Next Article in Journal
Influence of Pd-Layer Thickness on Bonding Reliability of Pd-Coated Cu Wire
Previous Article in Journal
Study on Electrical and Mechanical Properties of Double-End Supported Elastic Substrate Prepared by Wet Etching Process
Previous Article in Special Issue
Exploring the Capability of Cu-MoS2 Catalysts for Use in Electrocatalytic Overall Water Splitting
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ni3V2O8 Marigold Structures with rGO Coating for Enhanced Supercapacitor Performance

1
School of Mechanical Engineering, Yeungnam University, Gyeongsan 38541, Republic of Korea
2
School of Chemical Engineering, Yeungnam University, 280 Daehak-ro, Gyeongsan 38541, Republic of Korea
3
Division of Electronics and Electrical Engineering, Dongguk University—Seoul, 30 Pildong-ro, Jung-gu, Seoul 04620, Republic of Korea
4
Department of Electronic Engineering, Institute for Wearable Convergence Electronics, Kyung Hee University, 1732 Deogyeong-daero, Giheung-gu, Yongin 17104, Republic of Korea
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Micromachines 2024, 15(7), 930; https://doi.org/10.3390/mi15070930
Submission received: 10 June 2024 / Revised: 12 July 2024 / Accepted: 15 July 2024 / Published: 20 July 2024
(This article belongs to the Special Issue Electrochemical Supercapacitors for Energy Harvesting and Storage)

Abstract

:
In this work, Ni3V2O8 (NVO) and Ni3V2O8-reduced graphene oxide (NVO-rGO) are synthesized hydrothermally, and their extensive structural, morphological, and electrochemical characterizations follow subsequently. The synthetic materials’ crystalline structure was confirmed by X-ray diffraction (XRD), and its unique marigold-like morphology was observed by field emission scanning electron microscopy (FESEM). The chemical states of the elements were investigated via X-ray photoelectron spectroscopy (XPS). Electrochemical impedance spectroscopy (EIS), Galvanostatic charge–discharge (GCD), and cyclic voltammetry (CV) were used to assess the electrochemical performance. A specific capacitance of 132 F/g, an energy density of 5.04 Wh/kg, and a power density of 187 W/kg were demonstrated by Ni3V2O8-rGO. Key electrochemical characteristics were b = 0.67; a transfer coefficient of 0.52; a standard rate constant of 6.07 × 10−5 cm/S; a diffusion coefficient of 5.27 × 10−8 cm2/S; and a series resistance of 1.65 Ω. By employing Ni3V2O8-rGO and activated carbon, an asymmetric supercapacitor with a specific capacitance of 7.85 F/g, an energy density of 3.52 Wh/kg, and a power density of 225 W/kg was achieved. The series resistance increased from 4.27 Ω to 6.63 Ω during cyclic stability tests, which showed 99% columbic efficiency and 87% energy retention. The potential of Ni3V2O8-rGO as a high-performance electrode material for supercapacitors is highlighted by these findings.

1. Introduction

As the energy crisis continues to worsen, energy storage that makes use of renewable energy sources has emerged as one of the most promising strategies for mitigating the effects of both the energy crisis and environmental degradation. There are a variety of energy storage devices that are currently available, including lithium-ion batteries, sodium-ion batteries, zinc-ion batteries, and others. Supercapacitors have garnered more attention than the other types of capacitors due to the distinctive qualities that they possess [1,2]. Supercapacitors are highly beneficial for prolonging battery life and are used in a wide range of applications, such as heavy machinery, hybrid autos, and the development of small-scale electronic equipment [3,4]. Supercapacitors possess numerous benefits, such as a prolonged lifespan and a high power density, in addition to their uncomplicated charge storage theory and construction. Supercapacitors are additionally cost effective, secure, and require minimal maintenance [5]. Supercapacitors are commonly made using transition metal oxides, carbon, and conducting electrical polymers as active electrode materials [6,7,8]. Electric double-layer capacitors (EDLCs) and pseudocapacitors (PCs) are the two types of supercapacitor electrode materials. The PC mechanism is demonstrated by metal oxides and metal chalcogenides, whereas the EDLC mechanism is represented by carbon-based materials such as rGO and CNT. Each of these materials possesses distinct electrical, chemical, and structural properties that impact the lifespan and ultimate performance of the supercapacitor [9]. Among them vanadates, which exhibit robust pseudocapacitive activity, have been used as electrode materials [10]. Vanadium-based materials exhibit exceptional chemical stability and a broad range of oxidation states, resulting in excellent electrochemical performance in supercapacitors [11]. Vanadium oxide is a highly promising electroactive redox material for use in supercapacitors due to its abundant availability, diverse oxidation states, and cost effectiveness [12]. Nickel oxides possess exceptional attributes such as a high theoretical specific capacity of 1292 C/g, low toxicity, high capacitance, low production cost, an ecologically friendly nature, and a charge storage mechanism known as pseudocapacitance [13,14]. The nickel and vanadium oxides possess unique characteristics. Therefore, nickel vanadium oxides are considered highly promising materials for usage in supercapacitors due to their wide range of oxidation states, cheap production cost, and ease of synthesis [15,16,17,18]. Nickel vanadium oxide nanoparticles are synthesized using a different synthesis methods such as the solvothermal/hydrothermal method, the sonochemical method, co-precipitation, and sol–gel, etc. [19,20]. The nickel vanadium oxide (NVO) electrode’s electrical conductivity and surface area can be greatly increased by incorporating reduced grapheme oxide (rGO). In the present report, we opted for the hydrothermal method for the typical synthesis of NVO and NVO-rGO nanoparticles. Micro-structural and morphological analysis, elemental mapping analysis, and electrochemical performance analysis are studied, and it is concluded that synthesized NVO and NVO-rGO are promising candidates for supercapacitor application.

