Next Article in Journal
Insecticide Resistance and Its Management in Two Invasive Cryptic Species of Bemisia tabaci in China
Next Article in Special Issue
Nanoparticle-Polymer Surface Functionalizations for Capacitive Energy Storage: Experimental Comparison to First Principles Simulations
Previous Article in Journal
Diagnostic Accuracy of Leucine-Rich α-2-Glycoprotein 1 as a Non-Invasive Salivary Biomarker in Pediatric Appendicitis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Performance of WO3/SnO2 Nanocomposite Electrodes with Redox-Active Electrolytes for Supercapacitors

by
Tamiru Deressa Morka
and
Masaki Ujihara
*
Graduate Institute of Applied Science and Technology, National Taiwan University of Science and Technology, 43 Keelung Road, Taipei 10607, Taiwan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(7), 6045; https://doi.org/10.3390/ijms24076045
Submission received: 9 February 2023 / Revised: 1 March 2023 / Accepted: 6 March 2023 / Published: 23 March 2023
(This article belongs to the Special Issue Advances in Nanostructured Materials for Energy Storage Applications)

Abstract

:
For effective supercapacitors, we developed a process involving chemical bath deposition, followed by electrochemical deposition and calcination, to produce WO3/SnO2 nanocomposite electrodes. In aqueous solutions, the hexagonal WO3 microspheres were first chemically deposited on a carbon cloth, and then tin oxides were uniformly electrodeposited. The synthesized WO3/SnO2 nanocomposite was characterized by XRD, XPS, SEM, and EDX techniques. Electrochemical properties of the WO3/SnO2 nanocomposite were analyzed by cyclic voltammetry, galvanostatic charge-discharge tests, and electrochemical impedance spectroscopy in an aqueous solution of Na2SO4 with/without the redox-active electrolyte K3Fe(CN)6. K3Fe(CN)6 exhibited a synergetic effect on the electrochemical performance of the WO3/SnO2 nanocomposite electrode, with a specific capacitance of 640 F/g at a scan rate of 5 mV/s, while that without K3Fe(CN)6 was 530 F/g. The WO3/SnO2 nanocomposite catalyzed the redox reactions of [Fe(CN)6]3/[Fe(CN)6]4− ions, and the [Fe(CN)6]3−/[Fe(CN)6]4− ions also promoted redox reactions of the WO3/SnO2 nanocomposite. A symmetrical configuration of the nanocomposite electrodes provided good cycling stability (coulombic efficiency of 99.6% over 2000 cycles) and satisfied both energy density (60 Whkg−1) and power density (540 Wkg−1) requirements. Thus, the WO3/SnO2 nanocomposite prepared by this simple process is a promising component for a hybrid pseudocapacitor system with a redox-flow battery mechanism.

Graphical Abstract

1. Introduction

Clean and renewable energy is the basic criterion for sustainable development of the global economy and human lives throughout the world [1]. The massive increase in fossil fuel usage has brought two significant problems: the first is rapidly reduced fossil fuel reserves, and the second is environmental degradation, such as air/water pollution and climate change. Considering these issues, the need for renewable energy and new technology for energy storage are currently the top concerns around the world. In response to rising ecological concern and modern civilization, new environmentally friendly and cost-effective energy storage devices with excellent performance are required for customer products, such as hybrid automobiles, electronic vehicles, mobile phones, and laptop computers [2,3]. Supercapacitors, batteries, and fuel cells are the most common electrochemical energy storage devices used for this purpose. Among them, supercapacitors are emphasized because of their excellent power densities, fast charge/discharge, long-term cycling life, cost-effectiveness, good reversibility, and wide working temperature range compared to those of batteries; however, in terms of commercial application, supercapacitors are behind batteries [4,5,6]. The fundamental issue of supercapacitors is their lower energy density, especially when compared to batteries.
To solve this problem of supercapacitors, three different energy storage mechanisms have been developed: electric double-layer capacitors (EDLCs), pseudocapacitor, and battery-type capacitors [7]. Their energy storage mechanisms are explained as follows: the electric double-layer capacitor stores energy at the junction of the electrode and electrolyte by ion adsorption/desorption, the pseudocapacitor electrode stores energy on the surface of the electrode via Faradaic processes (using a fast and reversible redox process near the surface), and the battery-type capacitor has a specific potential window for the Faradaic processes used to enhance the capacitance [8]. During these electrochemical reactions, the amount of charge stored in the electrode is proportional to the total available surface of the active materials in the electrode. Therefore, porous materials with large specific surface areas are preferable for supercapacitors [1].
As active materials for pseudocapacitors, transition metal oxides (TMOs) and conducting polymers are widely used [9]. Transition metal oxides are known for high specific capacitance without sacrificing power density [10]. Among TMOs, RuO2, Co3O4, ZnO, MoO3, Fe2O3 MnO2, NiO, SnO2, WO3, and others have been investigated as active materials due to their appealing characteristics, such as high theoretical capacitance and low environmental impact [2,8,11,12,13]. Furthermore, composite TMOs have been designed for a pseudocapacitor with expectations of synergistic effects [6,7]. The TMOs in the composite electrode can produce unique structures with large surface areas, and the mixed valence metals can provide an extra valance state for the electrochemical process, which results in improved performance of the supercapacitor [3,5].
Recently, tungsten trioxide (WO3) has been considered a promising candidate for pseudocapacitors due to its numerous intrinsic features, such as high theoretical capacitance, high corrosion resistance, good chemical stability, low cost, and environmental friendliness [14], and a variety of composites with other TMOs, such as MnO2-WO3 [15], Ni-WO3 [16], Co-WO3 [17], TiO2/WO3 [18], WO3/Se [2], and V2O5/WO3 [19], have been examined. Moreover, SnO2 has attracted much attention as an electrode for energy storage or conversion due to its low cost, nontoxicity, high electrochemical activity, and chemical stability [20]. For successful insertion of SnO2 into WO3, excellent electrochemical supercapacitor performance is required [21,22]. Hence, motivated by the above considerations, we investigated the WO3/SnO2 nanocomposite as an active electrode material for supercapacitors.
Another approach to enhancing the capacity of the supercapacitor is to use a redox-active electrolyte. The effects of additive electrolytes, such as K3Fe(CN)6 [23,24], KI [25], and p-aminophenol [26], have been reported for aqueous systems. The basic role of these additives is to improve charge transport in the electrolyte solution, adjust the chemical state of the working electrode surface, and enhance Faradaic reactions. Synergistic effects in electrochemical processes, both in the active materials and electrolytes, can also be expected to enhance the efficiencies of redox reactions and improve the performance of supercapacitors [23]. The electrolyte additive potassium ferricyanide, K3Fe(CN)6, enables the electrode of the supercapacitor to achieve fast charging, slow discharge, and good cycling stability [27]. In this work, we investigated the preparation of a WO3/SnO2 nanocomposite through chemical bath deposition (CBD) and electrodeposition on a carbon cloth substrate, and analyzed its supercapacitor performance by using K3Fe(CN)6 added into a 1 M Na2SO4 aqueous electrolyte [28]. To the best of our knowledge, this is the first trial to synthesize a WO3/SnO2 nanocomposite and to determine its electrochemical supercapacitance performance with the redox-active electrolyte K3Fe (CN)6 with Na2SO4, and the system demonstrated a specific capacitance of 640 F/g. A symmetric supercapacitor using a WO3/SnO2 nanocomposite and an additive electrolyte was further assembled to demonstrate the potential of this system for efficient energy storage.
The reaction mechanisms for formation of WO3 were as follows (Equations (1) and (2)) [1,29]:
Na 2 WO 4 + 2 HCl + nH 2 O     H 2 WO 4 · nH 2 O + 2 HO 4 · nH 2 O
H 2 WO 4 · nH 2 O     WO 3 + ( n + 1 ) H 2 O
Crystals of tungsten oxide (WO3) are polymorphic and often adopt a hexagonal system [1,15]. Various synthetic aspects, such as the reaction temperature, process duration, precursor type, and chemical agent (e.g., (NH4)2SO4), can influence the structure and morphology. In comparison to other structures, this structure provides large tunnels for ion penetration, which improves the electrochemical performance [13]. Thus, WO3 is usually synthesized as a hexagonal system using reagents containing sulfate ions; in particular, ammonium sulfate is preferred. The NH4+ and SO42− ions act as stabilizing and capping agents, respectively, during the process. The radius of the NH4 ion is larger than those of alkali metal ions, and these differences in size influence the penetration of numerous ions, which implies that many SO42− ions can readily be adsorbed on the surface parallel to the WO3 axis [13,30]. The morphology and crystal structure of WO3 are also affected by the annealing temperature. The orthorhombic phase changes into an anhydrous hexagonal phase at 400 °C and turns into stable monoclinic WO3 with aggregation above 500 °C [31]. Therefore, annealing in this work was done at 400 °C for 2 h.

