Next Article in Journal
PMMA Thin Film with Embedded Carbon Quantum Dots for Post-Fabrication Improvement of Light Harvesting in Perovskite Solar Cells
Previous Article in Journal
Lanthanide-Doped SPIONs Bioconjugation with Trastuzumab for Potential Multimodal Anticancer Activity and Magnetic Hyperthermia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electroactive Ultra-Thin rGO-Enriched FeMoO4 Nanotubes and MnO2 Nanorods as Electrodes for High-Performance All-Solid-State Asymmetric Supercapacitors

by
Kugalur Shanmugam Ranjith
1,
Ganji Seeta Rama Raju
1,
Nilesh R. Chodankar
1,
Seyed Majid Ghoreishian
2,
Cheol Hwan Kwak
2,
Yun Suk Huh
2,* and
Young-Kyu Han
1,*
1
Department of Energy and Material Engineering, Dongguk University-Seoul, Seoul 04620, Korea
2
Department of Biological Engineering, Inha University, Incheon 22212, Korea
*
Authors to whom correspondence should be addressed.
Nanomaterials 2020, 10(2), 289; https://doi.org/10.3390/nano10020289
Submission received: 10 January 2020 / Revised: 4 February 2020 / Accepted: 5 February 2020 / Published: 9 February 2020
(This article belongs to the Section Energy and Catalysis)

Abstract

:
A flexible asymmetric supercapacitor (ASC) with high electrochemical performance was constructed using reduced graphene oxide (rGO)-wrapped redox-active metal oxide-based negative and positive electrodes. Thin layered rGO functionality on the positive and the negative electrode surfaces has promoted the feasible surface-active sites and enhances the electrochemical response with a wide operating voltage window. Herein we report the controlled growth of rGO-wrapped tubular FeMoO4 nanofibers (NFs) via electrospinning followed by surface functionalization as a negative electrode. The tubular structure offers the ultrathin-layer decoration of rGO inside and outside of the tubular walls with uniform wrapping. The rGO-wrapped tubular FeMoO4 NF electrode exhibited a high specific capacitance of 135.2 F g−1 in Na2SO4 neutral electrolyte with an excellent rate capability and cycling stability (96.45% in 5000 cycles) at high current density. Meanwhile, the hydrothermally synthesized binder-free rGO/MnO2 nanorods on carbon cloth (rGO-MnO2@CC) were selected as cathode materials due to their high capacitance and high conductivity. Moreover, the ASC device was fabricated using rGO-wrapped FeMoO4 on carbon cloth (rGO-FeMoO4@CC) as the negative electrode and rGO-MnO2@CC as the positive electrode (rGO-FeMoO4@CC/rGO-MnO2@CC). The rationally designed ASC device delivered an excellent energy density of 38.8 W h kg−1 with a wide operating voltage window of 0.0–1.8 V. The hybrid ASC showed excellent cycling stability of 93.37% capacitance retention for 5000 cycles. Thus, the developed rGO-wrapped FeMoO4 nanotubes and MnO2 nanorods are promising hybrid electrode materials for the development of wide-potential ASCs with high energy and power density.

Graphical Abstract

1. Introduction

Immense interest in the rapid development of portable flexible electronic tools has promoted the need for lightweight and high-performing energy storage devices to meet the requirement for a resilient power supply [1,2]. Among the electrochemical storage systems, supercapacitors have attracted significant interest as a future power source owing to their high-power densities and exceptionally long cycle lives [3,4]. However, the practical applications of supercapacitors have been generally limited due to their low energy densities. The design of asymmetric supercapacitors (ASCs) with two different sets of electrode systems has provided an efficient strategy for developing the operating voltage window and providing a high energy density to meet the demands of emerging technology [5]. While intensive efforts directed at ASCs have focused on device structures such as V2O5/carbon [6], MnO2/graphene or carbon [7], Co3O4@C@Ni3S2/activated carbon (AC) [8], MnO2/Fe3O4 [9], MnO2/MoS2 [10], CNT-MnO2/CNT-SnO2 [11], graphene/polypyrrole [12], and graphene–NiOH/graphene [13], most of the assembled supercapacitor devices have been configured with liquid electrolytes. Moreover, most of the ASC devices have been configured with pseudocapacitive materials as the positive electrode while carbon-based materials have commonly been used as negative electrodes [5,14]. Until now, intensive research has been focused on exploring the positive electrodes, while negative electrode materials were rarely studied apart from the carbon-based electrodes. Generally, ASCs using carbon-based negative electrodes have suffered due to the low energy density with a lower capacitance of the activated carbon (AC) in the aqueous electrolytes, especially in neutral conditions. Moreover, in comparison with the electrode surface features, the flexible electrode has held many advantages along with high flexibility and stability [15]. At this point, the flexible ASCs are expected to attract a huge demand, with their flexibility providing the key to portable energy storage systems.
Considerable attention has focused on the advantages and durability of Fe- and Mo-based electrodes for ASCs. In particular, the MoO-based electrode has attracted interest as a negative electrode system due to its low electrical resistivity, multiple valance states, rich chemistry, natural abundance, and affordable cost [16]. The incorporation of MoO3 with carbon results in a specific capacitance of 413.08 F g−1 at 8 A g−1 due to the great intercalating ability of MoO3 which promotes the large current density [17]. The MoO2/CC-based negative electrode in NaMnO2/CC//MoO2/CC ASCs exhibits an energy density of 0.92 mW h cm−3 and a maximum volume capacitance of 2.04 F cm−3 along with outstanding cycling stability with 97.33% capacitance retention even after 6000 cycles [18]. The construction of graphene-induced MoO3 nanosheets as a negative electrode supported the short diffusion path and promoted the electron transport pathway. In addition, the inclusion of graphene on both electrode surfaces provided the hybridized interlayers with an extra interface and promoted the pseudocapacitive reaction and rate capability [19]. With the graphene-promoted electrode surface, the GrMnO2/GrMoO3 ASC device exhibited a specific capacitance of 307 F g−1 with an energy density of 42.6 W h kg−1 at a power density of 276 W kg−1. In addition, the Fe2O3-based electrode systems have received significant attention because of their multiple valance states, rich redox chemistry, high specific capacitance, functional capability, and high stability as negative electrodes, high theoretical capacitance, environmentally friendly properties, and natural abundance, which make them suitable for commercialization. The design of Fe2O3 nanotubes as the negative electrode in MnO2/Fe2O3 ASCs results in excellent stability with a wide potential window of 0.0–1.6 V along with an outstanding energy density [20]. The richness of the Fe phase has been shown to promote the redox-active hydroxides in the electrode surface [21]. The construction of nanostructured electrode materials with tubular assemblies has significant advantages, including high surface area, effective heat transfer ability and a favorable electrode–electrolyte interface [22,23,24,25,26]. Decorating carbon nanotube (CNT) sponges with Fe2O3 horns to generate a porous electrode has resulted in a high specific capacitance of 296 F g−1 owing to the combined influence of the electric double-layer capacitance (EDLC) of the CNTs and the pseudocapacitance of Fe2O3 [27]. The inclusion of the ternary FeMoO4 network has demonstrated the advantage of promoting the individual functional properties to provide a high-performing electrode. The design of molybdenum-containing transition metal oxides has been promoted as an advanced anode material through the high oxidation of molybdate [27]. The high theoretical capacitance of 992.3 mA h g−1, attributed to the high valance of Fe3+ and Mo6+, has prompted interest in FeMoO4-based electrodes as the anode material in ASCs. The reduced graphene oxide (rGO)–FeMoO4 nanocomposite has provided a specific capacitance of 135 F g−1 and the promotional impact of rGO was found to induce lower electrochemical resistance and a longer cycle life [28]. Electrospinning is a feasible approach for designing one-dimensional fiber assembly even in binary or ternary compositions [29,30]. Performing experiments using polymerics with different physical parameters led to control of the morphology of the electrospun fibers using the effects of external or post-thermal treatments [31,32]. Tuning the surface area by controlling the morphology and inducing a highly conductive functional network has enhanced the potential of the heterostructural electrode system.
In the present work, the heterostructural rGO/FeMoO4 hybrid nanotube network was constructed via electrospinning followed by a hydrothermal technique. The fabrication of tubular FeMoO4 NFs with the twin-walled rGO functionalities results in the design of a hybrid network with a high surface area. Benefiting from the tubular assembly and the layered nanostructures of FeMoO4 and rGO flakes, the prepared FeMoO4/rGO electrodes exhibited superior electrochemical performances with a high specific capacitance of 135.2 F g−1 at 2 A g−1. Considering the advantages of the FeMoO4/rGO electrode material, a flexible ASC was fabricated using polyvinyl alcohol (PVA)/Na2SO4 electrolyte with FeMoO4/rGO and MnO2/rGO as negative and positive electrodes, respectively. For the first time, new ASC systems based on MnO2 and FeMoO4 were fabricated using a cost-effective, facile and scalable approach. In particular, the design of the tubular FeMoO4 NFs with highly conductive ultra-thin rGO features resulted in efficient electrochemical responses of the active surfaces due to the high internal surface area and short diffusion path of the hybrid electrode surface. The MnO2 nanorods were directly grown on the carbon cloth (CC) which facilitates the effective electrical mobility. The CC serves as a flexible current collector and as a scaffold for the heterostructural network. The ASC assembly of rGO–FeMoO4/rGO–MnO2 displayed a specific capacitance of 54.4 F g−1 at 2 A g−1 and delivered a wide voltage window of 0.0–1.8 V. With the advantage of a high-performing negative electrode with a wide potential window, the assembled ASC device exhibited a high energy density of 38.8 W h kg−1 at a power density of 1344.5 W kg−1 and a high capacitance retention of 93.37%. This work demonstrates that the FeMoO4/rGO hybrid electrodes have extended the voltage window compared to the carbon-based negative electrodes and have a broad potential application in flexible ASCs.