2. Experimental Section

Ni3V2O8 (NVO) and Ni3V2O8-rGO (NVO-rGO) composite nanoparticles were synthesized via a straightforward hydrothermal process employing ammonium fluoride (AF) and urea (U). To synthesize the Ni3V2O8 nanoparticles, a solution was prepared by dissolving 0.1 M of nickel nitrate and 0.05 M of vanadium chloride in 40 mL of solvent in a 250 mL beaker. The solution was stirred regularly for 30 min to ensure homogeneity. Subsequently, a solution containing 1.2 M urea and 0.6 M ammonium fluoride was added to the previous solution and stirred for an additional 30 min to achieve homogeneity. The entire solution was thereafter transferred into a hydrothermal stainless steel reactor with a capacity of 100 mL. The reactor was carefully sealed and then placed in an oven set at a temperature of 140 °C for a reaction period of 12 h. Following the completion of the reaction, the reactor underwent natural cooling, and the resulting product was filtered using filter paper. The product had many washes with ethanol and water, each performed more than four times. Subsequently, it was dried in an oven at a temperature of 60 °C for the duration of one night. Finally, the product was annealed at a temperature of 400 °C for a period of 4 h. The product that underwent the process of annealing was given the name NVO-AFU. The synthesis of the Ni3V2O8-rGO product followed a similar procedure, except for the inclusion of 40 mg of graphene oxide (GO) at the beginning. The name given to this product is NVO-AFU-rGO. Ultimately, the products were utilized for additional investigations into their structural, surface, elemental, and electrochemical capabilities. These investigations were conducted on both the powdered form of the goods and the electrodes prepared from them, using the same naming convention.
PANalytical XRD with CuKα radiation was used to investigate the crystal structure and phase development of the microparticles. Field-emission scanning electron microscopy (FE-SEM; S-4800 HITACHI, Ltd., Chiyoda, Japan) examined nanoparticle surface morphology and elemental mapping. Surface chemical composition was investigated by XPS (K-alpha, Thermo Scientific, Altrincham, UK). A ZIVE SP5 electrochemical workstation was used to perform electrochemical measurements, including CV, EIS, GCD, LSV, and cyclic stability measurements, using a platinum counter electrode, microparticles as the working electrode, and an Ag/AgCl reference electrode (WonAtech, Seoul, Korea). To investigate the electrochemical performance of NVO-AFU and NVO-AFU-rGO in a three-electrode setup, NVO-AFU and NVO-AFU-rGO powders were used. The Ni foam was ultrasonically cleaned for 30 min with acetone, ethanol, and water. Slurries were prepared utilizing active material (NVO-AFU and NVO-AFU-rGO powders), PVDF, and carbon black in an 80:10:10 ratio. The prepared slurries were drop-coated onto pre-cleaned Ni foam and dried at 60 °C overnight. For the two-electrode configuration, NVO-AFU-rGO electrodes, termed the positive electrodes, were prepared in the same way, while activated carbon electrodes were prepared in the same way but with activated carbon instead of NVO-AFU-rGO powder. Activated carbon served as the negative electrode in the two-electrode setup. To investigate the ASC device’s electrochemical performance, the positive and negative electrodes were wrapped in paraffin paper and separated with filter paper in a 2 M KOH electrolyte.