2. Results

2.1. Characterization of WO3, SnO2, and WO3/SnO2 Nanocomposite Thin Films

The structural characteristics and crystal phases of the WO3, SnO2, and WO3/SnO2 nanocomposites were analyzed by XRD (Figure 1). The XRD peaks for WO3 were assigned to the hexagonal system, except for the one marked with an asterisk; this came from the carbon cloth substrate (Figure 1a). The diffraction peaks at 2θ = 14.25°, 23.28°, 28.2°, 34.8°, 37.28°, 44.30°, 50.28°, 52.40°, and 56.03° were assigned to the (100), (110), (200), (112), (202), (212), (004), (220), and (204) planes, respectively (JCPDS 85-2459). The sharp and intense peaks suggested the highly crystalline nature of the WO3.
The XRD pattern for the synthesized SnO2 showed sharp diffraction peaks at 30.7°, 32.1°, 43.9°, 45.1°, and 55.5°, which corresponded to the (221), (101), (111), (210), and (220) planes of the tetragonal phase (JCPDS 41-1445), respectively, except for the one marked with an asterisk, which came from the carbon cloth substrate (Figure 1b) [32]. Figure 1c shows the diffraction peaks for the WO3/SnO2 nanocomposite at 2θ = 14.25°, 23.28°, 28.20°, 30.71°, 32.10°, 34.80°, 37.28°, 43.20°, 44.30°, 45.10°, 50.28°, 52.40°, 55.20°, and 56.03°, which were assigned to the (100), (110), (200), (112), (202), (212), (004), (220), and (204) planes of WO3 and (221), (101), (111), (210), and (220) planes of SnO2, respectively. The diffraction peaks, denoted with diamond symbols, were indexed to the tetragonal structure of SnO2 [33]; similarly, the other peaks were associated with the hexagonal crystal phase of WO3 [1]. Therefore, this diffraction pattern was composed of characteristic peaks for both WO3 and SnO2, as observed in (Figure 1a,b), which confirmed successful preparation of the WO3/SnO2 nanocomposite. No new significant peak was observed, and thus, the WO3 and SnO2 components were separately crystallized in the nanocomposite. The presence of sharp diffraction peaks suggested the highly crystalline nature of the materials deposited on the carbon cloth substrate. The average sizes of the WO3 and SnO2 crystallites in the WO3/SnO2 nanocomposite were determined by the Scherrer equation, as shown below in (Equation (3)) [6,34,35,36]:
D = K/βcosθ
where D is the crystallite size (nm), β is the full width half maximum (FWHM, radians) of the peak corresponding to the plane, K = 0.9 (Sherrer constant), λ = 0.15406 nm (wavelength of the X-ray source), and θ is obtained from the 2θ value corresponding to the diffraction peak [35]. These findings showed that the average crystallite sizes for the WO3 synthesized in the nanocomposite were 7.5 nm, with diffraction peaks of 2θ = 23.28° and 28.2° for the corresponding (110) and (200) planes, respectively. The average crystallite size of the SnO2 synthesized in the nanocomposite was estimated to be 3.9 nm, with diffraction peaks at 2θ = 32.10°, with a corresponding (101) plane [37,38].
The chemical compositions and surfaces of the WO3, SnO2, and WO3/SnO2 nanocomposites were further investigated using XPS (Figure 2 and Table 1). The survey spectrum for the WO3/SnO2 nanocomposite showed the coexistence of W, Sn, O, and C elements, and no other impurity peak was significantly detected (Figure 2a).
The high-resolution W 4f spectrum for WO3 was deconvoluted into two peaks (Figure 2b). The peaks at 36.08 eV and 38.21 eV were assigned to W 4f7/2 and W 4f5/2 binding energies, respectively, corresponding to the W6+ oxidation state. The separation between the W 4f7/2 and W 4f5/2 peaks was 2.13 eV, which was in good agreement with an earlier report [1]. Figure 2c shows the O 1s peak at 531.1 eV, which was assigned to the lattice oxygens of W–O bonds in WO3. The peak at 532.01 eV was due to an oxygen deficiency, the peak at 532.7 eV was assigned to chemisorbed oxygen, and the peak at 533.6 eV was attributed to water and hydroxide. Figure 3d shows the high-resolution Sn 3d spectrum of pure SnO2, which was deconvoluted into two peaks with binding energies of 487.6 eV and 496.05 eV. These peaks were assigned to the Sn 3d5/2 and Sn 3d3/2 binding energies, respectively, for the oxidation state Sn4+. The splitting between the Sn 3d5/2 and Sn 3d3/2 peaks was 8.45 eV, which was in good agreement with a previous study [4]. As displayed in (Figure 2e), the O 1s spectrum was deconvoluted into three peaks at 531.5 eV for lattice SnO2, chemisorbed oxygen at 532.7 eV, and water and hydroxide at 534.3 eV. In the WO3/SnO2 nanocomposite, the W 4f spectrum showed two peaks at 35.74 eV and 38.08 eV, which corresponded to W 4f7/2 and W 4f5/2 binding energies, respectively, and an oxidation state of W+6 (Figure 2f). The spin–orbit separation of W 4f7/2 and W 4f5/2 was 2.24 eV, which agreed with previous reports [39]. The difference in the W 4f doublet separation for the WO3 and WO3/SnO2 nanocomposites (2.13 eV and 2.24 eV, respectively) was also reported in an earlier study [2]: the W 4f7/2 peak was shifted by −0.24 eV, and the W 4f5/2 peak was shifted by −0.13 eV, which suggested an interaction between WO3 and SnO2 (Table 1). Moreover, the Sn 3d spectrum of the WO3/SnO2 composite had two peaks at 487.69 eV and 496.12 eV, which corresponded to the characteristic Sn 3d5/2 and Sn 3d3/2 binding energies, respectively. The spin–orbit separation between the Sn 3d5/2 and Sn 3d3/2 peaks was 8.43 eV (Figure 2g), which agreed with earlier reports [40]. Comparing pure SnO2 and WO3/SnO2, the Sn 3d5/2 and Sn 3d3/2 peaks were slightly shifted by 0.02 eV, which could have been due to an interaction between WO3 and SnO2. The O 1s spectrum was deconvoluted into four peaks (Figure 2h) at 531.09 eV, 531.76 eV, 532.8 eV, and 534.3 eV. The main peak at 531.76 eV was attributed to lattice oxygens in the metal oxides WO3 and SnO2 [33,37,40], the peak at 531.09 eV was assigned to oxygen defects or vacancies, the peak at 532.8 eV was related to chemisorbed oxygen species [41], and the peak at 534.3 eV was attributed to water and hydroxides on the surface of the WO3/SnO2 nanocomposite, as reported previously [6,32,40]. The binding energies of the lattice oxygens and oxygen vacancies in the WO3/SnO2 nanocomposite were shifted from those of the single components by −0.34 eV and 0.25 eV, respectively, which suggested an interaction between WO3 and SnO2 in the nanocomposite. The binding energy for absorbed water/hydroxides in the nanocomposite was similar to that of SnO2, which suggested that the surface of the nanocomposite was covered by SnO2.
The morphologies of the materials were observed by SEM. WO3 was composed of aggregated monodisperse microspheres (diameter: ~1 μm) with rough surfaces (Figure 3a,b). The spaces between these microspheres would allow electrolyte ions to diffuse into the porous nanostructure [15,42]. The roughness of the microsphere surface increased the number of active sites available for electrochemical reactions. Moreover, the SnO2 exhibited nanorod structures with ~60 nm widths and ~200 nm lengths (Figure 3c,d). The applied potential and the deposition time are crucial for successful formation of SnO2 nanorods [43,44]. In this work, SnO2 nanorods were synthesized from SnCl2 on a cathode (−0.9 V vs. Ag/AgCl), and the formation of SnO2 crystals can be attributed to electrochemical deposition of Sn(OH)2, followed by air oxidation during calcination, as shown below (Equations (4)–(7)) [45]:
SnCl 2 + 2 OH     Sn ( OH ) 2 + 2 Cl
Sn ( OH ) 2     SnO + H 2 O
SnO + ½   O 2     SnO 2
Sn ( OH ) 2 + ½   O 2 + H 2 O     Sn ( OH ) 4     SnO 2 + 2 H 2 O
The SEM images in (Figure 3e,f) indicate the surface morphology of the WO3/SnO2 nanocomposite. Thin and wrinkled films were uniformly formed on the substrate. These films seemed to cover microspheres, which could be the WO3 first deposited on the carbon cloth substrate. This morphology suggested that SnO2 could not form coarse nanorods on the rough surfaces of WO3, so smaller SnO2 crystals formed a network on the surface [46]. In the cross-sectional views (Figure 3g,h), the surface of WO3 was covered by nanorods of SnO2 [47]. These small crystals and the fine nanorods provide a large surface area, which could enhance the electrochemical performance of this nanocomposite [32].
The elemental compositions of the synthesized materials were measured with EDS (Figure 4). In the pure WO3 microspheres, the atomic ratio of W to O was 1:3.09, which agreed with the theoretical ratio of 1:3. In the SnO2 nanorods, the atomic ratio of Sn to O was 1:1.58, which was lower than the theoretical ratio of 1:2, suggesting that unreacted Sn2+ provided defects or oxygen vacancies in the crystal lattice. The WO3/SnO2 nanocomposite consisted of W, Sn, and O, with atomic percentages of W (18.45%), Sn (6.31%), and O (75.24%). The Sn content was significantly lower than the W content, which was supported by the SEM observation (Figure 3e,f). The elemental maps (Figure 4d–g) showed similar patterns for the components, which suggests homogeneous coverage of SnO2 on the WO3 microspheres.

