molybdenum trioxide/graphene nanoribbon ternary nanocomposite with enhanced capacitive performance as an electrode for supercapacitor. J. Solid State Electrochem. 2016, 20, 691–698. [Google Scholar] [CrossRef]
  51. Yu, Y.; Xu, A.; Zhang, Y.; Li, W.; Qin, Y. Evaporation-induced hydrated graphene/polyaniline/carbon cloth integration towards high mass loading supercapacitor electrodes. Chem. Eng. J. 2022, 445, 136727. [Google Scholar] [CrossRef]
  52. Verma, S.K.; Kar, P.; Yang, D.J.; Choudhury, A. Poly(m-aminophenol)/functionalized multi-walled carbon nanotube nanocomposite-based alcohol sensors. Sens. Actuators B 2015, 219, 199–208. [Google Scholar] [CrossRef]
  53. Ahmad, M.W.; Anand, S.; Shalini, K.; Ul-Islum, M.; Yang, D.-J.; Choudhury, A. MnMoO4 nanorods-encapsulated carbon nanofibers hybrid mat as binder-free electrode for flexible asymmetric supercapacitors. Mater. Sci. Semicond. Process. 2021, 136, 106176. [Google Scholar] [CrossRef]
  54. Jadhav, C.D.; Pandit, B.; Karade, S.S.; Sankapal, B.R.; Chavan, P.G. Enhanced field emission properties of V2O5/MWCNTs nanocomposite. Appl. Phys. A 2018, 124, 794. [Google Scholar] [CrossRef]
  55. Chen, C.; Sun, C.; Gao, Y. Electrosynthesis of poly (aniline-co-p-aminophenol) having electrochemical properties in a wide pH range. Electrochim. Acta 2008, 53, 3021–3028. [Google Scholar] [CrossRef]
  56. Kar, P.; Pradhan, N.C.; Adhikari, B. Induced doping by sodium ion in poly(m-aminophenol) through the functional groups. Synth. Met. 2010, 160, 13–14. [Google Scholar] [CrossRef]
  57. Zoromba, M.S.; Abdel-Aziz, M.H.; Bassyouni, M.; Attar, A.; Al-Hossainy, A.F. Synthesis and characterization of Poly (ortho-aminophenol-co-para-toluidine) and its application as semiconductor thin film. J. Mol. Struct. 2021, 1225, 129131. [Google Scholar] [CrossRef]
  58. Ehsani, A.; Mahjani, M.G.; Jafarian, M. Electrosynthesis of poly ortho aminophenol films and nanoparticles: A comparative study. Synth. Met. 2012, 162, 199–204. [Google Scholar] [CrossRef]
  59. Kundu, S.; Satpati, B.; Mukherjee, M.; Kar, T.; Pradhan, S.K. Hydrothermal synthesis of polyaniline intercalated vanadium oxide xerogel hybrid nanocomposites: Effective control of morphology and structural characterization. New J. Chem. 2017, 41, 3634–3645. [Google Scholar] [CrossRef]
  60. Choudhury, A.; Kim, J.-H.; Yang, K.-S.; Yang, D.-J. Facile synthesis of self-standing binder-free vanadium pentoxide-carbon nanofiber composites for high-performance supercapacitors. Electrochim. Acta 2016, 213, 400–407. [Google Scholar] [CrossRef]
  61. Far, H.M.; Donthula, S.; Taghvaee, T.; Saeed, A.M.; Garr, Z.; Sotiriou-Leventis, C.; Leventis, N. Air-oxidation of phenolic resin aerogels: Backbone reorganization, formation of ring-fused pyrylium cations, and the effect on microporous carbons with enhanced surface areas. RSC Adv. 2017, 7, 51104–51120. [Google Scholar] [CrossRef]
  62. Paul, S.; Choi, K.S.; Lee, D.J.; Sudhagar, P.; Kang, Y.S. Factors affecting the performance of supercapacitors assembled with polypyrrole/multi-walled carbon nanotube composite electrodes. Electrochim. Acta 2012, 78, 649–655. [Google Scholar] [CrossRef]
  63. Asen, P.; Shahrokhian, S.; Zad, A.I. One step electrodeposition of V2O5/polypyrrole/graphene oxide ternary nanocomposite for preparation of a high-performance supercapacitor. Int. J. Hydrogen Energy 2017, 42, 21073–21085. [Google Scholar] [CrossRef]