2. Materials and Methods

2.1. Preparation of rGO Ultra-Thin Nanoflakes

Two-dimensional (2D) well-dispersed graphene oxide (GO) was synthesized via the Hummers method [33]. A 3:1 mixture of H2O2 and NH3OH (40 mL) was added dropwise to a homogeneous dispersion of GO (60 mL, 1 mg mL−1) in a round-bottomed flask and reflexed at 200 °C for 24 h. Finally, a well-dispersed ultra-thin rGO solution was obtained.

2.2. Preparation of FeMoO4/rGO Hybrids

The FeMoO4 nanotubes were prepared via a facile electrospinning process followed by thermal annealing. In a typical process, 1.3 g of polyacrylonitrile (PAN), poly(methyl methacrylate) PMMA (8:2) in DMF (10 mL) was stirred for 3 h to form the homogeneous precursor solution. Further, FeCl3 (20 wt.% with polymeric source) and MoCl3 (10 wt.% with polymeric source) were added to the polymeric solution and stirred overnight to attain a homogeneous clear solution. The polymeric sol was then loaded into a plastic syringe with a 23-gauge injection needle, which was connected to the high-voltage power supply (Nano EC). Under the flow rate of 1 mL h−1 with 12 kV of applied voltage, the electrospun fibers were collected in the aluminum foil-covered drum collector which was maintained at a distance of 15 cm from the needle. The as-spun Fe- and Mo-loaded PAN/PMMA composite NFs were calcinated at 500 °C for 3 h in the air to yield the pale reddish hollow FeMoO4 NFs. The effective decomposition of PMMA created the tubular morphology, while the later decomposition of PAN provided a 1D scaffold with a fibrous morphology. For comparison, the electrospinning was also performed without PMMA in the precursor solution to obtain the solid NFs. A 0.3 g mass of as-prepared NFs was subjected to the surface functionalization using (3-aminopropyl)trimethoxysilane (ATS, 1 mL) in isopropyl alcohol (IPA, 60 mL) at 60 °C overnight, and the unattached silane functionalities were removed by multiple centrifugations. The ultra-thin rGO nanoflakes were coated onto the FeMoO4 NF surface via the facile hydrothermal technique. First, a well-dispersed 30 mL of rGO solution (0.6 mg/mL) was dissolved in 30 mL of deionized (DI) water and 0.3 g of hollow FeMoO4 NFs were then dispersed in the solution under mild stirring. Finally, the solution was transferred to the Teflon-lined autoclave and heated at 140 °C for 6 h. Under the thermal treatment, the well-dispersed ultrathin rGO was functionally attached to the hollow FeMoO4 nanofibrous surface and the strong bonding interaction with the functionalized FeMoO4 NFs was favored. The precipitated rGO–FeMoO4 NFs were centrifuged, washed three times with DI water and ethanol, and vacuum dried at 50 °C for 12 h.

2.3. Preparation of MnO/rGO Hybrids

The cleaned carbon cloth (CC) with dimensions of 2 × 1 cm (4 numbers) were immersed in aqueous KMnO4 solution (35 mL, 0.84 mg/mL) and dispersed in 15 mL of rGO solution (0.6 mg/mL) was then added dropwise under constant stirring. The homogeneous aqueous dispersion was obtained via the hydrothermal process in a Teflon-lined stainless-steel container which was sealed and maintained at 160 °C for 12 h. After cooling down, the CC was collected and cleaned with DI water and dried at 110 °C for 12 h in a vacuum. For comparison, un-modified MnO2 (without rGO) was also prepared on the CC.

2.4. Characterization Methods

The structural analyses of the as-prepared MnO2 and FeMoO4 fibers were performed by powder X-ray diffraction (X’Pert-PRO MRD, Philips, The Netherlands). The morphologies of the active materials were examined by high-resolution field emission microscopy (HRSEM, Hitachi, Japan) and high-resolution transmission electron microscopy (FETEM, JEM-2100F, JEOL, Japan), and the elemental compositions were analyzed by energy-dispersive X-ray spectroscopy (EDX, Oxford Instruments, High Wycombe, UK). The structural compositions and electronic states were analyzed by X-ray photoelectron spectrometry (XPS, Thermo Fisher Scientific, K-Alpha, USA) with non-monochromatic Al K-α radiation (1486.6 eV). The Brunauer–Emmett–Teller (BET) specific surface area (SBET) was measured by the multipoint BET method (ASAP ZOZO, Micromeritics, GA, USA).

2.5. Electrochemical Measurement

The FeMoO4 NF-based working electrode was prepared by mixing the electroactive NFs, carbon black and polyvinylidene fluoride (PVDF) binder in a mass ratio of 80:10:10 with N-methyl pyrrolidinone (NMP) and grinding to form a homogeneous slurry. The resulting slurry was coated onto the CC (1 × 1 cm2), followed by drying at 70 °C for 12 h in a vacuum oven. As the MnO2 was directly grown on the carbon fabric, it was utilized as a binder-free electrode. All the electrochemical measurements were performed at room temperature on the Metrohm Auto lab workstation (PGSTAT302N). The Ag/AgCl electrode and platinum foil were used as a reference and counter electrodes, respectively, in a three-electrode cell and 1 M Na2SO4 used as an aqueous electrolyte. Electrochemical tests such as cyclic voltammetry (CV) and galvanostatic charge–discharge (GCD) were performed at various scan rates and current densities between −0.8 and 0 V and between 0 and 1 V vs. Ag/AgCl for the rGO/FeMnO4 and rGO/MnO2 electrodes, respectively. Electrochemical impedance spectroscopy (EIS) was performed with the potential amplitude of 5 mV over the frequency range of 100 kHz to 0.01 Hz. The specific capacitance was calculated from the GCD curve according to the equation:
C = I   Δ t m   Δ V
where C is the specific capacitance (F g−1), Δt is the total discharge time (s), I is the constant discharge current (A), m is the mass of the active material in the electrode (g), and ΔV is the applied potential window (V).