2.2. Electrochemical Measurements

2.2.1. Electrochemical Measurements without a Redox-Active Electrolyte

The electrochemical behaviors of the synthesized materials were analyzed with CV and GCD measurements using the non-redox electrolyte Na2SO4 (Figure 5 and Figure 6).
Figure 5a shows the CV curves for WO3 in the potential window −0.2 to 0.7 V vs. Ag/AgCl with various scan rates of 5–50 mV/s. These CV curves demonstrated nonrectangular shapes and medium oxidation peaks at approximately 0.1 V, which indicated a typical combination of pseudocapacitive behavior and electrical double-layer capacitance [48,49]. The oxidation peak clearly increased as the scan rate was increased, indicating the high reactivity of the WO3 electrode [50,51]. The specific capacitance was determined from the CV curves by using (Equation (8)) [9,52,53,54]:
Ccv ( Fg 1 ) = 1 m ν ( V f V i ) V i V f i d V
where Ccv is the gravimetric specific capacitance (F/g) calculated from the measurement, i is the voltammetric current, m is the mass of the active material (mg), ν is the scan rate (mV/s), Vf and Vi are the bounds of the potential window. The Ccv of the WO3 electrode was 240, 180, 150, 120, 98, and 84 F/g at scan rates of 5, 10, 20, 30, 40, and 50 mV−1, respectively.
Under the same conditions, the CV curves for SnO2 displayed quasi-rectangular shapes, with redox peaks at 0.54–0.69 V (oxidative) and 0.12–0.37 V (reductive), which indicated the more pseudocapacitive nature of the SnO2 electrode (Figure 5b). As the scan rate was increased, the CV curves for SnO2 exhibited more rectangular shapes, and the difference between the redox peaks increased, which suggested relatively slow redox reactions occurring on the surfaces of the SnO2 nanorods. The specific capacitance values for the SnO2 electrodes were calculated from the CV curves and were 140, 120, 90, 60, 50, and 45 F/g at scan rates of 5, 10, 20, 30, 40, and 50 mV/s, respectively, as shown in (Figure 5e). The CV curves for the WO3/SnO2 nanocomposite showed no significant redox peak (Figure 5c), which suggested the complementary work of the two components and a synergistic effect arising from interactions between WO3 and SnO2 (see Figure 2 and Table 1). This WO3/SnO2 nanocomposite electrode generated greater areas under the CV curves, which indicated a higher specific capacitance than the single-component electrodes (WO3 and SnO2) (Figure 5e). The specific capacitances of the WO3/SnO2 nanocomposite were calculated as 530, 410, 280, 210, 160, and 150 Fg−1 at scan rates of 5, 10, 20, 30, 40, and 50 mV s−1, respectively. Comparing the CV curves and the specific capacitances of the three electrodes, the WO3/SnO2 nanocomposite electrode had a higher specific capacitance, especially at low scan rates (Figure 5d,e). This enhanced performance for the nanocomposite electrode could be explained by activation of the SnO2 via the synergistic effect and the smaller crystals of SnO2 formed on WO3 (Figure 4). The small SnO2 crystals could have enhanced the surface area of the electrode and shortened the electron pathway from the surface of SnO2 to WO3 [55].
Figure 6a illustrates the GCD curves of the WO3 electrode in the potential window—0.2 to 0.7 V vs. Ag/AgCl at various current densities. The nonlinear nature of the charge/discharge processes confirmed the pseudocapacitive nature and the high activity at low potential (<0.1 V). The specific capacitance was calculated from the GCD curves and Equation (9) [9,53,54]:
C G ( Fg 1 ) = Ig d t / m Δ V ( t )
where CG is the gravimetric capacitance (F/g) from the GCD measurement, Ig is the given current, d t is the discharging time (s), and Δ V ( t ) is the potential window as a function of t. The CG of WO3 electrode was calculated to be 120 Fg−1 at a current density of 1.25 Ag−1. The SnO2 electrode exhibited two steps in the charge process and one step in the discharge process (Figure 6b). At 1.25 Ag−1, the specific capacitance of the SnO2 electrode was 54 Fg−1. Figure 6c shows the GCD curves for the WO3/SnO2 nanocomposite electrode at different current densities. The shapes of the GCD curves were more triangular, which indicated that the nanocomposite electrode exhibited better pseudocapacitive nature than the others, as was also suggested by the CV measurements (Figure 6). The specific capacitance of the WO3/SnO2 nanocomposite was determined to be 180, 120, 80, and 68 Fg−1 at the various current densities 1.25, 1.88, 2.82, and 5.00 Ag−1, respectively. Figure 6d provides a comparison of the GCD curves for the SnO2, WO3, and WO3/SnO2 electrodes at the same current density of 1.25 Ag−1. Figure 6e shows the EIS results for the WO3, SnO2, and WO3/SnO2 nanocomposite electrodes, and the magnified EIS data for the electrode are shown in (Figure 6f). The charge transfer resistance values (Rct) calculated for the WO3, SnO2, and WO3/SnO2 nanocomposite electrodes were 2.12 Ω, 1.94 Ω, and 1.72 Ω, respectively. The lower Rct of the WO3/SnO2 nanocomposite electrode indicated rapid charge transfer in the nanocomposite. The resistance of the electrode–electrolyte interference (Rs), which corresponds to the x-axis intercept in the high-frequency region, was calculated to be 2.37 Ω, 2.32 Ω, and 1.87 Ω for the WO3, SnO2, and WO3/SnO2 nanocomposite electrodes, respectively. The low Rs suggested that the WO3/SnO2 nanocomposite contacted the electrolytes with large surface areas because of its small grain sizes (Figure 4).

2.2.2. Effects of Redox-Active Electrolyte

Electrochemical analyses of the WO3, SnO2, and WO3/SnO2 electrodes were performed in the presence of the redox additive 0.2 M K3Fe (CN)6 in the 1 M Na2SO4 aqueous electrolyte. First, CV measurements were conducted with the carbon cloth, WO3, SnO2, and WO3/SnO2 nanocomposite electrodes (Figure 7a).
While the carbon cloth was almost inactive, the CV curves for the metal oxide electrodes exhibited quasi-rectangular shapes, with a pair of redox peaks from [Fe (CN)6]3−/[Fe (CN)6]4− at 0.35–0.43 V and 0.18–0.26 V. The WO3/SnO2 nanocomposite electrode had the largest area at a scan rate of 50 mV/s, which indicated the highest specific capacitance. The WO3/SnO2 nanocomposite electrode also resulted in higher intensities of redox peaks than the other electrodes, which indicated that this electrode exhibited the highest activity for the redox reactions of [Fe(CN)6]3−/[Fe(CN)6]4−. Moreover, considering the high peak intensities, the difference between the oxidation and reduction potentials was smaller than those of the other electrodes. The reduction peak for [Fe(CN)6]3− with the WO3/SnO2 nanocomposite electrode was at the highest potential among the electrodes, while the oxidation potential for [Fe(CN)6]4− with the WO3/SnO2 nanocomposite electrode was higher than the others. These behaviors suggested that the WO3/SnO2 nanocomposite catalyzed the redox reactions of [Fe(CN)6]3−/[Fe(CN)6]4−, which is preferable for energy storage because of the low overpotential. The catalytic activity of the WO3/SnO2 nanocomposite could be due to interactions between WO3 and SnO2, as suggested by the XPS measurements (Figure 2 and Table 1). Therefore, the CV curves generated with two sets of conditions were compared, i.e., with K3Fe (CN)6 and without K3Fe (CN)6 (Figure 7b). With K3Fe (CN)6, the areas of the CV curves were larger, which indicated that the redox additive increased the charge storage capacity. Using the CV curves generated at different scan rates (Figure 7c–e), the specific capacitance of the electrode in the redox-active electrolyte was calculated. At scan rates of 5, 10, 20, 30, 40, and 50 mV/s, the respective specific capacitance values were 440, 360, 280, 240, 190, and 150 F/g for the WO3 electrode; 310, 220, 150, 130, 120, and 110 F/g for the SnO2 electrode; and 640, 520, 340, 330, 310, and 290 F/g for the WO3/SnO2 nanocomposite electrode. Moreover, the addition of K3Fe(CN)6 resulted in more pseudocapacitive (more rectangular CV curves) for the WO3 and WO3/SnO2 nanocomposite electrodes. In details, weak and broad reduction bands newly appeared at −0.1–0.0 V for the WO3 electrode and 0.0–0.1 V for the WO3/SnO2 electrode in the presence of K3Fe(CN)6 (Figure 7b,c,e). The pseudocapacitive behavior suggested that K3Fe(CN)6 provided the additional redox reactions of the [Fe(CN)6]3−/[Fe(CN)6]4− couple to increase the specific capacitance of these electrodes [56,57], and also promoted the redox reactions of the WO3 component. That is, these metal oxides (WO3 and WO3/SnO2 nanocomposites) transferred electrons to the [Fe(CN)6]3−/[Fe(CN)6]4− pair in the electrolyte solution via their own redox reactions. The plots of the specific capacitance for various electrodes against the different scan rates showed that only the WO3/SnO2 nanocomposite electrode exhibited a larger enhancement of the specific capacitance by K3Fe(CN)6 at the high scan rate versus the low scan rate (Figure 7f). This enhancement at the high scan rate was due to the catalytic activity of the WO3/SnO2 nanocomposite in facilitating the redox reactions of [Fe(CN)6]3−/[Fe(CN)6]4−, as mentioned above.
The kinetics of the energy storage process were analyzed from the perspectives of peak current and scan rate (Figure 7g). In general, the relationship between the peak current and the scan rate can be described as in (Equation (10)) [58,59]:
Ip = aνb
where Ip is the peak oxidation current (mA), ν is the scan rate (mV/s), and both a and b are adjustable parameters. The b value is obtained from the slope of a plot of log Ip versus log ν. When the charge storage process is a surface-controlled capacitive process (ion adsorption/desorption), the value of b is 1. On the other hand, if diffusion is the rate-controlling process, b is 0.5 [60]. With and without the K3Fe(CN)6 electrolyte, the b values of the WO3/SnO2 nanocomposite electrode were 0.55 and 0.71, respectively. These b values suggested that addition of the K3Fe(CN)6 electrolyte favored diffusion control of the energy storage process because the approach of the redox-active species enabled charge transfer at the electrode surface. Moreover, the b values for the WO3 and SnO2 electrodes were 0.62 and 0.75, respectively. These b values were higher than that of the WO3/SnO2 nanocomposite electrode, which indicated that these single-component electrodes were less affected by diffusion of the redox-active species.
The GCD measurements made in the presence of the K3Fe(CN)6 electrolyte were not adequate to calculate the specific capacitance values of the synthesized electrodes because the discharge process took longer than the charge time (see Figure S1 in Supporting Information). This extension of the discharge process could be due to reduction of [Fe(CN)6]3− to [Fe(CN)6]4−, which could be used for the redox-flow battery system. Therefore, GCD measurements were conducted for stability testing (Figure 7h). The coulombic efficiency of the electrode was calculated with Equation (11):
η = td/tc × 100
where η is the coulombic efficiency, td is the discharging time, and tc is the charging time (s) [9]. After 2000 GCD cycles at a current density of 2.5 A/g in the presence of K3Fe(CN)6 electrolyte, the WO3/SnO2 nanocomposite electrode demonstrated good cycling stability, capacitance retention of 94.7%, and a coulombic efficiency of 99.9%. The inset graph presents the GCD curves for the first 10 and the last 10 cycles, and the shapes of the curves remained triangular, which indicated the pseudocapacitive nature and good capacitive behavior even after 2000 cycles. For comparison, the WO3 electrode exhibited lower capacitance retention (90.4%) under the same conditions. This difference indicated that K3Fe(CN)6 improved the stability of the electrode, not just the pseudocapacitor. EIS analyses of the WO3, SnO2, and WO3/SnO2 nanocomposite electrodes were carried out to determine the effects of K3Fe (CN)6 on the conductivities of these electrodes. In the electrolyte system containing 1 M Na2SO4 with 0.02 M K3Fe(CN)6, the Rct values of the WO3, SnO2, and WO3/SnO2 nanocomposite electrodes were 1.71 Ω, 1.69 Ω, and 1.56 Ω, respectively, and the Rs values of these electrodes were 2.78 Ω, 2.13 Ω, and 1.65 Ω, respectively (Table 2).
The WO3/SnO2 nanocomposite electrode exhibited a lower Rct than the other electrodes in the electrolyte systems with/without K3Fe(CN)6. For all electrodes used in this study, K3Fe(CN)6 drastically decreased Rct, which indicated that K3Fe(CN)6 facilitated charge transfer between the active materials. That is, K3Fe(CN)6 activated the electrical behaviors of the metal oxides used in this study. The Rs values also showed improvement, except for that of the WO3 electrode, which was attributed to an additional redox reaction occurring between the electrode surface and the electrolyte, in addition to the increased ion concentration of the electrolyte solution.