3. Result and Discussion

3.1. X-ray Diffraction Analysis

X-ray diffraction (XRD) technology was used to assess the structural characteristics of nanomaterials, particularly the phase, purity, and crystallinity. Moreover, it serves as a potent analytical tool with the ability to provide precise information regarding the chemical structure of a given sample as well as the size of its unit cells. Figure 1 displays the X-ray diffraction pattern of the NVO_AFU and NVO_AFU_rGO samples, represented by the purple and blue colors, respectively. The presence of diffraction planes at 35.7°, 43.8°, and 63.4°, corresponding to the crystallographic indices (221), (151), and (135), respectively, indicates the synthesis of the NVO_AFU phase. The presence of an additional diffraction plane (002) at an angle of 26.3° in the NVO_AFU_rGO sample indicates the successful development of the NVO_AFU_rGO composite. The diffraction peak observed at an angle of 26.3° corresponds to the (002) plane, which is the distinctive peak of reduced graphene oxide (rGO), similar to that reported by V. Chellappa et al. [21]. The establishment of the NVO_AFU and NVO_AFU_rGO phases is confirmed by JCPDS card #01-070-1394. The NVO possesses an orthorhombic crystal structure with the space group Cmca (space group number: 64). The NVO crystal has lattice parameters of a = 5.93 Å, b = 11.42 Å, and c = 8.24 Å.

3.2. Field Electron Scanning Electron Microscopy (FESEM) Analysis

The surface morphology of the synthesized sample has been investigated using field electron scanning electron microscopy (SEM), while the elemental composition was analyzed using energy-dispersive X-ray spectroscopy (EDS). Figure 2a–d displays the surface characteristics of NVO_AFU nanoparticles at magnifications of ×20k, 30k, 40k, and 100k. Figure 2a–d clearly illustrate that the NVO_AFU nanoparticles possess a homogeneous marigold flower-like structure with an average diameter of approximately 2 μm. Figure 2f–i display the scanning electron microscope (SEM) images of the NVO_AFU_rGO composite at magnifications of 20k, 30k, 40k, and 100k. Figure 2f–i clearly show the visible presence of rGO nano-sheets on the pellet of NVO_AFU nanoparticles. This verifies the successful synthesis of the NVO_AFU_rGO composite. Some aggregation is observed in the SEM micrographs of both NVO_AFU and NVO_AFU_rGO nanoparticles. Figure 2e,j display the EDS spectra of the NVO_AFU and NVO_AFU_rGO composites, respectively, along with the elemental composition inset. Figure 2 provides information on the materials. This validates the formation of the stoichiometric phase in the NVO_AFU and NVO_AFU_rGO composites. The energy-dispersive X-ray spectroscopy (EDS) results in Figure 2e show that the stoichiometric ratio of Ni:V:O in NVO_AFU is 45.26:25.95:28.80. Figure 2j shows that the Ni:V:C:O ratio for NVO_AFU_rGO is 13.42:10.81:43.24:35.52 stoichiometrically. This elemental mapping result provides additional confirmation of the creation of the NVO_AFU and NVO_AFU_rGO composite phases, without any presence of impurities. These findings provide more evidence to validate the XRD analysis.