2.6. Fabrication of the All-Solid-State ASC Device

The all-solid-state flexible ASC device was fabricated using rGO/MnO2@CC as the positive electrode and rGO/FeMoO4@CC as the negative electrode in the PVA/Na2SO4 gel electrolyte. The PVA/Na2SO4 gel was prepared by mixing 2 g of PVA and 0.5 M Na2SO4 in deionized (DI) water (20 mL) and heating at 85 °C with constant stirring until the mixture became clear. The gel electrolyte was first coated uniformly onto the rGO/MnO2@CC and rGO/FeMoO4@CC electrodes and dried for 2 h in the open air. Finally, the ASC device was fabricated by combining the two electrodes with the gel electrolyte and the device assembly was packed with insulting tape and dried overnight in a hood at room temperature. The loading mass of the active electrode materials was set according to the following equation:
m + / m   =   ( C   ×   Δ V ) / ( C +   ×   Δ V + )
where C is the specific capacitance, Δ V is the potential window, the + and − indicate the positive and negative electrode, respectively, and m is the mass of the active electrode material. The energy density (E) for the ASC was calculated according to E = CV2/(2 × 3.6), while the power density (P) was evaluated according to P = E/ Δ t (where Δ t is the discharge time in GCD).

3. Results

3.1. Constructing the Negative Electrode Material

Scheme 1 presents the growth strategy for preparing the tubular FeMoO4 NFs with the ultra-thin rGO functionalities. The co-polymeric precursors PAN and PMMA used for electrospinning with the Fe and Mo sources induce the tubular morphology by means of their two distinct decomposition temperatures. The as-spun FeMo–PAN/PMMA fibers were thermally decomposed at 500 °C for 3 h to prepare the tubular FeMoO4 NFs. In the second step, the thermally reduced ultra-thin GO was functionalized on the tubular FeMoO4 NFs via the hydrothermal process to improve the surface functionality of the NFs. Figure S1 presents the SEM images of the as-spun FeMo-PAN and FeMo-PAN/PMMA, indicating that the obtained NFs had diameters of 450 nm and 480 nm, respectively. Figure 1 shows the structural morphologies of the tubular FeMoO4 and rGO/FeMoO4 NFs indicated by the SEM and TEM analyses. The SEM image in Figure 1a shows that the FeMoO4 NF synthesized using the PAN scaffold has a fiber diameter of 250 nm with a solid core. The high-resolution SEM image in Figure 1b shows the dense grain structure of the FeMoO4 fiber surface with quite smooth surface features, whereas the addition of PMMA/PAN as a scaffold in the spinning process (Figure 1c) results in the formation of tubular FeMoO4 NFs having diameters of around 250 nm. The high-resolution image in Figure 1d shows the tubular FeMoO4 NF with a shell wall thickness of 20 nm. The inclusion of PMMA with the PAN scaffold leads to earlier thermal decomposition which favors the formation of the tubular structure, while the PAN network promotes the stability of the fibers during the annealing process. When PMMA is used alone as the scaffold, aggregation of the FeMoO4 nanograins leads to the complete collapse of the fibrous morphology after annealing (Figure S1). Figure 1e,f present the SEM images of the rGO-functionalized FeMoO4 tubular NFs which indicate the hybrid structure of an ultrathin rGO layer covering the tubular NFs and the wrinkled rGO on the FeMoO4 NF surface. Further, rGO functionality on the FeMoO4 tubular NFs was confirmed through the Raman scattering spectra. Figure S2 results in the peaks at 816, 904 and 949 cm−1 are assigned to FeMoO4 [28,34]. In addition, with the completion of rGO wrapping, rGO–FeMoO4 NFs displayed additional peaks at 1360 and 1609 cm−1, representing the D and G bands of graphene, thereby confirming the wrapped features of rGO along with the FeMoO4 NF [35]. The ultrathin rGO layer is tightly bonded with the FeMoO4 surface and promotes the stability and surface conductivity during electrolyte ion diffusion and enhances the electrochemical performance. The tubular structure promotes the rGO loading on the inner and outer surfaces of the tubular features, leading to the twin shell wall effect on the hybrid fibrous surface. The pristine thermally reduced rGO (Figure S3) consists of a few layers of rGO flakes with sizes in the range of 40 to 80 nm, which facilitate binding to the inner and outer walls of the FeMoO4 tubular NFs. Through the ATS-based surface functionalization, the rGO layers were selectively bound to the fiber surface without any aggregation.
The structural and morphological properties of the rGO–FeMoO4 tubular NFs were further examined by TEM and HRTEM analyses. The TEM images of the rGO nanoflakes (Figure S3) indicate the ultra-thin layered features, while the TEM images of the bare FeMoO4 nanotubes indicate the nanograin features of the tubular wall (Figure S4). The high magnification TEM images of the rGO–FeMoO4 nanotubes are presented in Figure 1g,h, the latter image showing the heterostructure of well-dispersed surface-functionalized ultrathin rGO layers on the tubular wall of the FeMoO4 NFs. The high-resolution TEM images of the tubular rGO–FeMoO4 NFs in Figure 1h also shows the decoration of ultrathin rGO layers on the tubular wall with a size of 15 nm. The HRTEM results showed that the nanograins had a lattice spacing of 0.316 nm in the (220) crystal plane of the FeMoO4 and that the rGO functionalization on the tubular wall resulted from layer formation over the nanograins (Figure S5). The structural crystallinity of the tubular fiber was investigated through the selected area electron diffraction (SAED) pattern (inset of Figure 1g), in which the diffuse ring pattern corresponds to the carbon features and the dots represent the crystalline nature of the FeMoO4 NF. The EDX results demonstrate the presence of Fe, Mo, O, and C on the fibrous surface (Figure S6). The corresponding elemental mapping (Figure 1h) confirms that the elemental distribution of C, Fe, Mo, and O on the fibrous surface is homogeneous and that the presence of the carbon features results in the rGO surface functionality and the decomposed PAN network in the form of carbon over the fibrous surface.
Figure 2a presents the XRD pattern of the rGO-functionalized FeMoO4 nanofibers, which is well-indexed to the monoclinic phase of β-FeMoO4 (Joint Committee on Powder Diffraction Standards (JCPDS) card No. 01-89-2367). The additional broad peak at 26° on the rGO/FeMoO4 nanofibers next to the FeMoO4 indicates the surface functionality of rGO on the FeMoO4 surface. The trace appearance of cubic Fe3O4 (JCPDS card No: 65-3107) indicates the richness of Fe ions on the tubular fiber surface which results in faster oxidation to form Fe3O4 on the fiber surface. The survey XPS spectrum confirmed the existence of Fe, Mo, C, and O on the fiber surface (Figure 2b). The Fe 2p core-level XPS spectrum (Figure 2c) indicates the presence of mixed oxidization states of Fe2+ (711.1 eV and 724.9 eV) and Fe3+ (713.3 eV and 727.6 eV) with the satellite peak at 718.2 eV [36]. Figure 2d indicates the high-resolution Mo 3d with two prominent peaks at 232.5 eV and 235.7 eV due to the binding energies of Mo 3d3/2 and Mo 3d5/2, respectively, corresponding to the Mo6+ oxidization states in FeMoO4 [37,38]. The high-resolution C 1s spectra were deconvoluted (Figure 2e) with a C–C vibration at 284.6 eV, oxygen-containing vibrations at 286.5 eV (C–O) and 288.7 eV (O−C=O), and the trace of C–N interactions at 285.7 eV [39]. The high-resolution O 1s spectra were deconvoluted into two peaks at 530.3 eV and 531.8 eV representing the lattice oxide vibration and chemisorbed oxygen with Fe and Mo ions, respectively [40]. The nitrogen adsorption–desorption isotherm was investigated to explore the specific surface area and the pore size distribution of the rGO-loaded FeMoO4 tubular nanofibers (Figure S7). The tubular fibers resulted in the BET specific surface area of 5.64 m2 g−1 and the isotherm demonstrates a type IV hysteresis loop, suggesting a mesoporous structure. The pore size distribution was calculated to be about 27.1 nm according to the Barrett–Joyner–Halenda (BJH) method, which confirms the mesoporous structure.
Figure 3a presents the CV curve of the FeMoO4 and rGO–FeMoO4 hybrid electrodes within the potential range of −0.8 to 0 V (vs. Ag/AgCl) at the scan rate of 10 mV s−1. The hybrid electrodes displayed noticeable redox peaks and demonstrated that, with the rGO feature, the FeMoO4 electrodes possessed larger integral areas than the pristine FeMoO4 tubular fibers and, hence, the higher specific capacitance (Figure S8). Figure 3b presents the CV curves of the tubular rGO–FeMoO4 electrodes at various scan rates ranging from 1 to 200 mV s−1. Even at the high scan rate, the CV curve of the rGO–FeMoO4 electrode maintained its typical shape, indicating the excellent rate capability. The GCD curve results indicate that, at the various scan rates within the potential window of −0.8 V to 0 V, the rGO–FeMoO4 electrode displays high coulombic efficiency and ideal capacitance behavior with a superior charge storage rate. Compared with the pristine tubular FeMoO4 electrode, the rGO–FeMoO4 electrodes displayed specific capacitances at similar current densities, comparable to the above CV results. However, there is only a small IR drop of 0.028 V at a high current density of 20 A g−1 (Figure 3d), indicating the excellent electrical conductivity and efficient mobility of electrolyte ions. Figure 3e indicates the calculated specific capacitance of the hybrid rGO–FeMoO4 electrode at various current densities from the GCD curve. The rGO–FeMoO4 hybrid electrode displayed a specific capacitance of 135.2 F g−1 at 2 A g−1, which is much greater than that of FeMoO4 (32.6 F g−1) at the same current density, and remains as high as 100.3 F g−1 even at the high current density of 20 A g−1, thus demonstrating good rate capability. The synergistic effect between the thin-layered rGO and FeMoO4 NFs increases the specific capacitance of the rGO-wrapped FeMoO4 electrode. Further, the rGO acts as a conductive channel and an active interface on the FeMoO4 nanofibers. The rGO–FeMoO4 hybrid electrodes with various rGO loadings have been compared (Figure S9) for the investigation. However, while the efficient loading of the rGO (60% and 20%) resulted in good rate performances as compared to others, their specific capacitances (82.5 F g−1 and 127.5 F g−1) were much lower than for the 40% loading of rGO on the FeMoO4 fibers (135.2 F g−1). Further, the 80% loading of rGO resulted in the lower specific capacitance of 57.5 F g−1. Highly-loaded rGO on FeMoO4 may completely arrest the active sites through high surface capping, which would restrict the utility of FeMoO4 in the hybrids. The electrochemical impedance spectroscopy (EIS) spectra in Figure 3f indicates the fundamental behavior of the electrode surface material. The high-frequency semicircle indicates the internal charge transfer resistance and the vertical line in the low-frequency region indicates the diffusion limit of the rGO-FeMoO4 hybrid electrodes. The Nyquist plot in Figure 3f shows that the rGO–FeMoO4 hybrid has a much higher phase angle in the lower frequency range, indicating the ideal capacitive nature and low charge transfer resistance of 1.73 Ω due to the fast ion diffusion via the tubular surface with the high surface conductive skeleton which accelerates the electrolyte ionic transport. In Figure 3g, the outstanding electrochemical stability of the rGO–FeMoO4 electrodes is indicated by the 3.55% loss of initial specific capacitance and 98.34% of its coulombic efficiency after 5000 cycles. The inset in Figure 3g, the overlapping function of CV and GCD curves, indicates excellent stability before and after the cycling stability and EIS. The cyclic stability results in a slight increase in charge transfer resistance of 0.11 Ω (Rs). The exceptional stability and promising electrochemical performance of the rGO–FeMoO4 electrodes were ascribed to the following factors: (i) the bimetallic features of Fe and Mo induced more redox sites to promote the electrochemical response of the fiber network; (ii) the integral configuration of rGO bound to the FeMoO4 surface decreased the internal charge resistance and promoted the efficient ion diffusion channel with the proper utilization of active material; and (iii) the tubular morphology provided the highly exposed surface active sites with shortened ionic diffusion.