2.2.3. Electrochemical Performance of Symmetric Systems

For practical application of the WO3/SnO2 nanocomposite electrode, a symmetric supercapacitor system with identical electrodes was constructed with an aqueous electrolyte containing 1 M NaSO4 and 0.02 M K3Fe(CN)6. The total mass of the WO3/SnO2 nanocomposite on both electrodes was ~3.4 mg (1.7 mg for each). To confirm the optimal voltage window of the device, CV curves were measured over different potential ranges (from 1.0 to 1.8 V) at a scan rate of 50 mVs−1, as shown in (Figure 8a). The CV profile showed that the working potential was extended to 1.8 V, and water hydrolysis occurred at 1.9 V. As the potential range was increased, the shapes of the CV curves changed from quasi-rectangular (up to 1.4 V) to rectangular with redox peaks (higher than 1.6 V). This change indicated that a potential range larger than 1.6 V converted the pseudocapacitive mechanism to an energy storage mechanism, including the redox reactions of [Fe(CN)6]3−/[Fe(CN)6]4− ions.
The CV curves of the symmetric WO3/SnO2 nanocomposite electrode systems with and without K3Fe(CN)6 showed nearly rectangular shapes, which indicated pseudocapacitive behavior, even at a high scan rate of 50 mV/s (Figure 8b,c). Comparing the two systems, the CV profile of the WO3/SnO2 nanocomposite electrode with K3Fe(CN)6 had a greater area than that without K3Fe(CN)6, as expected from the half-cell experiment above (Figure 7b). GCD measurements of the symmetric WO3/SnO2 nanocomposite electrodes with and without K3Fe (CN)6 were also performed with various current densities (Figure 8e,f), and the GCD curves generated at 1.25 A/g were compared (Figure 8g). The IR drop in the initial phase of discharge was decreased in the presence of K3Fe(CN)6 (dropped from 1.8 V to ~1.0 V without K3Fe(CN)6 and from 1.8 V to ~1.3 V with K3Fe(CN)6). The specific capacitance values were determined at a current density of 1.25 A/g and were 285 F/g with K3Fe(CN)6 and 186 F/g without K3Fe(CN)6. The coulombic efficiencies were low in these cases, which could be explained by a minor water hydrolysis and the diffusion of [Fe(CN)6]3−/[Fe (CN)6]4− ions. At higher potential than 1.6 V, the edge of water electrolysis started (Figure 8a), and the minor gas generation resulted in the energy loss. Using K3Fe(CN)6, the reaction products ([Fe(CN)6]3− and [Fe(CN)6]4−, mutually) diffused into the whole electrolyte solution in the cell, which decreased the availability of these ions for the discharge process [28].
The specific energy density (SE) and the specific power density (SP) of the WO3/SnO2 symmetric supercapacitor system with/without redox activity were important parameters for determining the feasibility of the device application. SE and SP were calculated using Equations (12) and (13) [53,54]:
S E ( Whkg 1 ) = I M ( 3600 ) 0 t d V d t
S p ( W   Kg 1 ) = E D   X   3600 t d
where M is the total mass of both the positive and negative electrodes, I is the current, and V is the potential as a function of discharge time (td). The Ragone diagram for the energy density vs. power density of the WO3/SnO2 nanocomposite electrodes with/without additive is shown in (Figure 8h). The WO3/SnO2 symmetric supercapacitor system with the K3Fe(CN)6 electrolyte exhibited the highest SE of 64 Whkg−1 at a SP of 542 Wkg−1, whereas without K3Fe(CN)6, the SE was 35 Whkg−1 for a SP of 468 Wkg−1 (in both cases, the current density was 1.25 A/g). Compared with previous reports, the symmetric device with the WO3/SnO2 nanocomposite electrodes exhibited excellent electrochemical performance in the Na2SO4/K3Fe(CN)6 electrolyte (Table 3). In the symmetric system, the active materials simultaneously worked as positive (anode) and negative (cathode) electrodes, and the GCD curves confirmed the reversible balanced redox reactions occurring at the anode and cathode [61,62]. Although the WO3/SnO2 ratio and the concentration of K3Fe(CN)6 should be optimized in future studies, the approach in this study would be useful for designing metal oxide nanocomposites for supercapacitors.

3. Discussion

In this study, microspheres of WO3, nanorods of SnO2, and WO3/SnO2 nanocomposite thin films were synthesized via chemical bath deposition, electrodeposition, and calcination. The hexagonal phase of WO3 and the tetragonal phase of SnO2 were observed in both the single-component and nanocomposite materials. The average WO3 crystal size in the nanocomposite was smaller than that in single-component WO3. An interaction between WO3 and the SnO2 in the nanocomposite was also suggested by the shifts in binding energies measured by XPS. SEM observations indicated the mesoporous surface structures of these nanomaterials, which are preferable for supercapacitor applications.
The CV measurements revealed that the specific capacitance of the WO3/SnO2 nanocomposite electrode reached 530 Fg−1 without K3Fe(CN)6 and 640 Fg−1 with K3Fe(CN)6 at a low scan rate of 5 mV/s. The improved capacitance in the presence of K3Fe(CN)6 was explained by the redox reactions of [Fe(CN)6]3−/[Fe(CN)6]4− ions, in addition to the redox reactions of the WO3/SnO2 nanocomposite. Moreover, the WO3/SnO2 nanocomposite and K3Fe(CN)6 could mutually facilitate their redox reactions. The improved specific capacitance was also observed for single-component electrodes (both of WO3 and SnO2). Therefore, this approach can be used for other electrodes using these metal oxides, although the reaction efficiency could be differed by the electrode species.
Using the symmetric configurations of WO3/SnO2 nanocomposite electrodes, the addition of K3Fe(CN)6 to the 1 M Na2SO4 electrolyte generated a high energy density and a power density of 64 Whkg−1 at 542 Wkg−1, respectively. The capacitive retention efficiencies for the WO3/SnO2 and WO3 electrodes with the K3Fe(CN)6 electrolyte was 94.7% and 90.4%, respectively, as indicated by GCD curves after 2000 cycles at 2.5 Ag−1. The higher cycling stability and the excellent coulombic efficiency (99.9%) of the WO3/SnO2 nanocomposite electrode relative to those of the pure WO3 electrode were due to the high catalytic activity of the WO3/SnO2 nanocomposite for the redox reactions of [Fe(CN)6]3−/[Fe(CN)6]4− ions. As a result, the binder-free WO3/SnO2 nanocomposite with a redox-active electrolyte constituted a promising system for pseudocapacitive energy storage devices. Although the optimal configuration of the WO3/SnO2 nanocomposite and the electrolyte system should be studied further in the future, this approach can also be used with other metal oxide nanocomposites in reactive electrolytes.

4. Materials and Methods

Sodium tungstate dihydrate (Na2WO4·2H2O) was purchased from Alfa Aesar India, and tin (II) chloride hydrate (SnCl2·2H2O) was purchased from Alfa Aesar United States. Hydrochloric acid (HCl, ~35% v/v) and ammonium sulfate ((NH4)2SO4) were purchased from Duksan. Carbon cloth (WOS1009) was obtained from CeTech Co., Ltd., and sodium sulfate (Na2SO4, anhydrous, ≥99%) was from Honeywell/Fluka. Potassium ferricyanide (K3Fe(CN)6, 99%) was purchased from Acros Organics. All of these chemicals were used without further purification. Aqueous solutions were freshly prepared with ultrapure water (resistivity of 18.2 MΩcm; Yamato, Japan) throughout the experiments. Characterization was performed over the range 2θ = 10° to 60° at room temperature with an X-ray diffractometer (XRD, Bruker, D2 phaser Karlsruhe, German) using a copper radiation source (λ = 0.154 nm), operating at 40 kV and 30 mA, with a step size of 0.05 and step time of 5 sec.; X-ray photoelectron spectroscopy (XPS, VG scientific ESCALAB 250, Birmingham, UK); and field emission scanning electron microscopy (FE-SEM, at 15 kV acceleration voltage) equipped with an energy-dispersive X-ray spectroscopy (EDS) analyzer (JSM-6500F, JEOL, Tokyo, Japan).