3.3. XPS Analysis

The surface chemistry of the as-synthesized NVO_rGO nano-composite was investigated by X-ray photoelectron spectroscopy (XPS). The XPS approach was used in Figure 3 to find out the binding energies of different chemical states of elements found on the surface of the NVO_rGO sample. Figure 3a displays the XPS survey scan of the NVO_rGO sample, which was produced using hydrothermal methods. The scan covers the energy range of 0–1200 eV.
The peaks seen in the survey spectra at 285.90 eV, 529.13 eV, 854.18 eV, and 517.07 eV indicate the presence of carbon, oxygen, nickel, and vanadium in the NVO_rGO composite that was formed. The Gaussian and Lorentzian peak fitting analysis approach was employed to investigate the different oxidation states of nickel, vanadium, oxygen, and carbon. The XPS spectra for the Ni2p orbital are displayed in Figure 3b. The saturation peak is represented by two moderately intense peaks seen at binding energies of 861.65 eV and 880.25 eV. Two prominent peaks are detected at 855.61 eV and 872.34 eV, representing the Ni 2p3/2 and Ni 2p1/2 energy levels, respectively. The distinctive peaks of Ni2+ are assigned to energy levels of 855.59 eV and 873.02 eV for Ni 2p3/2 and Ni 2p1/2, respectively. The distinctive peaks of Ni3+ are ascribed to occur at 857.85 eV and 874.79 eV for the Ni 2p3/2 and Ni 2p1/2 oxidation states, respectively [22,23,24]. Figure 3c displays the spectra pertaining to the C 1s area. The peaks detected at binding energies of 284.04 eV, 285.05 eV, and 286.18 eV in the C 1s region correspond to the carbonate groups C-C, C-O-C, and O-C=O, respectively [25,26]. Figure 3d shows the measured peaks corresponding to the binding energies of 530.09 eV, 530.46 eV, and 531.76 eV for the oxygen atoms OI, OII, and OIII, respectively, in the O 1s area of transition metal (oxy) hydroxides [27,28,29]. Figure 3e displays the V2p XPS spectra of the NVO_rGO nano-composite. The V 2p XPS spectra exhibit two clearly distinguishable peaks at 516.77 eV and 524.37 eV, which correspond to V 2p3/2 and V 2p1/2, respectively. The V 2p deconvoluted spectra exhibited four different peaks: 516.74 eV and 523.11 eV associated with the V4+ oxidation state and 517.49 eV and 524.43 eV related to the V5+ oxidation state [30,31,32,33,34,35]. The XPS results are consistent with the XRD and EDS results, providing confirmation of the presence and creation of Ni3V2O8-rGO.