3.2. Constructing the Positive Electrode Material

To fabricate the cathode material, rGO–MnO2 nanorods were grown directly on the CC via a facial hydrothermal process at 160 °C. The results of the SEM and TEM morphological analyses of the rGO–MnO2 are presented in Figure 4. As a binder-free electrode, the MnO2 nanoflakes were effectively grown on the CC as a hierarchical platform with self-stacked structures (Figure 4a,b). The composite features were more clearly characterized by the TEM analysis (Figure 4c,d). It can be observed that the individual nanoflakes were tightly tagged with the rGO layers on their surface. The lattice spacing of the MnO2 nanoflakes was around 0.69 nm, representing the (001) crystal planes of Birnessite MnO2 (Figure 4e) [41]. The SEAD pattern (Figure 4f) reveals the polycrystalline nature. In the XRD spectrum (Figure 4g) the broad band at 26.4° corresponds to the carbon substrate (CC) and the other diffraction peaks correspond to tetragonal α-MnO2. The trace appearance of a broad peak at 26.4° in the rGO–MnO2 samples results from the rGO oxide features on the electrode surface. The XPS survey spectrum of the rGO–MnO2 presented in Figure 4h indicates the existence of the elements Mn, O and C. The high-resolution Mn 2p spectrum (inset of Figure 4h) displays two prominent peaks at 641.6 eV and 653.4 eV separated by 11.8 eV, confirming the presence of MnO2 in the composite [42].
Figure 5a presents the CV curves of the hydrothermally grown free-standing MnO2 and rGO/MnO2 electrodes at a scan rate of 50 mV s−1. The rGO features on MnO2 promote the redox signal under the applied potential window. The CV curve of the rGO–MnO2 electrode at various scan rates from 1 to 200 mV s−1 (Figure 5b) within the potential window of 0–1 V reveals the typical rectangular profile with an excellent rate capability. Figure 5c presents the GCD curves of the rGO–MnO2 free-standing electrode at various current densities from 1 to 20 A g−1 within the potential window of 0–1 V. The GCD curves of the rGO–MnO2 electrode are highly linear with high coulombic efficiency, indicating ideal capacitive character. Thus, the rGO–MnO2 electrode displays high coulombic efficiency in a wide potential window along with the superior specific capacitance of 324.5 F g−1 at 1 A g−1 (Figure 5d), which is much more efficient than the pristine MnO2 (232.3 F g−1). The Nyquist plot (Figure 5e) displays a semicircle (Rs = 3.28 Ω) in the high-frequency region and a vertical plot in the low-frequency region, indicating the excellent rate capability. Figure 5f indicates the cycling stability of the rGO–MnO2 free-standing electrode with the capacitance retention of 90.26% after 5000 cycles at the current density of 10 A g−1.