  64. Chee, W.K.; Lim, H.N.; Harrison, I.; Chong, K.F.; Zainal, Z.; Ng, C.H.; Huang, N.M. Performance of flexible and binderless polypyrrole/graphene oxide/zinc oxide supercapacitor electrode in a symmetrical two-electrode configuration. Electrochim. Acta 2015, 157, 88–94. [Google Scholar] [CrossRef]
  65. Xiong, P.; Hu, C.; Fan, Y.; Zhang, W.; Zhu, J.; Wang, X. Ternary manganese ferrite/graphene/polyaniline nanostructure with enhanced electrochemical capacitance performance. J. Power Source 2014, 266, 384–392. [Google Scholar] [CrossRef]
  66. Moyseowicz, A.; Sliwak, A.; Miniach, E.; Gryglewicz, G. Polypyrrole/iron oxide/reduced graphene oxide ternary composite as a binderless electrode material with high cyclic stability for supercapacitors. Compos. Part B Eng. 2017, 109, 23–29. [Google Scholar] [CrossRef]
  67. Asen, P.; Shahrokhian, S. A High Performance Supercapacitor Based on Graphene/Polypyrrole/Cu2O–Cu(OH)2 Ternary Nanocomposite Coated on Nickel Foam. J. Phys. Chem. C 2017, 121, 6508–6519. [Google Scholar] [CrossRef]
  68. Du, Y.; Wang, X.; Li, B.; Lü, H.; Shen, Y. AQDS-guided growth of nanocilia-like polyaniline on graphene nanofiber as cathode material for high-performance asymmetric supercapacitors. Synth. Met. 2021, 272, 116660. [Google Scholar] [CrossRef]
  69. Shen, Z.M.; Luo, X.J.; Zhu, Y.Y.; Liu, Y.S. Facile co-deposition of NiO-CoO-PPy composite for asymmetric supercapacitors. J. Energy Storage 2022, 51, 104475. [Google Scholar] [CrossRef]
  70. Fan, L.Q.; Liu, G.J.; Wu, J.H.; Liu, L.; Lin, J.M.; Wei, Y.L. Asymmetric supercapacitor based on graphene oxide/polypyrrole composite and activated carbon electrodes. Electrochim. Acta 2014, 137, 26–33. [Google Scholar] [CrossRef]
  71. Poudel, M.B.; Shin, M.; Kim, H.J. Polyaniline-silver-manganese dioxide nanorod ternary composite for asymmetric supercapacitor with remarkable electrochemical performance. Int. J. Hydrogen Energy 2021, 46, 474–485. [Google Scholar] [CrossRef]
  72. Wang, J.G.; Yang, Y.; Huang, Z.H.; Kang, F. A high-performance asymmetric supercapacitor based on carbon and carbon–MnO2 nanofiber electrodes. Carbon 2013, 61, 190–199. [Google Scholar] [CrossRef]
  73. Misnon, I.I.; Jose, R. Synthesis and electrochemical evaluation of the PANI/δ-MnO2 electrode for high performing asymmetric supercapacitors. New J. Chem. 2017, 41, 6574–6584. [Google Scholar] [CrossRef]
  74. Karimi, A.; Kazeminezhad, I.; Naderi, L.; Shahrokhian, S. Construction of a ternary nanocomposite, polypyrrole/Fe–Co sulfide-reduced graphene oxide/nickel foam as a novel binder-free electrode for high-performance asymmetric supercapacitors. J. Phys. Chem. C 2020, 124, 4393–4407. [Google Scholar] [CrossRef]
  75. Han, L.; Tang, P.; Zhang, L. Hierarchical Co3O4@PPy@MnO2 core–shell–shell nanowire arrays for enhanced electrochemical energy storage. Nano Energy 2014, 7, 42–51. [Google Scholar] [CrossRef]
  76. Tang, P.; Han, L.; Zhang, L. Facile Synthesis of Graphite/PEDOT/MnO2 Composites on Commercial Supercapacitor Separator Membranes as Flexible and High-Performance Supercapacitor Electrodes. ACS Appl. Mater. Interfaces 2014, 6, 10506–10515. [Google Scholar] [CrossRef]
  77. Mak, W.F.; Wee, G.; Aravindan, V.; Gupta, N.; Mhaisalkar, S.G.; Madhavi, S. High-energy density asymmetric supercapacitor based on electrospun vanadium pentoxide and polyaniline nanofibers in aqueous electrolyte. J. Electrochem. Soc. 2012, 159, A1481–A1488. [Google Scholar] [CrossRef]
Figure 1. (a) FTIR spectra, (b) Raman spectra, and (c) XRD patterns of V2O5/amino-FG nanocomposite and V2O5/amino-FG/PmAP nanocomposite film.