3.4. Electrochemical Study

The distinctive morphology of Ni3V2O8 nanoparticles and Ni3V2O8-rGO composites, synthesized through the utilization of ammonium fluoride and urea, and an electrochemical study of the prepared electrode were investigated using a three-electrode setup for the purpose of energy storage in a 2 M KOH electrolyte. The data exhibit the CV, GCD, and EIS spectra of the Ni3V2O8 nanoparticles and Ni3V2O8-rGO composites.
The CV measurements were taken at a different scan rate of 5–100 mV/s, ranging from −0.1 to 0.5 V, with the Ag/AgCl electrode as the reference, as shown in Figure 4c,e. The Ni3V2O8 nanoparticles and Ni3V2O8-rGO composite electrodes both display two prominent redox peaks for oxidation and reduction, which indicate the characteristic electrochemical feature of a faradic redox process. The data illustrate the cyclic voltammetry (CV) characteristics of the Ni3V2O8 nanoparticles and Ni3V2O8-rGO composites as the scan rate increases from 5 to 100 mV/s. The two redox peaks are distinctly detected, with the anodic peak gradually shifting towards a positive direction and the cathodic peak shifting towards a negative direction, in addition to an increase in the peak current with the rise in scan rate. The Ni3V2O8-rGO composite electrode exhibits enhanced peak currents and a greater area under the CV profile, as shown in Figure 4a, suggesting faster faradic processes, reduced interfacial resistance, and increased specific capacitance in comparison to the Ni3V2O8 electrode. The enhancement is probably a result of the augmented conductivity of the electrode material facilitated by the rGO. The specific capacitances of the Ni3V2O8 nanoparticles and Ni3V2O8-rGO composite electrodes were calculated by analyzing the GCD profiles of the electrodes. Figure 4b illustrates the contrasting GCD patterns of the Ni3V2O8 nanoparticles and Ni3V2O8-rGO composites. The Ni3V2O8-rGO composite electrode demonstrates extended durations for both charging and discharging in comparison to the Ni3V2O8 electrode. The addition of rGO in the Ni3V2O8-rGO composite results in a greater specific capacitance compared to the Ni3V2O8 electrode. The electrodes’ specific capacitance (Cs), specific energy density (EDs), and power density (PDs) were determined using Equations (1)–(3) [36,37,38]. The measured specific capacitance was determined to be 132 F/g, with an energy density of 5.04 Wh/kg and a power density of 187.14 W/kg at a current density of 1 mA/cm2. Figure 4d,f also display the GCD profiles of the Ni3V2O8 nanoparticles and Ni3V2O8-rGO composite electrodes under various current densities.
Specific   Capacitance   ( C s ) = I × t m × V
Energy   Density   ( E D s ) = C s × V 2 7.2
Power   Density   ( P D s ) = E D s × 3600 t
where I,   t , m   a n d V are the current density, discharging time, loading mass and potential window of the GCD profile.
The composite electrode benefits from the marigold-like surface microstructure and the incorporation of rGO belts, resulting in improved capacitive performance and an impressive electrochemical energy storage system. In addition, the electrochemical impedance spectroscopy (EIS) technique was used to investigate the reaction kinetics and conductivity of both the Ni3V2O8 nanoparticle and Ni3V2O8-rGO composite electrodes.
This analysis sought to assess the interaction between the electrode and electrolyte. The EIS spectra of the Ni3V2O8 nanoparticle and Ni3V2O8-rGO composite electrodes were measured in the frequency range of 1 MΩ to 0.1 Ω at an AC voltage of 5 mV, as shown in Figure 5. The region where the EIS spectrum intersects the x-axis represents the series resistance (Rs) of the electrodes. The Ni3V2O8 and Ni3V2O8-rGO composite electrodes had series resistance values of 2.21 Ω and 1.65 Ω, respectively. The Ni3V2O8-rGO composite electrode exhibits superior electrochemical performance compared to the Ni3V2O8 electrode. This is likely attributed to the presence of rGO belts, which enhance conductivity by overlapping the marigold surface microstructure, as observed in the FESEM microstructure analysis. The linear trend observed in the EIS spectra signifies the ideal capacitive behavior and enhanced ion diffusion in the electrolyte. In addition, the kinetics of the electrochemical process were investigated by analyzing the b value, transfer coefficient (α), standard rate constant (k0), and diffusion coefficient (D) of the electrode material [39,40,41,42,43]. The CV profiles of the Ni3V2O8 and Ni3V2O8-rGO composite electrodes were utilized for the computation of these quantities. The b value of the electrodes was determined by plotting the logarithm of the current density (ip) against the logarithm of the scan rate (v), as shown in Figure 6a,b and Equation (4) [39,40]. The slope of this plot corresponds to the b value. The b value denotes the character of the present contribution, where a value of 0.5 signifies diffusion and a value of 1.0 signifies capacitive contributions.
i = a v b
i v = 2.69 × A × C × D × n
i p = 0.227 A C F n k 0 e x p n F R T ( E p E 0 )
where b, A, C, n, D, F, R, T, Ep and E0 are the b value, area of the electrode, concentration of the electrolyte, number of electron present, diffusion coefficient, Faraday’s constant, universal gas constant, temperature, peak potential and formal potential.
The b value for the Ni3V2O8-rGO electrode was measured to be 0.71 and 0.63, but the Ni3V2O8-rGO composite exhibited values of 0.67 and 0.57 for the cathodic and anodic peaks, respectively. Both values fall within the range of 0.5 to 1.0, indicating a blend of capacitive and diffusion influences. The Ni3V2O8-rGO composite has a greater diffusion contribution, potentially attributed to the marigold surface microstructure with rGO bands that facilitate ion diffusion throughout the charging and discharging cycle. In supercapacitors, where the current distribution in charge storage either occurs through capacitive or diffusion contributions, ion diffusion is an essential concept for energy storage applications. Supercapacitors are used to store energy. Ions contribute to the diffusion process, which is represented by the diffusion coefficient. This process is what causes the diffusion contribution. The diffusion coefficients of the electrode material were estimated from the CV profile by plotting the peak current (ip) versus the square root of the scan rate (v). This results in the calculation of the diffusion coefficient. Both the cathodic and anodic peaks for both electrodes are depicted in Figure 6c,d as a plot of the peak current vs. the square root of the scan rate. Equation (5) is used to compute the diffusion coefficient, which results in the values of 2.28 × 10−8 cm2/S and 2.45 × 10−8 cm2/S for the Ni3V2O8 electrode and 5.27 × 10−8 cm2/S and 8.3.7 × 10−8 cm2/S for the Ni3V2O8-rGO composite for the anodic and cathodic peaks, respectively, during the computation [39,41]. It is likely that the marigold surface microstructure, which coincides with the wrapping of the rGO belt, contributes for the higher diffusion coefficient of the Ni3V2O8-rGO composite electrode. Through the utilization of the standard rate constant (k0) and the transfer coefficient (α), one may estimate the reaction that occurs between the electrode and the electrolyte. To determine the k0 value and the transfer coefficient, Figure 6e,f are used alongside with Equation (6) for the computation. Figure 6e,f illustrate the relationship between the ln(ip) and Ep-E0 for both the cathodic and anodic values of both electrodes. Redox reactions, which can be reversible, irreversible, or quasi-reversible, are mechanisms that are responsible for charge storage. k0 is less than or equal to 10−5 for reversible redox reactions, k0 is more than or equal to 10−5 for irreversible redox reactions, and 10−1 is less than or equal to k0 for quasi-reversible redox processes [39,41,42,43,44]. The k0 values for the Ni3V2O8 electrode are 2 × 10−5 cm/S and 2.89 × 10−5 cm/S, and for the Ni3V2O8-rGO composite electrode, 6.07 × 10−5 cm/S and 4.82 × 10−5 cm/S for the anodic and cathodic peaks, respectively. Indicating that both electrodes display redox reactions through a mechanism that is quasi-reversible, these values are within the range of redox reactions that are considered to be quasi-reversible. Quasi-reversible redox processes are also represented by the transfer coefficient, which can range from 0.0 to 1.0 degrees. Both of the electrodes exhibit a quasi-reversible redox nature, as evidenced by the transfer coefficients of 0.76 and 0.52, respectively, for the Ni3V2O8 and Ni3V2O8-rGO composite forms of the material.