3.3. Construction of the rGO–FeMoO4@CC/rGO–MnO2@CC ASC Device

In view of the excellent electrochemical performances of the positive (rGO–MnO2) and negative (rGO–FeMoO4) electrodes in the aqueous electrolyte, the flexible ASC device was fabricated using rGO–MnO2 as the positive electrode and rGO–FeMoO4 as the negative electrode and Na2SO4/PVA solid gel electrolyte. The assembled device is depicted in Figure 6a. The mass loading of the ASC device was balanced by following the relationship of q+ = q prior to assembly. Figure 6b shows the working potential windows of both the rGO–MnO2 and rGO–FeMoO4 electrodes at the scan rate of 30 mV s−1 with the Na2SO4-based electrolyte. With the advantages of a wide potential window and high specific capacitance, the ASC device was expected to operate at potentials of up to 1.8 V. Under the fixed scan rate, Figure 6c presents the CV curves of the rGO–FeMoO4@CC/rGO–MnO2@CC ASC within various potential windows ranging from 0–0.8 V to 0–2.2 V (Figure 6c). The rGO–MnO2@CC/rGO–FeMoO4@CC ASC device provides evidence of synergistic electrochemical performances up to the potential window of 0–1.8 V. The typical CV curve (Figure 6d) of the rGO–MnO2@CC/rGO–FeMoO4@CC ASC device at various current densities within the potential window of 0–1.8 V demonstrates the promising pseudocapacitive storage behavior even at high current density.
Figure 6e presents the galvanic charge device (GCD) curve of the rGO–MnO2@CC/rGO–FeMoO4@CC ASC device at various current densities within the potential window of 0–1.8 V, which clearly indicates highly symmetric curves with a high coulombic efficiency (>95%) for all the respective current densities, thereby demonstrating the excellent reversible efficiency. The specific capacitance values of the rGO–MnO2@CC/rGO–FeMoO4@CC ASC estimated from the GCD curve are 57.7, 54.4, 53.4, 52.1, 50.2, 44.4, and 44.1 F g−1 at the respective current densities of 1, 2, 3, 4, 5, 10, and 20 A g−1. The loss in specific capacitance at high current density is ascribed to the increase in the internal resistance of the hybrid device assembly. Interestingly, the rGO–MnO2@CC/rGO–FeMoO4@CC ASC devices retained 76.4% of their initial capacitance on increasing the current density by 20 times, demonstrating the high rate capability of the ASC electrode surface. In order to understand the electron/ion transport performance, Figure 6f presents the Nyquist EIS spectra of the rGO–MnO2@CC/rGO–FeMoO4@CC ASC electrode surface. The internal resistance (Rs) was calculated as 6.72 Ω and the internal charge transfer resistance was calculated as 1.13 Ω from the semicircle in the high-frequency region. The results of the cyclic stability tests for the rGO–MnO2@CC/rGO–FeMoO4@CC ASC devices indicated that about 93.37% of the primary capacitance was retained after 5000 GCD cycles at the current density of 4 A g−1 (Figure 6h). The SEM images presented in Figure S10 indicates that the rGO–FeMoO4 NFs retained their morphology after the cyclic test, thus demonstrating their potential stability. The obtained results indicated the superior electrochemical performances of the rGO–MnO2@CC/rGO–FeMoO4@CC ASC devices along with long-term stability and a wider potential window than the previous reports summarized in Table S1.
Figure 7a presents a Ragone plot of the fabricated ASC device derived using the GCD results. The rGO–MnO2@CC/rGO–FeMoO4@CC device exhibited outstanding energy of 38.8 W h kg−1 at the power density of 1344.5 W kg−1. Further, at the high-power density of 26,872.9 W kg−1, the ASC hybrid device maintained an energy density of 29.9 W h kg−1, thus demonstrating an outstanding rate performance. In a comparative investigation, the energy and power densities of the rGO–MnO2@CC/rGO–FeMoO4@CC ASC were compared with those of recently reported ASCs such as Co3O4@MnO2/MEGO (17.7 W h kg−1 at 158 W kg−1) [43], ZnCo2O4@MnO2/AC (29.41 W h kg−1 at 628.4 W kg−1) [44], ZnCo2O4@NixCo2x(OH)6x/AC (26.2 W h kg−1 at 511.8 W kg−1) [45], CaMoO4/AC (18.7 W h kg−1 at 362 W kg−1) [46], MnO2/Fe2O3 (53.55 W h kg−1 at 1280 W kg−1) [47], MnO2–GNS/FeOOH–GNS CNTs (30.4 W h kg−1 at 237.6 W kg−1) [48], MnO2/FeOOH (24 W h kg−1 at 450 W kg−1) [49], and MnO2 nanowire/graphene (30.4 W h kg−1 at 100 W kg−1) [50] (Figure 7a and Table S1). A photographic image of three rGO–MnO2@CC/rGO–FeMoO4@CC ASC devices connected in series to light two red LEDs are presented in Figure 7b. The operation of the flexible ASC device under different bending positions, and the respective CV curves, are shown in Figure 7c. The CV curves of the rGO–MnO2@CC/rGO–FeMoO4@CC ASC with various bending angles are almost identical to that in the absence of any deformation.
The photographs of the ASC device in various bending positions demonstrate the high flexibility of the rGO–MnO2@CC/rGO–FeMoO4@CC ASC device. The obtained results demonstrated the fabrication of a high-performing flexible ASC device as an integrated hybrid smart textile. The notably promising performances of the rGO–MnO2@CC/rGO–FeMoO4@CC ASC device can be ascribed to the following causes: (i) the promotion of electroactive sites by the tubular and layered features of rGO-bonded FeMoO4 nanostructures, and (ii) the extended potential window of the electrode surface provided by the unique construction of rGO–FeMoO4 and rGO–MnO2 on the flexible CC. It is, therefore, demonstrated that the rGO–MnO2@CC/rGO–FeMoO4@CC ASC device fabricated at low cost holds great promise for highly efficient energy storage in flexible modern devices.

4. Conclusions

In summary, we have demonstrated the fabrication of a flexible hybrid rGO–FeMoO4@CC/rGO–MnO2@CC-based all-solid-state ASC device via a low-cost, simple strategy using electrospinning and hydrothermal processes. The as-prepared rGO-bonded tubular FeMoO4 NFs possessed a specific capacitance of 135.2 F g−1 due to their mixed oxide states with a high surface area. Tagging the ultrathin rGO layers onto the active material surface provided a surface-conductive channel, further increasing the effective path for electron transport and promoting the morphological stability of the active materials. The fabricated rGO–MnO2@CC/rGO–FeMoO4@CC ASC using rGO–MnO2@CC as the positive electrode and rGO–FeMoO4@CC as the negative electrode exhibited a specific energy density of 38.8 W h kg−1 with efficient cycling stability and excellent rate capability. These results deliver valuable insights on the fabrication of electrode materials with high surface-active sites using the rGO-based hybrid platform to increase the potential window and specific energy density of ASCs.

Supplementary Materials

The following are available online at https://www.mdpi.com/2079-4991/10/2/289/s1: ESI includes the SEM images of the as-prepared and annealed FeMo-polymer fibers with different polymer; Raman spectra of the pristine and rGO functionalized FeMoO4 tubular nanofibers; TEM image of the ultra-thin layered rGO nanoflakes, pristine tubular FeMoO4 nanofibers, and rGO wrapped tubular FeMoO4 nanofibers; EDAX and N2 adsorption–desorption spectra of the rGO wrapped tubular FeMoO4 nanofibers; electrochemical results of the tubular FeMoO4 nanofibers with different rGO ratio; morphological stability of rGO wrapped tubular FeMoO4 nanofibers after cycling performances.

Author Contributions

Conceptualization, K.S.R.; performed the experiments K.S.R., S.M.G., and C.H.K.; analyzed the data, K.S.R.; contributed for discussion, K.S.R., G.S.R.R., and N.R.C.; writing original draft preparation, K.S.R.; writing review and editing, Y.S.H. and Y.-K.H.; funding acquisition, Y.-K.H. All authors have read and agreed to the published version of the manuscript.