4.1. Chemical Bath Deposition of WO3 Nanosphere Thin Films

WO3 nanospheres were deposited on a carbon cloth substrate by the CBD method. First, a carbon cloth substrate (1 × 1 cm2) was treated with 1 M HNO3, acetone, ethanol, and water sequentially with ultrasonication for 20 min and then dried overnight in a vacuum. In a typical preparation of WO3 nanospheres, Na2WO4·H2O (0.82 g) was dissolved in 50 mL of water. After stirring for 20 min, 1 M HCl was added dropwise to produce a pH of 1.2 ± 1.0. After stirring for 30 min, (NH4)2SO4 (1.32 g) was added to the above solution. After stirring for 10 min, 40 mL of the above mixture was transferred to a 100 mL beaker, and the carbon cloth substrate was immersed in it. Then, the beaker was heated on a hotplate at 80 °C for 4 h. The carbon cloth was removed from the bath, rinsed several times with water, and dried overnight at room temperature. Finally, the film was calcined at 400 °C for 2 h. The mass of WO3 loaded onto the carbon cloth substrate was measured by subtracting the weight of the original carbon cloth substrate from the weight of the product and determined to be 1.2 mg.

4.2. Electrochemical Deposition of SnO2 Nanorods

SnO2 nanorods were synthesized through electrochemical deposition. First, a carbon cloth substrate with an area of 1 × 1 cm2 was treated with 1 M HNO3, acetone, ethanol, and water through ultrasonication and sequentially dried overnight at ambient temperature. Next, an aqueous solution of SnCl2 (0.112 g) was prepared in 50 mL of water, and 2.6 mL of HCl was added. After stirring for 30 min, 25 mL of the solution was transferred to a 100 mL beaker, and a three-electrode system was set up: carbon cloth was used as the working electrode (WE), Ag/AgCl in 3 M Cl was used as the reference electrode (RE), and a platinum wire was used as the counter electrode (CE). Then, in a solution of SnCl2 with HCl, potentiostatic electrodeposition was performed at an applied potential of −0.9 V for 10 min using a galvanostat/potentiostat (Hokuto-Denko, Model HA-151B) at ambient temperature. After deposition, the WE was rinsed with water and dried overnight at ambient temperature. Finally, the deposited material on the carbon cloth was calcined at 400 °C for 2 h. The deposited mass was measured by subtracting the weight of the original carbon cloth substrate from the weight of the product and determined to be 0.4 mg.

4.3. Synthesis of a WO3/SnO2 Nanocomposite Electrode

WO3/SnO2 nanocomposite thin films were synthesized through CBD and the electrochemical deposition method. First, the carbon cloth substrate of area (1 × 1 cm2) was sequentially treated by ultrasonication in 1 M HNO3, acetone, ethanol, and ultrapure water and then dried overnight at ambient temperature. WO3 microspheres were deposited on the carbon cloth by CBD, and then the carbon cloth/WO3 was immersed in the SnCl2/HCl solution (25 mL, 0.112 g, and 0.65 M) in a 100 mL beaker [63]. Then, the three-electrode system was set up: the WE were carbon cloth supported by WO3 nanospheres, the RE was Ag/AgCl in 3 M Cl, and the CE was a platinum wire. Then, potentiostatic electrodeposition was performed at an applied potential of −0.9 V for 10 min using a galvanostat/potentiostat (Hokuto-Denko, Model HA-151B) at ambient temperature. After deposition, the WE were rinsed with water and then dried overnight at ambient temperature. Finally, the deposited material on the carbon cloth was calcined at 400 °C for 2 h on a hot plate. The deposited mass was measured by subtracting the weight of the original carbon cloth substrate from the weight of the product and was determined to be 1.6 mg.

4.4. Electrochemical Analyses

All electrochemical analyses were performed using a standard three-electrode system in an electrochemical cell connected to an electrochemical workstation (ZAHNER mass system, model xpot-26366, Germany). An Ag/AgCl electrode in 3 M KCl was used as the RE; a Pt wire as CE; and the WO3, SnO2, and WO3/SnO2 nanocomposites formed on carbon cloth were used as WEs. An aqueous solution of 1 M Na2SO4 with/without 0.02 M K3Fe (CN)6 was used as the electrolyte for all electrochemical measurements. The electrochemical performance of WO3 and SnO2, as well as the WO3/SnO2 nanocomposite electrodes, was studied using cyclic voltammetry (CV) over the range −0.2 V to 0.7 V vs. Ag/AgCl at different scan rates and galvanostatic charge-discharge (GCD) at different current densities. Electrochemical impendence spectroscopy (EIS) was performed with an amplitude of 10 mV over the frequency range 100 kHz to 100 mHz in an open circuit and described with a Nyquist plot.

5. Conclusions

In this work, we developed a preparation method of WO3/SnO2 nanocomposite on carbon cloth for supercapacitor electrodes. WO3 nanospheres were first deposited on a carbon cloth substrate by a chemical bath deposition method, and then a SnO2 layer was formed by an electrochemical deposition and calcination. The WO3/SnO2 nanocomposite on carbon cloth exhibited a higher specific capacitance (530 F/g at 5 mV/s scan rate) than that of single-component electrodes (240 F/g for WO3 and 140 F/g for SnO2) in an aqueous electrolyte of 1M Na2SO4. The addition of 0.02 M K3Fe(CN)6 redox-active electrolyte into the 1M Na2SO4 electrolyte further improved the charge storage performance of electrodes (440 F/g for WO3, 310 F/g for SnO2, and 640 F/g for WO3/SnO2 nanocomposite). These results show the synergistic effects of WO3/SnO2 nanocomposites and the advantages of using redox-active electrolytes for supercapacitor systems. Using a symmetric configuration of WO3/SnO2 nanocomposite electrodes, the electrolyte solution of 0.02 MK3Fe(CN)6 and 1 M Na2SO4 electrolyte demonstrated a high energy density and a power density of 64 Whkg−1 at 542 Wkg−1, respectively. The capacitive retention efficiency for the WO3/SnO2 and WO3 electrodes with the K3Fe(CN)6 electrolyte was 94.7% after 2000 cycles at 2.5 Ag−1. Thus, the WO3/SnO2 nanocomposite electrode grown on carbon cloth is a promising candidates as a binder-free electrode for the high performance of a supercapacitor device, and the supercapacitor system using redox-active K3Fe(CN)6 electrolyte was proposed. These strategies developed in this study can also be applied to other metal oxide nanocomposites to improve the performance of supercapacitors.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms24076045/s1.

Author Contributions

Conceptualization, T.D.M. and M.U.; methodology, M.U.; software, M.U.; formal analysis, T.D.M.; investigation, T.D.M.; resources, M.U.; data curation, T.D.M.; writing—original draft preparation, T.D.M.; writing—review and editing, M.U.; visualization, T.D.M.; supervision, M.U.; project administration, M.U.; funding acquisition, M.U. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Science and Technology of the Republic of China, [MOST 109-2221-E-011-062-].