3.5. Asymmetric Supercapacitor (ASC)

The Ni3V2O8-rGO composite electrode demonstrates remarkable electrochemical performance in a three-electrode configuration study. To explore its practical application, the Ni3V2O8-rGO composite electrode was further investigated using a two-electrode configuration. An asymmetric supercapacitor (ASC) device was fabricated using the Ni3V2O8-rGO composite as the positive electrode and activated carbon (AC) as the negative electrode. The ASC assembly involved wrapping the Ni3V2O8-rGO composite electrode and the AC electrode together with paraffin paper, sandwiching filter paper between them, and using 2M KOH as the electrolyte. The electrochemical performance of the ASC was evaluated through cyclic voltammetry (CV) and galvanostatic charge–discharge (GCD) measurements at different scan rates, current densities, and potentials.
Figure 7a,b illustrate the CV and GCD profiles of the ASC device across different potential ranges from 1.2 to 1.8 V. The CV profile remained stable without any noticeable changes, confirming that the optimal potential for the ASC is 1.8 V. Maintaining a potential of 1.8 V, the rate capability of the ASC was studied by performing CV measurements at various scan rates ranging from 10 to 100 mV/s, as shown in Figure 7c. The consistent increase in current with higher scan rates, without any alteration in the CV profile shape, demonstrated the strong capacitive performance of the device. Figure 7d shows GCD profiles of the ASC at a constant potential of 1.8 V with different current densities ranging from 8 to 14 mA/cm2. The absence of significant distortion with increasing current density indicates that the ASC can operate stably within the potential range of 1.2–1.8 V. Using Equations (1)–(3), the specific capacitance (Cs), energy density (EDs), and power density (PDs) of the ASC were calculated [38,39,40]. The ASC exhibited a specific capacitance of 7.85 F/g, an energy density of 3.56 Wh/kg, and a power density of 225 W/kg at a current density of 8 mA/cm2. The specific capacitance of the ASC decreased with increasing current density due to the non-ideal diffusion nature, which limited the number of ions participating in the process over a given time while using less active material. Electrochemical impedance spectroscopy (EIS) was employed to analyze the electrode and electrolyte interaction of the ASC device before and after stability testing. Figure 7e shows the EIS spectra of the ASC before and after stability testing. In the high-frequency region, the intercept at the x-axis indicates the series resistance (Rs). The ASC exhibited Rs values of 4.27 Ω before stability and 6.63 Ω after stability. The increase in Rs after stability might be attributed to surface modification following prolonged cycling. Additionally, EIS analysis provided insights into the charge transfer resistance (Rct) and the Warburg impedance, which related to ion diffusion. Before stability testing, the Nyquist plot of the ASC showed a smaller semicircle at high frequencies, indicating lower Rct, and a steeper slope in the low-frequency region, signifying efficient ion diffusion.
After 10,000 cycles, the semicircle’s diameter increased, and the slope in the low-frequency region decreased, suggesting that prolonged cycling slightly impeded the charge transfer and ion diffusion processes. The long-term practical capability of the ASC was assessed by performing GCD cyclic stability tests over 10,000 cycles. Figure 7f illustrates the cyclic stability of the ASC device. Initially, the capacitance of the ASC increased up to approximately 2000 cycles, likely due to the activation of the electrode material and the diffusion of ions within the electrode. Subsequently, the discharge capacity decreased with further cycling, possibly due to surface microstructure modification during the charging and discharging processes. After 7000 GCD cycles, only a minor decrease in discharge capacity was observed, resulting in approximately 87% cyclic stability over 10,000 cycles. The columbic efficiency of the ASC device was measured at 99% over 10,000 cycles in Figure 7f, indicating excellent cyclic stability and columbic efficiency, making the ASC a potential candidate for energy storage applications. The improved performance of the Ni3V2O8-rGO composite electrode in the ASC device can be attributed to the synergistic effects of the Ni3V2O8 nanoparticles and the rGO sheets. The rGO sheets enhance electrical conductivity and provide a robust framework for the Ni3V2O8 nanoparticles, facilitating faster electron transport and ion diffusion. The marigold-like surface microstructure of the Ni3V2O8 nanoparticles, combined with the high surface area and excellent conductivity of the rGO sheets, results in superior electrochemical performance. Furthermore, the ASC’s performance metrics, such as specific capacitance, energy density, and power density, are highly competitive with other state-of-the-art supercapacitor reported in the literature. The specific capacitance of 7.85 F/g, energy density of 3.56 Wh/kg, and power density of 225 W/kg underscore the potential of the Ni3V2O8-rGO composite electrode for high-performance energy storage devices. The outstanding cyclic stability and columbic efficiency of the ASC device, maintaining 87% capacitance retention and 99% efficiency over 10,000 cycles, highlight its robustness and reliability for long-term applications. The slight increase in series resistance and charge transfer resistance after prolonged cycling suggests that while the device remains highly efficient, further optimization of the electrode materials and device architecture could further enhance stability and performance. The Ni3V2O8-rGO composite electrode exhibits exceptional electrochemical properties in both three-electrode and two-electrode configurations. The successful integration of Ni3V2O8 nanoparticles with rGO sheets leads to enhanced capacitive performance, high energy and power densities, and excellent long-term stability. The development of ASCs using these composite electrodes holds significant promise for advanced energy storage systems, offering a viable solution for various applications requiring an efficient and durable supercapacitor. The comparative study were listed in Table 1.