Funding

We acknowledge financial support from the Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Science, ICT & Future Planning (2019R1F1A1061477 and NRF- 2014R1A5A1009799). Further, this work was supported by the Dongguk University Research Fund of 2019 (S-2019-G0001-00050).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dubal, D.P.; Ayyad, O.; Ruiz, V.; Gómez-Romero, P. Hybrid energy storage: The merging of battery and supercapacitor chemistries. Chem. Soc. Rev. 2015, 44, 1777–1790. [Google Scholar] [CrossRef]
  2. Noori, A.; Kady, M.F.E.; Rahmanifar, M.S.; Kaner, R.B.; Mousavi, M.F. Towards establishing standard performance metrics for batteries, supercapacitors and beyond. Chem. Soc. Rev. 2019, 48, 1272–1341. [Google Scholar] [CrossRef]
  3. Wang, F.; Wu, X.; Yuan, X.; Liu, Z.; Zhang, Y.; Fu, L.; Zhun, Y.; Zhou, Q.; Wu, Y.; Huang, W. Latest advances in supercapacitors: From new electrode materials to novel device designs. Chem. Soc. Rev. 2017, 46, 6816–6854. [Google Scholar] [CrossRef]
  4. Miller, E.E.; Hua, Y.; Tezel, F.H. Materials for energy storage: Review of electrode materials and methods of increasing capacitance for supercapacitors. J. Energy Storage 2018, 20, 30–40. [Google Scholar] [CrossRef]
  5. Shao, Y.; El-Kady, M.F.; Sun, J.; Li, Y.; Zhang, Q.; Zhu, M.; Wang, H.; Dunn, B.; Kaner, R.B. Design and Mechanisms of Asymmetric Supercapacitors. Chem. Rev. 2018, 118, 9233–9280. [Google Scholar]
  6. Jiang, H.; Cai, X.; Qian, Y.; Zhang, C.; Zhou, L.; Liu, W.; Li, B.; Lai, L.; Huang, W. V2O5 embedded in vertically aligned carbon nanotube arrays as free-standing electrodes for flexible supercapacitors. J. Mater. Chem. A 2017, 5, 23727–23736. [Google Scholar] [CrossRef]
  7. Zhai, T.; Wang, F.; Yu, M.; Xie, S.; Liang, C.; Li, C.; Xiao, F.; Tang, R.; Wu, Q.; Lu, X.; et al. 3D MnO2-graphene composites with large areal capacitance for high-performance asymmetric supercapacitors. Nanoscale 2013, 5, 6790–6796. [Google Scholar] [CrossRef]
  8. Kong, D.; Cheng, C.; Wang, Y.; Wong, J.I.; Yang, Y.; Yang, H.Y. Three-dimensional Co3O4@C@Ni3S2 sandwich-structured nanoneedle arrays: Towards high-performance flexible all-solid-state asymmetric supercapacitors. J. Mater. Chem. A 2015, 3, 16150–16161. [Google Scholar] [CrossRef]
  9. Sun, J.; Huang, Y.; Fu, C.; Huang, Y.; Zhu, M.; Tao, X.; Zhi, C.; Hu, H. A high performance fiber-shaped PEDOT@MnO2//C@Fe3O4 asymmetric supercapacitor for wearable electronics. J. Mater. Chem. A 2016, 4, 14877–14883. [Google Scholar] [CrossRef]
  10. Yang, X.; Niu, H.; Jiang, H.; Wang, Q.; Qu, F. A high energy density all-solid-state asymmetric supercapacitor based on MoS2/graphene nanosheets and MnO2/graphene hybrid electrodes. J. Mater. Chem. A 2016, 4, 11264–11275. [Google Scholar] [CrossRef]
  11. Ng, K.C.; Zhang, S.; Peng, C.; Chen, G.Z. Individual and Bipolarly Stacked Asymmetrical Aqueous Supercapacitors of CNTs/SnO2 and CNTs/MnO2 Nanocomposites. J. Electrochem. Soc. 2009, 156, A846–A853. [Google Scholar] [CrossRef]
  12. Liu, G.; Shi, Y.; Wang, L.; Song, Y.; Gao, S.; Liu, D.; Fan, L. Reduced graphene oxide/polypyrrole composite: An advanced electrode for high-performance symmetric/asymmetric supercapacitor. Carbon Lett. 2019, 1–9. [Google Scholar] [CrossRef]
  13. Yan, J.; Fan, Z.; Sun, W.; Ning, G.; Wei, T.; Zhang, Q.; Zhang, R.; Zhi, L.; Wei, F. Advanced Asymmetric Supercapacitors Based on Ni(OH)2/Graphene and Porous Graphene Electrodes with High Energy Density. Adv. Funct. Mater. 2012, 22, 2632–2641. [Google Scholar] [CrossRef]
  14. Najib, S.; Erdem, E. Current progress achieved in novel materials for supercapacitor electrodes: Mini-review. Nanoscale Adv. 2019, 1, 2817–2827. [Google Scholar] [CrossRef] [Green Version]
  15. Dong, L.; Xu, C.; Li, Y.; Huang, Z.H.; Kang, F.; Yang, Q.H.; Zhao, X. Flexible electrodes and supercapacitors for wearable energy storage: A review by category. J. Mater. Chem. A 2016, 4, 4659–4685. [Google Scholar] [CrossRef]
  16. Hercule, K.M.; Wei, Q.; Khan, A.M.; Zhao, Y.; Tian, X.; Mai, L. Synergistic effect of hierarchical nanostructured MoO2/Co(OH)2 with largely enhanced pseudocapacitor cyclability. Nanoscale Lett. 2013, 13, 5685–5691. [Google Scholar] [CrossRef]
  17. Joshi, A.; Sahu, V.; Singh, G.; Sharma, R.K. Performance enhancement of a supercapacitor negative electrode based on loofah sponge derived oxygen-rich carbon through encapsulation of MoO3 nanoflowers. Sustain. Energy Fuels 2019, 3, 1248–1257. [Google Scholar] [CrossRef]
  18. Lu, X.F.; Huang, Z.X.; Tong, Y.X.; Li, G.R. Asymmetric supercapacitors with high energy density based on helical hierarchical porous NaxMnO2 and MoO2. Chem. Sci. 2016, 7, 510–517. [Google Scholar] [CrossRef] [Green Version]
  19. Chang, J.; Jin, M.; Yao, F.; Kim, T.H.; Le, V.T.; Yue, H.; Gunes, F.; Li, B.; Ghosh, A.; Xie, S.; et al. Asymmetric Supercapacitors Based on Graphene/MnO2 Nanospheres and Graphene/MoO3 Nanosheets with High Energy Density. Adv. Funct. Mater. 2013, 23, 5074–5083. [Google Scholar] [CrossRef]
  20. Yang, P.; Ding, Y.; Lin, Z.; Chen, Z.; Li, Y.; Qiang, P.; Ebrahimi, M.; Mai, W.; Wong, C.P.; Wang, Z.L. Low-cost high performance solid-state asymmetric supercapacitors based on MnO2 nanowires and Fe2O3 nanotubes. Nanoscale Lett. 2014, 14, 731–736. [Google Scholar] [CrossRef]
  21. Xie, Y.; Chen, Y.; Zhou, Y.; Unruh, K.M.; Xiao, J.Q. A negative working potential supercapacitor electrode consisting of a continuous nanoporous Fe-Ni network. Nanoscale 2016, 8, 11875–11881. [Google Scholar] [CrossRef] [PubMed]
  22. Xia, X.; Zhang, Y.; Chao, D.; Xiong, Q.; Fan, Z.; Tong, X.; Tu, J.; Zhang, H.; Fan, H.J. Tubular TiC fibre nanostructures as supercapacitor electrode materials with stable cycling life and wide-temperature performance. Energy Environ. Sci. 2015, 8, 1559–1568. [Google Scholar] [CrossRef]
  23. Sun, L.; Wang, X.; Wang, Y.; Xiao, D.; Cai, W.; Jing, Y.; Wang, Y.; Hu, F.; Zhang, Q. In-situ Functionalization of Metal Electrodes for Advanced Asymmetric Supercapacitors. Front. Chem. 2019, 7, 512. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Nikkhah, Z.; Karimipour, A.; Safaei, M.R.; Tehrani, P.F.; Goodarzi, M.; Dahari, M.; Wongwises, S. Forced convective heat transfer of water/functionalized multi-walled carbon nanotube nanofluids in a microchannel with oscillating heat flux and slip boundary condition. Int. Commun. Heat Mass Transf. 2015, 68, 69–77. [Google Scholar] [CrossRef]
  25. Moradikazerouni, A.; Hajizadeh, A.; Reza Safaei, M.; Afrand, M.; Yarmand, H.; Binti, N.W.; Zulkifli, M. Assessment of thermal conductivity enhancement of nano-antifreeze containing single-walled carbon nanotubes: Optimal artificial neural network and curve-fitting. Phys. A Stat. Mech. Appl. 2019, 521, 138–145. [Google Scholar] [CrossRef]
  26. Sun, J.; Li, W.; E, L.; Xu, Z.; Ma, C.; Wu, Z.; Liu, S. Ultralight carbon aerogel with tubular structures and N-containing sandwich-like wall from kapok fibers for supercapacitor electrode materials. J. Power Sources 2019, 438, 227030. [Google Scholar] [CrossRef]
  27. Wang, F.; Robert, R.; Chernova, N.A.; Pereira, N.; Omenya, F.; Badway, F.; Hua, X.; Ruotolo, M.; Zhang, R.; Wu, L.; et al. Conversion Reaction Mechanisms in Lithium Ion Batteries: Study of the Binary Metal Fluoride Electrodes. J. Am. Chem. Soc. 2011, 133, 18828–18836. [Google Scholar] [CrossRef]
  28. Wang, Y.; He, P.; Lei, W.; Dong, F.; Zhang, T. Novel FeMoO4/graphene composites based electrode materials for supercapacitors. Compos. Sci. Technol. 2014, 103, 16–21. [Google Scholar] [CrossRef]
  29. Xue, J.; Wu, T.; Dai, Y.; Xia, Y. Electrospinning and Electrospun Nanofibers: Methods, Materials, and Applications. Chem. Rev. 2019, 119, 5298–5415. [Google Scholar] [CrossRef]
  30. Lee, H.G.; Gopalan, A.I.; Lee, B.C.; Lee, K.P. Facile synthesis of functionalized graphene-palladium nanoparticle incorporated multicomponent TiO2 composite nanofibers. Mater. Chem. Phys. 2015, 154, 125–136. [Google Scholar] [CrossRef]
  31. Peng, S.; Li, L.; Hu, Y.; Srinivasan, M.; Cheng, F.; Chen, J.; Ramakrishna, S. Fabrication of Spinel One-Dimensional Architectures by Single-Spinneret Electrospinning for Energy Storage Applications. ACS Nanoscale 2015, 9, 1945–1954. [Google Scholar] [CrossRef] [PubMed]
  32. Gopalan, A.I.; Komathi, S.; Muthuchamy, N.; Lee, K.P.; Whitcombe, M.J.; Dhana, L.; Anand, G.S. Functionalized conjugated polymers for sensing and molecular imprinting applications. Progr. Polym. Sci. 2019, 88, 1–129. [Google Scholar]
  33. Ranjith, K.S.; Manivel, P.; Rajendrakumar, R.T.; Uyar, T. Multifunctional ZnO nanorod-reduced graphene oxide hybrids nanocomposites for effective water remediation: Effective sunlight driven degradation of organic dyes and rapid heavy metal adsorption. Chem. Eng. J. 2017, 325, 588–600. [Google Scholar] [CrossRef] [Green Version]
  34. Singh, R.; Kumar, M.; Khajuria, H.; Sharma, S.; Sheikh, H.N. Studies on hydrothermal synthesis of photolumniscent rare earth (Eu3+ & Tb3+) doped NG@FeMoO4 for enhanced visible light photodegradation of methylene blue dye. Solid State Sci. 2018, 76, 38–47. [Google Scholar]
  35. Nagaraju, D.H.; Wang, Q.; Beaujuge, P.; Alshareef, H.N. Two-dimensional heterostructures of V2O5 and reduced graphene oxide as electrodes for high energy density asymmetric supercapacitors. J. Mater. Chem. A 2014, 2, 17146–17152. [Google Scholar] [CrossRef]
  36. Pham, M.H.; Dinh, C.T.; Vuong, G.T.; Ta, N.D.; Do, T.O. Visible light-induced hydrogen generation using a hollow photocatalyst with two cocatalysts separated on two surface sides. Phys. Chem. Chem. Phys. 2014, 16, 5937–5941. [Google Scholar] [CrossRef]