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

This investigation was partly supported by the Graduate Institute of Applied Science and Technology, National Taiwan University of Science and Technology.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shinde, P.A.; Lokhande, A.C.; Patil, A.M.; Lokhande, C.D. Facile synthesis of self-assembled WO3 nanorods for high-performance electrochemical capacitor. J. Alloys Compd. 2019, 770, 1130–1137. [Google Scholar] [CrossRef]
  2. Barik, R.; Yadav, A.K.; Jha, S.N.; Bhattacharyya, D.; Ingole, P.P. Two-dimensional tungsten oxide/selenium nanocomposite fabricated for flexible supercapacitors with higher operational voltage and their charge storage mechanism. ACS Appl. Mater. Interfaces 2021, 13, 8102–8119. [Google Scholar] [CrossRef] [PubMed]
  3. Debelo, T.T.; Ujihara, M. Effect of simultaneous electrochemical deposition of manganese hydroxide and polypyrrole on structure and capacitive behavior. J. Electroanal. Chem. 2020, 859, 113825. [Google Scholar] [CrossRef]
  4. Mevada, C.; Mukhopadhyay, M. High mass loading tin oxide-ruthenium oxide-based nanocomposite electrode for supercapacitor application. J. Energy Storage 2020, 31, 101587. [Google Scholar] [CrossRef]
  5. Kariper, İ.; Tezel, F.M.; Gökkuş, Ö. Synthesize of WO3 thin film supercapacitor and its characterization. Phys. Lett. A 2021, 388, 127059. [Google Scholar] [CrossRef]
  6. Abebe, E.M.; Ujihara, M. Influence of Temperature on ZnO/Co3O4 Nanocomposites for High Energy Storage Supercapacitors. ACS Omega 2021, 6, 23750. [Google Scholar] [CrossRef]
  7. Abebe, E.M.; Ujihara, M. Simultaneous Electrodeposition of Ternary Metal Oxide Nanocomposites for High-Efficiency Supercapacitor Applications. ACS Omega 2022, 7, 17161–17174. [Google Scholar] [CrossRef] [PubMed]
  8. Kang, Y.; Li, Z.; Xu, K.; He, X.; Wei, S.; Cao, Y. Hollow SnO2 nanospheres with single-shelled structure and the application for supercapacitors. J. Alloys Compd. 2019, 779, 728–734. [Google Scholar] [CrossRef]
  9. Debelo, T.T.; Ujihara, M. Electrodeposition of binder-free polypyrrole on a three-dimensional flower-like nanosheet of manganese oxides containing pyrrole derivative for supercapacitor electrode. J. Mater. Chem. A 2022, 278, 125690. [Google Scholar] [CrossRef]
  10. Poudel, M.B.; Lohani, P.C.; Kim, A.A. Synthesis of silver nanoparticles decorated tungsten oxide nanorods as high-performance supercapacitor electrode. Chem. Phys. Lett. 2022, 804, 139884. [Google Scholar] [CrossRef]
  11. Nithya, V.; Arul, N.S. Review on α-Fe2O3 based negative electrode for high performance supercapacitors. J. Power Sources 2016, 327, 297–318. [Google Scholar] [CrossRef]
  12. Shinde, P.A.; Lokhande, A.C.; Chodankar, N.R.; Patil, A.M.; Kim, J.H.; Lokhande, C.D. Temperature dependent surface morphological modifications of hexagonal WO3 thin films for high performance supercapacitor application. Electrochim. Acta 2017, 224, 397–404. [Google Scholar] [CrossRef]
  13. Ji, S.-H.; Chodankar, N.R.; Kim, D.-H. Aqueous asymmetric supercapacitor based on RuO2-WO3 electrodes. Electrochim. Acta 2019, 325, 134879. [Google Scholar] [CrossRef]
  14. Chu, J.; Lu, D.; Wang, X.; Wang, X.; Xiong, S. WO3 nanoflower coated with graphene nanosheet: Synergetic energy storage composite electrode for supercapacitor application. J. Alloys Compd. 2017, 702, 568–572. [Google Scholar] [CrossRef]
  15. Shinde, P.A.; Lokhande, V.C.; Patil, A.M.; Ji, T.; Lokhande, C.D. Single-step hydrothermal synthesis of WO3-MnO2 composite as an active material for all-solid-state flexible asymmetric supercapacitor. Int. J. Hydrogen Energy 2018, 43, 2869–2880. [Google Scholar] [CrossRef]
  16. Kumar, R.D.; Andou, Y.; Karuppuchamy, S. Synthesis and characterization of nanostructured Ni-WO3 and NiWO4 for supercapacitor applications. J. Alloys Compd. 2016, 654, 349–356. [Google Scholar] [CrossRef]
  17. Kumar, R.D.; Karuppuchamy, S. Microwave mediated synthesis of nanostructured Co-WO3 and CoWO4 for supercapacitor applications. J. Alloys Compd. 2016, 674, 384–391. [Google Scholar] [CrossRef]
  18. Hai, Z.; Akbari, M.K.; Xue, C.; Xu, H.; Solano, E.; Detavernier, C.; Hu, J.; Zhuiykov, S. Atomically-thin WO3/TiO2 heterojunction for supercapacitor electrodes developed by atomic layer deposition. Compos. Commun. 2017, 5, 31–35. [Google Scholar] [CrossRef]
  19. Periasamy, P.; Krishnakumar, T.; Sandhiya, M.; Sathish, M.; Chavali, M.; Siril, P.F.; Devarajan, V. Preparation and comparison of hybridized WO3–V2O5 nanocomposites electrochemical supercapacitor performance in KOH and H2SO4 electrolyte. Mater. Lett. 2019, 236, 702–705. [Google Scholar] [CrossRef]
  20. Xu, Y.; Wang, L.; Zhou, Y.; Guo, J.; Zhang, S.; Lu, Y. Synthesis of heterostructure SnO2/graphitic carbon nitride composite for high-performance electrochemical supercapacitor. J. Electroanal. Chem. 2019, 852, 113507. [Google Scholar] [CrossRef]
  21. Shinde, P.A.; Lokhande, V.C.; Ji, T.; Lokhande, C.D. Facile synthesis of hierarchical mesoporous weirds-like morphological MnO2 thin films on carbon cloth for high performance supercapacitor application. J. Colloid Interface Sci. 2017, 498, 202–209. [Google Scholar] [CrossRef] [PubMed]
  22. Shinde, P.A.; Lokhande, V.C.; Patil, A.M.; Ji, T.; Lokhande, C.D. Design and synthesis of hierarchical mesoporous WO3-MnO2 composite nanostructures on carbon cloth for high-performance supercapacitors. J. Electrochem. Sci. Technol. 2017, 3, 35–48. [Google Scholar] [CrossRef]
  23. Veerasubramani, G.K.; Krishnamoorthy, K.; Kim, S.J. Improved electrochemical performances of binder-free CoMoO4 nanoplate arrays@ Ni foam electrode using redox additive electrolyte. J. Power Sources 2016, 306, 378–386. [Google Scholar] [CrossRef]
  24. Ojha, G.P.; Pant, B.; Acharya, J.; Park, M. Prussian red anions immobilized freestanding three-dimensional porous carbonaceous networks: A new avenue to attain capacitor-and faradic-type electrodes in a single frame for 2.0 V hybrid supercapacitors. ACS Sustain. Chem. Eng. 2022, 10, 2994–3006. [Google Scholar] [CrossRef]
  25. Sankar, K.V.; Selvan, R.K. Improved electrochemical performances of reduced graphen oxide based supercapacitor using redox additive electrolyte. Carbon 2015, 90, 260–273. [Google Scholar] [CrossRef]
  26. Huang, X.; Wang, Q.; Chen, X.Y.; Zhang, Z.J. The effects of amine/nitro/hydroxyl groups on the benzene rings of redox additives on the electrochemical performance of carbon-based supercapacitors. Phys. Chem. Chem. Phys. 2016, 18, 10438–10452. [Google Scholar] [CrossRef]
  27. Han, X.; Jiang, T.; Chen, X.; Jiang, D.; Xie, K.; Jiang, Y.; Wang, Y. Electrolyte additive induced fast-charge/slow-discharge process: Potassium ferricyanide and potassium persulfate for CoO-based supercapacitors. J. Colloid Interface Sci. 2020, 576, 505–513. [Google Scholar] [CrossRef]
  28. Lee, J.; Choudhury, S.; Weingarth, D.; Kim, D.; Presser, V. High performance hybrid energy storage with potassium ferricyanide redox electrolyte. ACS Appl. Mater. Interfaces 2016, 8, 23676–23687. [Google Scholar] [CrossRef]
  29. Lokhande, V.; Lokhande, A.; Namkoong, G.; Kim, J.H.; Ji, T. Charge storage in WO3 polymorphs and their application as supercapacitor electrode material. Results Phys. 2019, 12, 2012–2020. [Google Scholar] [CrossRef]
  30. Zheng, F.; Song, S.; Lu, F.; Li, R.; Bu, N.; Liu, J.; Li, Y.; Hu, P.; Zhen, Q. Hydrothermal preparation, growth mechanism and supercapacitive properties of WO 3 nanorod arrays grown directly on a Cu substrate. CrystEngComm 2016, 18, 3891–3904. [Google Scholar] [CrossRef]
  31. Gao, L.; Wang, X.; Xie, Z.; Song, W.; Wang, L.; Wu, X.; Qu, F.; Chen, D.; Shen, G. High-performance energy-storage devices based on WO3 nanowire arrays/carbon cloth integrated electrodes. J. Mater. Chem. A 2013, 1, 7167–7173. [Google Scholar] [CrossRef]
  32. Nayak, A.K.; Ghosh, R.; Santra, S.; Guha, P.K.; Pradhan, D. Hierarchical nanostructured WO 3–SnO 2 for selective sensing of volatile organic compounds. Nanoscale 2015, 7, 12460–12473. [Google Scholar] [CrossRef] [PubMed]
  33. Zhu, Y.; Wang, H.; Liu, J.; Yin, M.; Yu, L.; Zhou, J.; Liu, Y.; Qiao, F. High-performance gas sensors based on the WO3-SnO2 nanosphere composites. J. Alloys Compd. 2019, 782, 789–795. [Google Scholar] [CrossRef]