Figure 1. (a) FTIR spectra, (b) Raman spectra, and (c) XRD patterns of V2O5/amino-FG nanocomposite and V2O5/amino-FG/PmAP nanocomposite film.
Nanomaterials 13 00642 g001
Figure 2. (a) XPS survey spectra XPS spectra of V2O5/amino-FG/PmAP nanocomposite film; the high-resolution spectra of (b) V 2p, (c) O 1s, and (d) N 1s, along with the fitting peaks of as-prepared V2O5/amino-FG/PmAP nanocomposite film.
Figure 2. (a) XPS survey spectra XPS spectra of V2O5/amino-FG/PmAP nanocomposite film; the high-resolution spectra of (b) V 2p, (c) O 1s, and (d) N 1s, along with the fitting peaks of as-prepared V2O5/amino-FG/PmAP nanocomposite film.
Nanomaterials 13 00642 g002
Figure 3. FESEM images of binary V2O5/amino-FG nanocomposite (ac) and V2O5/amino-FG/PmAP nanocomposite film (df) under different magnifications. HR-TEM images of V2O5/amino-FG/PmAP nanocomposite film (g), Selected area electron diffraction (SAED) pattern (h), and EDX elemental mapping of V2O5/amino-FG/PmAP nanocomposite (i).
Figure 3. FESEM images of binary V2O5/amino-FG nanocomposite (ac) and V2O5/amino-FG/PmAP nanocomposite film (df) under different magnifications. HR-TEM images of V2O5/amino-FG/PmAP nanocomposite film (g), Selected area electron diffraction (SAED) pattern (h), and EDX elemental mapping of V2O5/amino-FG/PmAP nanocomposite (i).
Nanomaterials 13 00642 g003
Figure 4. (a) CV curves of V2O5/amino-FG/PmAP and V2O5/amino-FG/PVDF nanocomposite in 1 M KCl and 1 M Li2SO4 at scan rate of 25 mV/s, (b) CV curves of V2O5/amino-FG/PmAP electrode in 1 M KCl electrolyte at different sweep rates, (c) Linear fitting curves between ln (i) Vs. ln (sweep rate), (d) Charge–discharge profiles of V2O5/amino-FG/PmAP nanocomposite at different current densities, (e) Variations of specific capacitance as function of current density for V2O5/amino-FG/PmAP and V2O5/amino-FG/PVDF nanocomposites film, and (f) Specific capacitance retention versus cyclic number for V2O5/amino-FG/PmAP electrode.
Figure 4. (a) CV curves of V2O5/amino-FG/PmAP and V2O5/amino-FG/PVDF nanocomposite in 1 M KCl and 1 M Li2SO4 at scan rate of 25 mV/s, (b) CV curves of V2O5/amino-FG/PmAP electrode in 1 M KCl electrolyte at different sweep rates, (c) Linear fitting curves between ln (i) Vs. ln (sweep rate), (d) Charge–discharge profiles of V2O5/amino-FG/PmAP nanocomposite at different current densities, (e) Variations of specific capacitance as function of current density for V2O5/amino-FG/PmAP and V2O5/amino-FG/PVDF nanocomposites film, and (f) Specific capacitance retention versus cyclic number for V2O5/amino-FG/PmAP electrode.