  37. Zhang, Z.; Hu, C.; Hashim, M.; Chen, P.; Xiong, Y.; Zhang, C. Synthesis and magnetic property of FeMoO4 nanorods. Mater. Sci. Eng. B 2011, 176, 756–761. [Google Scholar] [CrossRef]
  38. Zhang, Z.; Li, W.; Ng, T.W.; Kang, W.; Lee, C.S.; Zhang, W. Iron(II) molybdate (FeMoO4) nanorods as a high-performance anode for lithium-ion batteries: Structural and chemical evolution upon cycling. J. Mater. Chem. A 2015, 3, 20527–20534. [Google Scholar] [CrossRef]
  39. Lin, L.; Song, X.; Chen, Y.; Rong, M.; Zhao, T.; Jiang, Y.; Wang, Y.; Chen, X. One-pot synthesis of highly greenish-yellow fluorescent nitrogen-doped graphene quantum dots for pyrophosphate sensing via competitive coordination with Eu3+ ions. Nanoscale 2015, 7, 15427–15433. [Google Scholar] [CrossRef]
  40. Gou, Y.; Liu, Q.; Liu, Z.; Asiri, A.M.; Sun, X.; Hu, J. FeMoO4 nanorod array: A highly active 3D anode for water oxidation under alkaline conditions. Inorg. Chem. Front. 2018, 5, 665–668. [Google Scholar] [CrossRef]
  41. Zhou, J.; Yu, L.; Sun, M.; Yang, S.; Ye, F.; He, J.; Hao, Z. Novel Synthesis of Birnessite-Type MnO2 Nanostructure for Water Treatment and Electrochemical Capacitor. Ind. Eng. Chem. Res. 2013, 52, 9586–9593. [Google Scholar] [CrossRef]
  42. Xu, H.; Hu, X.; Yang, H.; Sun, Y.; Hu, C.; Huang, Y. Flexible Asymmetric Micro-Supercapacitors Based on Bi2O3 and MnO2 Nanoflowers: Larger Areal Mass Promises Higher Energy Density. Adv. Energy Mater. 2015, 5, 1401882. [Google Scholar] [CrossRef]
  43. Huang, M.; Zhang, Y.; Li, F.; Zhang, L.; Wen, Z.; Liu, Q. Facile synthesis of hierarchical Co3O4@MnO2 core-shell arrays on Ni foam for asymmetric supercapacitors. J. Power Sources 2014, 252, 98–106. [Google Scholar] [CrossRef]
  44. Yu, D.; Zhang, Z.; Meng, Y.; Teng, Y.; Wu, Y.; Zhang, X.; Sun, Q.; Tong, W.; Zhao, X.; Liu, X. The synthesis of hierarchical ZnCo2O4@MnO2 core-shell nanosheet arrays on Ni foam for high-performance all-solid-state asymmetric supercapacitors. Inorg. Chem. Front. 2018, 5, 597–604. [Google Scholar] [CrossRef]
  45. Fu, W.; Wang, Y.; Han, W.; Zhang, Z.; Zha, H.; Xie, E. Construction of hierarchical ZnCo2O4@NixCo2x(OH)6x core/shell nanowire arrays for high-performance supercapacitors. J. Mater. Chem. A 2016, 4, 173–182. [Google Scholar] [CrossRef]
  46. Bhagwan, J.; Hussain, S.K.; Yu, J.S. Facile Hydrothermal Synthesis and Electrochemical Properties of CaMoO4 Nanoparticles for Aqueous Asymmetric Supercapacitors. ACS Sustain. Chem. Eng. 2019, 7, 12340–12350. [Google Scholar] [CrossRef]
  47. Liu, W.; Zhu, M.; Liu, J.; Li, X.; Liu, J. Flexible asymmetric supercapacitor with high energy density based on optimized MnO2 cathode and Fe2O3 anode. Chin. Chem. Lett. 2019, 30, 750–756. [Google Scholar] [CrossRef]
  48. Long, C.; Jiang, L.; Wei, T.; Yan, J.; Fan, Z. High-performance asymmetric supercapacitors with lithium intercalation reaction using metal oxide-based composites as electrode materials. J. Mater. Chem. A 2014, 2, 16678–16686. [Google Scholar] [CrossRef]
  49. Jin, W.H.; Cao, G.T.; Sun, J.Y. Hybrid supercapacitor based on MnO2 and columned FeOOH using Li2SO4 electrolyte solution. J. Power Sources 2008, 175, 686–691. [Google Scholar] [CrossRef]
  50. Kim, M.; Hwang, Y.; Kim, J. Graphene/MnO2-based composites reduced via different chemical agents for supercapacitors. J. Power Sources 2013, 239, 225–233. [Google Scholar] [CrossRef]
Scheme 1. Schematic illustration of the strategic growth of reduced graphene oxide (rGO)-functionalized bimetallic tubular nanofibers (NFs) (rGO–FeMoO4) via electrospinning followed by solution-based thermal functionalization.
Scheme 1. Schematic illustration of the strategic growth of reduced graphene oxide (rGO)-functionalized bimetallic tubular nanofibers (NFs) (rGO–FeMoO4) via electrospinning followed by solution-based thermal functionalization.
Nanomaterials 10 00289 sch001
Figure 1. SEM images of (a,b) the FeMoO4 solid NFs; (c,d) the FeMoO4 tubular NFs; (e,f) the rGO–FeMoO4 tubular NFs; TEM images (g,h) of the rGO–FeMoO4 NF (inset: the SAED pattern of rGO-FeMoO4); (i) high-resolution TEM image of rGO–FeMoO4 NF; (j) elemental mapping images of the hybrid NF.
Figure 1. SEM images of (a,b) the FeMoO4 solid NFs; (c,d) the FeMoO4 tubular NFs; (e,f) the rGO–FeMoO4 tubular NFs; TEM images (g,h) of the rGO–FeMoO4 NF (inset: the SAED pattern of rGO-FeMoO4); (i) high-resolution TEM image of rGO–FeMoO4 NF; (j) elemental mapping images of the hybrid NF.
Nanomaterials 10 00289 g001
Figure 2. Physical characterization of rGO–FeMoO4: (a) XRD pattern of FeMoO4-based NFs and the XPS spectra of the rGO–FeMoO4 NFs; (b) full survey spectrum; (c) Fe 2p; (d) Mo 3d; (e) C 1s (inset: N 1s spectra); (f) O 1s.
Figure 2. Physical characterization of rGO–FeMoO4: (a) XRD pattern of FeMoO4-based NFs and the XPS spectra of the rGO–FeMoO4 NFs; (b) full survey spectrum; (c) Fe 2p; (d) Mo 3d; (e) C 1s (inset: N 1s spectra); (f) O 1s.
Nanomaterials 10 00289 g002
Figure 3. Three-electrode electrochemical performance of the tubular FeMoO4, rGO and rGO–FeMoO4 electrodes in 1 M Na2SO4 aqueous solution: (a) CV curve for the tubular FeMoO4 and rGO–FeMoO4 at 10 mV s−1; (b) CV curves for rGO–FeMoO4 at various scan rates; (c) GCD profile, and (d) IR drop for rGO–FeMoO4 at various current densities; (e) rate performances for FeMoO4 and rGO–FeMoO4 at various current densities; (f) Nyquist plot of the rGO–FeMoO4 electrode; and (g) the cyclic stability and coulombic efficiency at 1 A g−1. The inset images in (g) present the GCD, CV, and EIS spectra before and after cycling performances.
Figure 3. Three-electrode electrochemical performance of the tubular FeMoO4, rGO and rGO–FeMoO4 electrodes in 1 M Na2SO4 aqueous solution: (a) CV curve for the tubular FeMoO4 and rGO–FeMoO4 at 10 mV s−1; (b) CV curves for rGO–FeMoO4 at various scan rates; (c) GCD profile, and (d) IR drop for rGO–FeMoO4 at various current densities; (e) rate performances for FeMoO4 and rGO–FeMoO4 at various current densities; (f) Nyquist plot of the rGO–FeMoO4 electrode; and (g) the cyclic stability and coulombic efficiency at 1 A g−1. The inset images in (g) present the GCD, CV, and EIS spectra before and after cycling performances.
Nanomaterials 10 00289 g003
Figure 4. (a,b) SEM (inset: low magnification view); (ce) TEM images, and (f) corresponding SAED pattern of the rGO–MnO2 NRs; (g) Powder XRD spectra, and (h) the XPS full scan survey spectra of the rGO–MnO2@CC. The inset in (h) presents the high-resolution Mn 2p of rGO–MnO2 NRs.
Figure 4. (a,b) SEM (inset: low magnification view); (ce) TEM images, and (f) corresponding SAED pattern of the rGO–MnO2 NRs; (g) Powder XRD spectra, and (h) the XPS full scan survey spectra of the rGO–MnO2@CC. The inset in (h) presents the high-resolution Mn 2p of rGO–MnO2 NRs.
Nanomaterials 10 00289 g004
Figure 5. (a) CV curve of the rGO, MnO2 and rGO–MnO2 free-standing electrodes measured at the scan rate of 50 mV s−1; (b) CV curves of the rGO–MnO2 electrodes at various scan rates (1–200 mV s−1); (c) GCD curves and (d) rate performances for the rGO–MnO2 electrode at various current densities; (e) Nyquist plot and (f) cycling stability of the rGO–MnO2 electrode.
Figure 5. (a) CV curve of the rGO, MnO2 and rGO–MnO2 free-standing electrodes measured at the scan rate of 50 mV s−1; (b) CV curves of the rGO–MnO2 electrodes at various scan rates (1–200 mV s−1); (c) GCD curves and (d) rate performances for the rGO–MnO2 electrode at various current densities; (e) Nyquist plot and (f) cycling stability of the rGO–MnO2 electrode.
Nanomaterials 10 00289 g005
Figure 6. (a) Schematic diagram of the constructed solid-state ASC device; (b) CV curves of the individual rGO–FeMoO4@CC and rGO–MnO2@CC electrodes at a scan rate of 30 mV s−1; (c) CV curves of the rGO–FeMoO4@CC/rGO–MnO2@CC within various potential windows at a scan rate of 50 mV s−1; (d) CV curves at various scan rates and (e) GCD curves at various current densities for the rGO–FeMoO4@CC/rGO–MnO2@CC ASC; (f) rate performances of FeMoO4@CC/rGO–MnO2@CC ASC; (g) comparative Nyquist plots before and after 5000 cycles; and (h) cyclic stability and coulombic efficiency of the rGO–FeMoO4@CC/rGO–MnO2@CC ASC at 4 A g−1.
Figure 6. (a) Schematic diagram of the constructed solid-state ASC device; (b) CV curves of the individual rGO–FeMoO4@CC and rGO–MnO2@CC electrodes at a scan rate of 30 mV s−1; (c) CV curves of the rGO–FeMoO4@CC/rGO–MnO2@CC within various potential windows at a scan rate of 50 mV s−1; (d) CV curves at various scan rates and (e) GCD curves at various current densities for the rGO–FeMoO4@CC/rGO–MnO2@CC ASC; (f) rate performances of FeMoO4@CC/rGO–MnO2@CC ASC; (g) comparative Nyquist plots before and after 5000 cycles; and (h) cyclic stability and coulombic efficiency of the rGO–FeMoO4@CC/rGO–MnO2@CC ASC at 4 A g−1.
Nanomaterials 10 00289 g006
Figure 7. (a) Ragone plot comparing the rGO–MnO2@CC/rGO–FeMoO4@CC ASC device with previously reported results; (b) photographic image of LEDs lit up by three ASCs connected in series; (c) CV curves of the all-solid-state ASC under various bending angles, and (d) photographic images of the all-solid-state ASC bent at various positions.
Figure 7. (a) Ragone plot comparing the rGO–MnO2@CC/rGO–FeMoO4@CC ASC device with previously reported results; (b) photographic image of LEDs lit up by three ASCs connected in series; (c) CV curves of the all-solid-state ASC under various bending angles, and (d) photographic images of the all-solid-state ASC bent at various positions.
Nanomaterials 10 00289 g007