4. Conclusions

The method of hydrothermal synthesis and comprehensive characterization of Ni3V2O8 and Ni3V2O8-reduced graphene oxide (Ni3V2O8-rGO) showed significant advances in the development of electrode materials for supercapacitors. An X-ray diffraction (XRD) investigation verified that the produced materials have a crystalline structure. Furthermore, field emission scanning electron microscopy (FESEM) shows a unique marigold-like shape, which suggests that the materials have a large surface area that is advantageous for electrochemical applications. The XPS examination yielded valuable information about the chemical states, thus enhancing our understanding of the material’s composition. The electrochemical investigations demonstrated the exceptional efficiency of Ni3V2O8-rGO, which displayed a specific capacitance of 132 F/g, an energy density of 5.04 Wh/kg, and a power density of 187 W/kg. The effective electrochemical kinetics of the system were highlighted by key metrics, including a b value of 0.67, a transfer coefficient of 0.52, a standard rate constant of 6.07 × 10−5 cm2/S, a diffusion coefficient of 5.27 × 10−8 cm/s, and a series resistance of 1.65 Ω. An asymmetric supercapacitor was constructed by combining Ni3V2O8-rGO and activated carbon (AC), resulting in a specific capacitance of 7.85 F/g, an energy density of 3.52 Wh/kg, and a power density of 225 W/kg. The device exhibited exceptional cyclic stability, with a columbic efficiency of 99% and 87% energy retention after 10,000 cycles. The series resistance increased moderately from 4.27 Ω to 6.63 Ω. The results confirm that Ni3V2O8-rGO has the capability to be used as an excellent electrode material, which can be applied in advanced energy storage systems of the future.

Author Contributions

M.A.Y.—Conceptualization, methodology, P.J.M.—validation software, V.K.—formal analysis, A.M.T.—original draft preparation, S.A.B.—writing—review and editing, S.D.D.—data curation and D.-K.S.—supervision funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Technology Innovation Program (#20010170), funded by the Ministry of Trade, Industry, and Energy (MOTIE, the Republic of Korea).

Data Availability Statement

All the relevant data are included in this published article.