Nanomaterials 13 00642 g004
Figure 5. (a) CV curves of V2O5/amino-FG/PmAP electrode in 1 M Li2SO4 electrolyte at different sweep rates, (b) Linear fitting curves between ln (i) Vs. ln (sweep rate), (c) Charge–discharge profiles of V2O5/amino-FG/PVDF nanocomposites film in 1 M Li2SO4 electrolyte at different current densities, (d) Variations of specific capacitance as function of current density for V2O5/amino-FG/PmAP and V2O5/amino-FG/PVDF nanocomposites film, (e) Specific capacitance retention versus cyclic number for V2O5/amino-FG/PmAP electrode, and (f) EIS curves of V2O5/amino-FG/PmAP with 1 M KCl and 1 M Li2SO4 electrolyte and V2O5/amino-FG/PVDF electrodes in 1 M KCl with the equivalent electrical circuit model.
Figure 5. (a) CV curves of V2O5/amino-FG/PmAP electrode in 1 M Li2SO4 electrolyte at different sweep rates, (b) Linear fitting curves between ln (i) Vs. ln (sweep rate), (c) Charge–discharge profiles of V2O5/amino-FG/PVDF nanocomposites film in 1 M Li2SO4 electrolyte at different current densities, (d) Variations of specific capacitance as function of current density for V2O5/amino-FG/PmAP and V2O5/amino-FG/PVDF nanocomposites film, (e) Specific capacitance retention versus cyclic number for V2O5/amino-FG/PmAP electrode, and (f) EIS curves of V2O5/amino-FG/PmAP with 1 M KCl and 1 M Li2SO4 electrolyte and V2O5/amino-FG/PVDF electrodes in 1 M KCl with the equivalent electrical circuit model.
Nanomaterials 13 00642 g005
Figure 6. (a) CV curves of V2O5/amino-FG/PmAP//CC ASC device in 1 M KCl electrolyte with different potential windows ranging from 0–0.9 V to 0–1.4 V, (b) CV curves of the ASC device at different sweep rates, (c) Charge–discharge profiles of ASC device in 1 M KCl electrolyte with different potential windows ranging from 0–0.9 V to 0–1.4 V, (d) Charge–discharge (CD) cycling curves for the ASC device at different current densities, (e) Specific capacitance of ASC device over a range of current densities, and (f) Capacitance retention versus cyclic number for the ASC device in 1 M KCl.
Figure 6. (a) CV curves of V2O5/amino-FG/PmAP//CC ASC device in 1 M KCl electrolyte with different potential windows ranging from 0–0.9 V to 0–1.4 V, (b) CV curves of the ASC device at different sweep rates, (c) Charge–discharge profiles of ASC device in 1 M KCl electrolyte with different potential windows ranging from 0–0.9 V to 0–1.4 V, (d) Charge–discharge (CD) cycling curves for the ASC device at different current densities, (e) Specific capacitance of ASC device over a range of current densities, and (f) Capacitance retention versus cyclic number for the ASC device in 1 M KCl.
Nanomaterials 13 00642 g006
Figure 7. (a) CV curves of V2O5/amino-FG/PmAP//CC ASC device in 1 M Li2SO4 electrolyte with different potential windows ranging from 0–1.2 V to 0–1.7 V, (b) CV curves of the ASC device with 1 M Li2SO4 at different sweep rates, (c) Charge–discharge profiles of ASC device in 1 M Li2SO4 with different potential windows, (d) Charge–discharge (CD) cycling curves for the ASC device at different current densities, (e) Specific capacitance of ASC device over a range of current densities, and (f) Capacitance retention versus cyclic number for the ASC device in 1 M Li2SO4.