Share and Cite

MDPI and ACS Style

Ranjith, K.S.; Raju, G.S.R.; R. Chodankar, N.; Ghoreishian, S.M.; Kwak, C.H.; Huh, Y.S.; Han, Y.-K. Electroactive Ultra-Thin rGO-Enriched FeMoO4 Nanotubes and MnO2 Nanorods as Electrodes for High-Performance All-Solid-State Asymmetric Supercapacitors. Nanomaterials 2020, 10, 289. https://doi.org/10.3390/nano10020289

AMA Style

Ranjith KS, Raju GSR, R. Chodankar N, Ghoreishian SM, Kwak CH, Huh YS, Han Y-K. Electroactive Ultra-Thin rGO-Enriched FeMoO4 Nanotubes and MnO2 Nanorods as Electrodes for High-Performance All-Solid-State Asymmetric Supercapacitors. Nanomaterials. 2020; 10(2):289. https://doi.org/10.3390/nano10020289

Chicago/Turabian Style

Ranjith, Kugalur Shanmugam, Ganji Seeta Rama Raju, Nilesh R. Chodankar, Seyed Majid Ghoreishian, Cheol Hwan Kwak, Yun Suk Huh, and Young-Kyu Han. 2020. "Electroactive Ultra-Thin rGO-Enriched FeMoO4 Nanotubes and MnO2 Nanorods as Electrodes for High-Performance All-Solid-State Asymmetric Supercapacitors" Nanomaterials 10, no. 2: 289. https://doi.org/10.3390/nano10020289

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