  34. Darwish, A.M.; Eisa, W.H.; Shabaka, A.A.; Talaat, M.H. Investigation of factors affecting the synthesis of nano-cadmium sulfide by pulsed laser ablation in liquid environment. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2016, 153, 315–320. [Google Scholar] [CrossRef] [PubMed]
  35. Hasanin, M.S.; Mostafa, A.M.; Mwafy, E.A.; Darwesh, O.M. Eco-friendly cellulose nano fibers via first reported Egyptian Humicola fuscoatra Egyptia X4: Isolation and characterization. Environ. Nanotechnol. Monit. Manag. 2018, 10, 409–418. [Google Scholar] [CrossRef]
  36. Geleta, T.A.; Imae, T. Influence of additives on zinc oxide-based dye sensitized solar cells. Bull. Chem. Soc. Jpn. 2020, 93, 611. [Google Scholar] [CrossRef]
  37. Shao, H.; Huang, M.; Fu, H.; Wang, S.; Wang, L.; Lu, J.; Wang, Y.; Yu, K. Hollow WO3/SnO2 hetero-nanofibers: Controlled synthesis and high efficiency of acetone vapor detection. Front. Chem. 2019, 7, 785. [Google Scholar] [CrossRef] [PubMed]
  38. Wu, C.-M.; Motora, K.G.; Chen, G.-Y.; Kuo, D.-H.; Gultom, N.S. Highly efficient MoS2@CsxWO3 nanocomposite hydrogen gas sensors. Front. Mater. 2022, 9, 18. [Google Scholar] [CrossRef]
  39. Claros, M.; Setka, M.; Jimenez, Y.P.; Vallejos, S. AACVD synthesis and characterization of iron and copper oxides modified ZnO structured films. J. Nanomater. 2020, 10, 471. [Google Scholar] [CrossRef] [Green Version]
  40. Motsoeneng, R.G.; Kortidis, I.; Ray, S.S.; Motaung, D.E. Designing SnO2 nanostructure-based sensors with tailored selectivity toward propanol and ethanol vapors. J. Alloys Compd. 2019, 4, 13696–13709. [Google Scholar] [CrossRef] [Green Version]
  41. Ciftyürek, E.; Šmíd, B.; Li, Z.; Matolín, V.; Schierbaum, K. Spectroscopic understanding of SnO2 and WO3 metal oxide surfaces with advanced synchrotron based; XPS-UPS and near ambient pressure (NAP) XPS surface sensitive techniques for gas sensor applications under operational conditions. J. Sens. 2019, 19, 4737. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Farhadian, M.; Sangpour, P.; Hosseinzadeh, G. Morphology dependent photocatalytic activity of WO3 nanostructures. J. Energy Chem. 2015, 24, 171–177. [Google Scholar] [CrossRef]
  43. Han, C.; Zhang, B.; Zhao, K.; Meng, J.; He, Q.; He, P.; Yang, W.; Li, Q.; Mai, L. Oxalate-assisted formation of uniform carbon-confined SnO2 nanotubes with enhanced lithium storage. Chem. Commun. 2017, 53, 9542–9545. [Google Scholar] [CrossRef] [PubMed]
  44. Liu, M.; Xu, P.; Wang, G.; Yan, J. SnO2 Nanorod arrays grown on carbon cloth as a flexible binder-free electrode for high-performance lithium batteries. J. Electron. Mater. 2019, 48, 8206–8211. [Google Scholar] [CrossRef]
  45. Zhao, Y.-F.; Sun, Y.-P.; Yin, X.; Yin, G.-C.; Wang, X.-M.; Jia, F.-C.; Liu, B. Effect of surfactants on the microstructures of hierarchical SnO2 blooming nanoflowers and their gas-sensing properties. Nanoscale Res. Lett. 2018, 13, 54. [Google Scholar] [CrossRef] [PubMed]
  46. Gupta, S.P.; Nishad, H.H.; Chakane, S.D.; Gosavi, S.W.; Late, D.J.; Walke, P.S. Phase transformation in tungsten oxide nanoplates as a function of post-annealing temperature and its electrochemical influence on energy storage. Nanoscale Adv. 2020, 2, 4689–4701. [Google Scholar] [CrossRef]
  47. Huang, H.; Tan, O.; Lee, Y.; Tse, M.; Guo, J.; White, T. In situ growth of SnO2 nanorods by plasma treatment of SnO2 thin films. J. Nanotechnol. 2006, 17, 3668. [Google Scholar] [CrossRef]
  48. Li, K.-D.; Chen, P.-W.; Chang, K.-S.; Hsu, S.-C.; Jan, D.-J. Indium-Zinc-Tin-Oxide Film Prepared by Reactive Magnetron Sputtering for Electrochromic Applications. Materials 2018, 11, 2221. [Google Scholar] [CrossRef] [Green Version]
  49. Yang, F.; Jia, J.; Mi, R.; Liu, X.; Fu, Z.; Wang, C.; Liu, X.; Tang, Y. Fabrication of WO3· 2H2O/BC hybrids by the radiation method for enhanced performance supercapacitors. Front. Chem. 2018, 6, 290. [Google Scholar] [CrossRef] [Green Version]
  50. Wang, S.; Xu, H.; Hao, T.; Wang, P.; Zhang, X.; Zhang, H.; Xue, J.; Zhao, J.; Li, Y. In situ XRD and operando spectra-electrochemical investigation of tetragonal WO3-x nanowire networks for electrochromic supercapacitors. NPG Asia Mater. 2021, 13, 51. [Google Scholar] [CrossRef]
  51. Ashraf, M.; Shah, S.S.; Khan, I.; Aziz, M.A.; Ullah, N.; Khan, M.; Adil, S.F.; Liaqat, Z.; Usman, M.; Tremel, W. A High-Performance Asymmetric Supercapacitor Based on Tungsten Oxide Nanoplates and Highly Reduced Graphene Oxide Electrodes. Chem. A Eur. J. 2021, 27, 6973–6984. [Google Scholar] [CrossRef] [PubMed]
  52. Angelin, M.D.; Rajkumar, S.; Merlin, J.P.; Xavier, A.R.; Franklin, M.; Ravichandran, A. Electrochemical investigation of Zr-doped ZnO nanostructured electrode material for high-performance supercapacitor. Ionics 2020, 26, 5757–5772. [Google Scholar] [CrossRef]
  53. Mathis, T.S.; Kurra, N.; Wang, X.; Pinto, D.; Simon, P.; Gogotsi, Y. Energy storage data reporting in perspective—Guidelines for interpreting the performance of electrochemical energy storage systems. Adv. Energy Mater. 2019, 9, 1902007. [Google Scholar] [CrossRef]
  54. Pham, D.T.; Baboo, J.P.; Song, J.; Kim, S.; Jo, J.; Mathew, V.; Alfaruqi, M.H.; Sambandam, B.; Kim, J. Facile synthesis of pyrite (FeS2/C) nanoparticles as an electrode material for non-aqueous hybrid electrochemical capacitors. Nanoscale 2018, 10, 5938–5949. [Google Scholar] [CrossRef] [PubMed]
  55. Nagaraju, P.; Alsalme, A.; Alkathiri, A.M.; Jayavel, R. Rapid synthesis of WO3/graphene nanocomposite via in-situ microwave method with improved electrochemical properties. J. Phys. Chem. Solids 2018, 120, 250–260. [Google Scholar] [CrossRef]
  56. Huang, Y.; Li, M.; Chen, S.; Sun, P.; Lv, X.; Li, B.; Fang, L.; Sun, X. Constructing aqueous Zn//Ni hybrid battery with NiSe nanorod array on nickel foam and redox electrolytes for high-performance electrochemical energy storage. Appl. Surf. Sci. 2021, 562, 150222. [Google Scholar] [CrossRef]
  57. Chowdhury, A.; Biswas, S.; Singh, T.; Chandra, A. Redox mediator induced electrochemical reactions at the electrode-electrolyte interface: Making sodium-ion supercapacitors a competitive technology. Electrochem. Sci. Adv. 2022, 2, e2100030. [Google Scholar] [CrossRef]
  58. Kim, H.-S.; Cook, J.B.; Lin, H.; Ko, J.S.; Tolbert, S.H.; Ozolins, V.; Dunn, B. Oxygen vacancies enhance pseudocapacitive charge storage properties of MoO3−x. Nat. Mater. 2017, 16, 454–460. [Google Scholar] [CrossRef]
  59. Dong, X.; Chen, L.; Liu, J.; Haller, S.; Wang, Y.; Xia, Y. Environmentally-friendly aqueous Li (or Na)-ion battery with fast electrode kinetics and super-long life. Sci. Adv. 2016, 2, e1501038. [Google Scholar] [CrossRef] [Green Version]
  60. Teli, A.M.; Beknalkar, S.A.; Pawar, S.A.; Dubal, D.P.; Dongale, T.D.; Patil, D.S.; Patil, P.S.; Shin, J.C. Effect of concentration on the charge storage kinetics of nanostructured MnO2 thin-film supercapacitors synthesized by the hydrothermal method. Energies 2020, 13, 6124. [Google Scholar] [CrossRef]
  61. Sahoo, M.K.; Rao, G.R. A high energy flexible symmetric supercapacitor fabricated using N-doped activated carbon derived from palm flowers. Nanoscale Adv. 2021, 3, 5417–5429. [Google Scholar] [CrossRef] [PubMed]
  62. Tian, J.; Lin, B.; Sun, Y.; Zhang, X.; Yang, H. Porous WO3@CuO composites derived from polyoxometalates@ metal organic frameworks for supercapacitor. Mater. Lett. 2017, 206, 91–94. [Google Scholar] [CrossRef]
  63. Gao, L.; Qu, F.; Wu, X. Hierarchical WO3@SnO2 core–shell nanowire arrays on carbon cloth: A new class of anode for high-performance lithium-ion batteries. J. Mater. Chem. A 2014, 2, 7367–7372. [Google Scholar] [CrossRef]
Figure 1. XRD patterns for (a) WO3, (b) SnO2, and (c) WO3/SnO2 nanocomposites. The clubs (♣) and diamonds (♦) represent peaks of CC and SnO2, respectively.
Figure 1. XRD patterns for (a) WO3, (b) SnO2, and (c) WO3/SnO2 nanocomposites. The clubs (♣) and diamonds (♦) represent peaks of CC and SnO2, respectively.
Ijms 24 06045 g001
Figure 2. XPS data for WO3, SnO2, and the WO3/SnO2 nanocomposite: (a) survey spectrum of the WO3/SnO2 nanocomposite, high-resolution (b) W 4f spectrum of WO3, (c) O 1s spectrum of WO3, (d) Sn 3d spectrum of SnO2, (e) O 1s spectrum of SnO2, (f) W 4f spectrum of WO3/SnO2, (g) Sn 3d spectrum of WO3/SnO2, and (h) O 1s spectrum of the WO3/SnO2 nanocomposite.