Figure 7. (a) CV curves of V2O5/amino-FG/PmAP//CC ASC device in 1 M Li2SO4 electrolyte with different potential windows ranging from 0–1.2 V to 0–1.7 V, (b) CV curves of the ASC device with 1 M Li2SO4 at different sweep rates, (c) Charge–discharge profiles of ASC device in 1 M Li2SO4 with different potential windows, (d) Charge–discharge (CD) cycling curves for the ASC device at different current densities, (e) Specific capacitance of ASC device over a range of current densities, and (f) Capacitance retention versus cyclic number for the ASC device in 1 M Li2SO4.
Nanomaterials 13 00642 g007
Figure 8. (a) Nyquist plots of V2O5/amino-FG/PmAP//CC ASC device in 1 M KCl and 1 M Li2SO4, (b) Ragone plots of the ASC devices in 1 M KCl and 1 M Li2SO4 and earlier reported similar types of ASC devices [24,61,62,63,64,65,66,67,68], and (c) CV profiles of as-assembled asymmetric device in KCl electrolyte after bending at different angles.
Figure 8. (a) Nyquist plots of V2O5/amino-FG/PmAP//CC ASC device in 1 M KCl and 1 M Li2SO4, (b) Ragone plots of the ASC devices in 1 M KCl and 1 M Li2SO4 and earlier reported similar types of ASC devices [24,61,62,63,64,65,66,67,68], and (c) CV profiles of as-assembled asymmetric device in KCl electrolyte after bending at different angles.
Nanomaterials 13 00642 g008
Table 1. Comparison of the reported ASCs with recent devices using asymmetric configuration design.
Table 1. Comparison of the reported ASCs with recent devices using asymmetric configuration design.
ASCs Cell ConfigurationElectrolyteED (Wh/kg)PD (W/kg)Ref.
AnodeCathode
NiO-CoO-PPyAC2 M KOH36801[69]
GO/PPyAC1 M Na2SO421.4453.9[70]
PANI/Ag@MnO2AC2 M KOH49.771599.75[71]
MnO2@CNFACNF1 M Na2SO48.72080[72]
PANI–MnO2AC6 M KOH20400[73]
PPy/FeCoS-rGOrGO3 M KOH28.3810[74]
Co3O4/PPy/MnO2AC1 M NaOH34.380[75]
Gra/PEDOT/MnO2AC0.5 M Na2SO431.490[76]
V2O5 NFPANI NF3 M KCl26.7220[77]
V2O5/amino-FG/PmAPAC1 M KCl264607This work
V2O5/amino-FG/PmAPAC1 M Li2SO429.66655This work
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

Hossain, S.S.; Dey, B.; Ali, S.S.; Choudhury, A. Fabrication of Flexible Poly(m-aminophenol)/Vanadium Pentoxide/Graphene Ternary Nanocomposite Film as a Positive Electrode for Solid-State Asymmetric Supercapacitors. Nanomaterials 2023, 13, 642. https://0-doi-org.brum.beds.ac.uk/10.3390/nano13040642

AMA Style

Hossain SS, Dey B, Ali SS, Choudhury A. Fabrication of Flexible Poly(m-aminophenol)/Vanadium Pentoxide/Graphene Ternary Nanocomposite Film as a Positive Electrode for Solid-State Asymmetric Supercapacitors. Nanomaterials. 2023; 13(4):642. https://0-doi-org.brum.beds.ac.uk/10.3390/nano13040642

Chicago/Turabian Style

Hossain, SK Safdar, Baban Dey, Syed Sadiq Ali, and Arup Choudhury. 2023. "Fabrication of Flexible Poly(m-aminophenol)/Vanadium Pentoxide/Graphene Ternary Nanocomposite Film as a Positive Electrode for Solid-State Asymmetric Supercapacitors" Nanomaterials 13, no. 4: 642. https://0-doi-org.brum.beds.ac.uk/10.3390/nano13040642

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

Article Metrics

Back to TopTop