Figure 2. XPS data for WO3, SnO2, and the WO3/SnO2 nanocomposite: (a) survey spectrum of the WO3/SnO2 nanocomposite, high-resolution (b) W 4f spectrum of WO3, (c) O 1s spectrum of WO3, (d) Sn 3d spectrum of SnO2, (e) O 1s spectrum of SnO2, (f) W 4f spectrum of WO3/SnO2, (g) Sn 3d spectrum of WO3/SnO2, and (h) O 1s spectrum of the WO3/SnO2 nanocomposite.
Ijms 24 06045 g002
Figure 3. FE-SEM images of (a,b) WO3 grown on a carbon cloth, (c,d) SnO2 grown on a carbon cloth, (e,f) WO3/SnO2 nanocomposite, and (g,h) cross-sectional image of WO3/SnO2 nanocomposite at low and high magnification, respectively.
Figure 3. FE-SEM images of (a,b) WO3 grown on a carbon cloth, (c,d) SnO2 grown on a carbon cloth, (e,f) WO3/SnO2 nanocomposite, and (g,h) cross-sectional image of WO3/SnO2 nanocomposite at low and high magnification, respectively.
Ijms 24 06045 g003
Figure 4. EDS spectra of synthesized (a) WO3 microspheres, (b) SnO2 nanorods, and (c) WO3/SnO2 nano-composites; (d) SEM image and elemental maps of WO3/SnO2 nanocomposites for (e) O, (f) W, and (g) Sn.
Figure 4. EDS spectra of synthesized (a) WO3 microspheres, (b) SnO2 nanorods, and (c) WO3/SnO2 nano-composites; (d) SEM image and elemental maps of WO3/SnO2 nanocomposites for (e) O, (f) W, and (g) Sn.
Ijms 24 06045 g004
Figure 5. CV curves for (a) WO3, (b) SnO2, and (c) WO3/SnO2 nanocomposite electrodes at various scan rates from 5 mV/s to 50 mV/s with 1 M Na2SO4 electrolyte; (d) comparison of CV curves for WO3, SnO2, and WO3/SnO2 at 50 mV/s; (e) specific capacitance vs. scan rate for WO3, SnO2, and WO3/SnO2 from CV curves.
Figure 5. CV curves for (a) WO3, (b) SnO2, and (c) WO3/SnO2 nanocomposite electrodes at various scan rates from 5 mV/s to 50 mV/s with 1 M Na2SO4 electrolyte; (d) comparison of CV curves for WO3, SnO2, and WO3/SnO2 at 50 mV/s; (e) specific capacitance vs. scan rate for WO3, SnO2, and WO3/SnO2 from CV curves.
Ijms 24 06045 g005
Figure 6. GCD curves determined for electrodes with 1 M Na2SO4 electrolyte and at different current densities: (a) WO3, (b) SnO2, and (c) WO3/SnO2; (d) comparison of GCD curves for WO3, SnO2, and WO3/SnO2 with the same current density of 1.25 A/g; (e) EIS for WO3, SnO2, and WO3/SnO2 nanocomposite electrodes; and (f) magnified EIS for the inset in (e), and the inset in (e) is the model circuit used for the analysis.
Figure 6. GCD curves determined for electrodes with 1 M Na2SO4 electrolyte and at different current densities: (a) WO3, (b) SnO2, and (c) WO3/SnO2; (d) comparison of GCD curves for WO3, SnO2, and WO3/SnO2 with the same current density of 1.25 A/g; (e) EIS for WO3, SnO2, and WO3/SnO2 nanocomposite electrodes; and (f) magnified EIS for the inset in (e), and the inset in (e) is the model circuit used for the analysis.
Ijms 24 06045 g006
Figure 7. CV curves for (a) various electrodes at scan rate of 50 mV/s with K3Fe(CN)6 and (b) WO3 and the WO3/SnO2 nanocomposite electrodes with/without K3Fe(CN)6 at scan rate of 30 mV/s; CV curves of (c) WO3 electrode, (d) SnO2 electrode, and (e) WO3/SnO2 nanocomposite electrode with K3Fe(CN)6 at various scan rates; (f) specific capacitance of various electrodes with/without K3Fe(CN)6 at various scan rates; (g) double logarithmic plots of scan rate vs. peak current for WO3, SnO2, and the WO3/SnO2 nanocomposite; (h) cycling stability and coulombic efficiency versus cycle number for the WO3/SnO2 nanocomposite electrode with K3Fe(CN)6; (i) EIS analyses of WO3, SnO2, and WO3/SnO2 nanocomposite electrodes with K3Fe(CN)6 over the frequency range 100 kHz to 100 mHz.
Figure 7. CV curves for (a) various electrodes at scan rate of 50 mV/s with K3Fe(CN)6 and (b) WO3 and the WO3/SnO2 nanocomposite electrodes with/without K3Fe(CN)6 at scan rate of 30 mV/s; CV curves of (c) WO3 electrode, (d) SnO2 electrode, and (e) WO3/SnO2 nanocomposite electrode with K3Fe(CN)6 at various scan rates; (f) specific capacitance of various electrodes with/without K3Fe(CN)6 at various scan rates; (g) double logarithmic plots of scan rate vs. peak current for WO3, SnO2, and the WO3/SnO2 nanocomposite; (h) cycling stability and coulombic efficiency versus cycle number for the WO3/SnO2 nanocomposite electrode with K3Fe(CN)6; (i) EIS analyses of WO3, SnO2, and WO3/SnO2 nanocomposite electrodes with K3Fe(CN)6 over the frequency range 100 kHz to 100 mHz.
Ijms 24 06045 g007
Figure 8. (a) CV curves of the WO3/SnO2 symmetric electrode with different applied potential windows at 50 mV/s, symmetric CV curves of the WO3/SnO2 symmetric electrode at different scan rates (b) without the redox-active electrolyte and (c) with the redox-active electrolyte; (d) comparison of the CV curves for the WO3/SnO2 symmetric device with and without the redox-active electrolyte at 50 mV/s; GCD curve for the WO3/SnO2 symmetric electrode at different current densities (e) without the redox-active electrolyte and (f) with the redox-active electrolyte; (g) comparison of the GCD curve-based symmetric electrode at 1.25 A/g; (h) Ragone plots of the WO3/SnO2 symmetric device.
Figure 8. (a) CV curves of the WO3/SnO2 symmetric electrode with different applied potential windows at 50 mV/s, symmetric CV curves of the WO3/SnO2 symmetric electrode at different scan rates (b) without the redox-active electrolyte and (c) with the redox-active electrolyte; (d) comparison of the CV curves for the WO3/SnO2 symmetric device with and without the redox-active electrolyte at 50 mV/s; GCD curve for the WO3/SnO2 symmetric electrode at different current densities (e) without the redox-active electrolyte and (f) with the redox-active electrolyte; (g) comparison of the GCD curve-based symmetric electrode at 1.25 A/g; (h) Ragone plots of the WO3/SnO2 symmetric device.
Ijms 24 06045 g008
Table 1. Binding energies (eV) for W, Sn, and O in WO3, SnO2, and WO3/SnO2 nanocomposites determined by deconvolution of XPS peaks.
Table 1. Binding energies (eV) for W, Sn, and O in WO3, SnO2, and WO3/SnO2 nanocomposites determined by deconvolution of XPS peaks.
W 4f7/2W 4f5/2Doublet Separation
WO336.0838.212.13
WO3/SnO235.8438.082.24
Sn 3d5/2Sn 3d3/2
SnO2487.60496.058.45
WO3/SnO2487.69496.128.43
O in Lattice Metal OxideO Deficiency (Defects)Chemisorbed OO in Water, Hydroxide (OH/H2O)
WO3531.43532.01532.72533.60
SnO2531.50-532.70534.38
WO3/SnO2531.09531.76532.80534.30
Table 2. Charge Transfer Resistance (Rct) and Electrochemical Resistance (Rs) of Different Electrodes with/without K3Fe(CN)6.
Table 2. Charge Transfer Resistance (Rct) and Electrochemical Resistance (Rs) of Different Electrodes with/without K3Fe(CN)6.
ElectrodeWith K3Fe(CN)6Without K3Fe(CN)6
Rs (Ω)Rct (Ω)Rs (Ω)Rct (Ω)
WO32.781.712.372.12
SnO22.131.692.321.94
WO3/SnO21.651.561.871.72
Table 3. Comparison of the electrochemical performance for WO3-based active electrode materials with the parameters for supercapacitor applications.
Table 3. Comparison of the electrochemical performance for WO3-based active electrode materials with the parameters for supercapacitor applications.
ElectrodeSpecific Capacitance
(Fg−1)
Current Density
(Ag−1)
Energy Density
(Whkg−1)
Power Density
(Wkg−1)
Stability
(Retention, Cycles)
ElectrolyteRef.
MnO2-WO3657@5 mVs−1NA10.865092%, 2000Na2SO4[21]
RuO2-WO3NANA16.92540NANA[12]
WO3-MnO2/WO388.64 mAcm−224.1391595%, 2500CMC-Na2SO4[14]
WO3-RGO4951NANA85%, 10000.5 M H2SO4[13]
WO3@CuO248.21NANA85.2%, 15006 M KOH[60]
WO3/Se(ASC)0.8580.20.0470.34574%, 4000PVA-H2SO4[2]
WO3/SnO2530@5 mVs−1NA35468NA1 M Na2SO4This work
WO3/SnO2640@5 mVs−1NA6454294.7%, 20000.02 M K3Fe(CN)6/1 M Na2SO4This 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

Morka, T.D.; Ujihara, M. Enhanced Performance of WO3/SnO2 Nanocomposite Electrodes with Redox-Active Electrolytes for Supercapacitors. Int. J. Mol. Sci. 2023, 24, 6045. https://doi.org/10.3390/ijms24076045

AMA Style

Morka TD, Ujihara M. Enhanced Performance of WO3/SnO2 Nanocomposite Electrodes with Redox-Active Electrolytes for Supercapacitors. International Journal of Molecular Sciences. 2023; 24(7):6045. https://doi.org/10.3390/ijms24076045

Chicago/Turabian Style

Morka, Tamiru Deressa, and Masaki Ujihara. 2023. "Enhanced Performance of WO3/SnO2 Nanocomposite Electrodes with Redox-Active Electrolytes for Supercapacitors" International Journal of Molecular Sciences 24, no. 7: 6045. https://doi.org/10.3390/ijms24076045

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