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] [PubMed]
  2. Lokhande, C.D.; Dubal, D.P.; Joo, O.S. Metal oxide thin film based supercapacitors. Curr. Appl. Phys. 2011, 11, 255–270. [Google Scholar] [CrossRef]
  3. Sagadevan, S.; Marlinda, A.R.; Johan, M.R.; Umar, A.; Fouad, H.; Alothman, O.Y.; Khaled, U.; Akhtar, M.S.; Shahid, M.M. Reduced graphene/nanostructured cobalt oxide nanocomposite for enhanced electrochemical performance of supercapacitor applications. J. Colloid. Interface Sci. 2019, 558, 68–77. [Google Scholar] [CrossRef]
  4. Yang, B.; Wang, J.; Zhang, X.; Wang, J.; Shu, H.; Li, S.; He, T.; Lan, C.; Yu, T. Applications of battery/supercapacitor hybrid energy storage systems for electric vehicles using perturbation observer based robust control. J. Power Sources 2020, 448, 227444. [Google Scholar] [CrossRef]
  5. Husain, A.; Ahmad, S.; Mohammad, F. Electrical conductivity and alcohol sensing studies on polythiophene/tin oxide nanocomposites. J. Sci. Adv. Mater. Devices 2020, 5, 84–94. [Google Scholar] [CrossRef]
  6. Nasir, F.; Mohammad, M.A. Investigation of Device Dimensions on Electric Double Layer Microsupercapacitor Performance and Operating Mechanism. IEEE Access 2020, 8, 28367–28374. [Google Scholar] [CrossRef]
  7. Veerakumar, P.; Sangili, A.; Manavalan, S.; Thanasekaran, P.; Lin, K.C. Research Progress on Porous Carbon Supported Metal/Metal Oxide Nanomaterials for Supercapacitor Electrode Applications. Ind. Eng. Chem. Res. 2020, 59, 6347–6374. [Google Scholar] [CrossRef]
  8. Zhang, Y.; Feng, H.; Wu, X.; Wang, L.; Zhang, A.; Xia, T.; Dong, H.; Li, X.; Zhang, L. Progress of electrochemical capacitor electrode materials: A review. Int. J. Hydrogen Energy 2009, 34, 4889–4899. [Google Scholar] [CrossRef]
  9. Bi, W.; Wang, J.; Jahrman, E.P.; Seidler, G.T.; Gao, G.; Wu, G.; Cao, G. Interface Engineering V2O5 Nanofibers for High-Energy and Durable Supercapacitors. Small 2019, 15, e1901747. [Google Scholar] [CrossRef]
  10. Fu, M.; Zhuang, Q.; Zhu, Z.; Zhang, Z.; Chen, W.; Liu, Q.; Yu, H. Facile synthesis of V2O5/graphene composites as advanced electrode materials in supercapacitors. J. Alloys Compd. 2021, 862, 158006. [Google Scholar] [CrossRef]
  11. Tian, M.; Li, R.; Liu, C.; Long, D.; Cao, G. Aqueous Al-Ion Supercapacitor with V2O5 Mesoporous Carbon Electrodes. ACS Appl. Mater. Interfaces 2019, 11, 15573–15580. [Google Scholar] [CrossRef]
  12. Brisse, A.L.; Stevens, P.; Toussaint, G.; Crosnier, O.; Brousse, T. Ni(OH)2 and NiO based composites: Battery type electrode materials for hybrid supercapacitor devices. Materials 2018, 11, 1178. [Google Scholar] [CrossRef] [PubMed]
  13. Shown, I.; Ganguly, A.; Chen, L.C.; Chen, K.H. Conducting polymer-based flexible supercapacitor. Energy Sci. Eng. 2015, 3, 2–26. [Google Scholar] [CrossRef]
  14. Du, X.; Lin, Z.; Wang, X.; Zhang, K.; Hu, H.; Dai, S. Electrode Materials, Structural Design, and Storage Mechanisms in Hybrid Supercapacitors. Molecules 2023, 28, 6432. [Google Scholar] [CrossRef] [PubMed]
  15. Srinivasan, R.; Elaiyappillai, E.; Anandaraj, S.; Duvaragan, B.K.; Johnson, P.M. Study on the electrochemical behavior of BiVO4/PANI composite as a high performance supercapacitor material with excellent cyclic stability. J. Electroanal. Chem. 2020, 861, 113972. [Google Scholar] [CrossRef]
  16. Packiaraj, R.; Venkatesh, K.S.; Devendran, P.; Bahadur, S.A. Nallamuthu, Structural, morphological and electrochemical studies of nanostructured BiVO4 for supercapacitor application. Mater. Sci. Semicond. Process. 2020, 115, 105122. [Google Scholar] [CrossRef]
  17. Choi, J.; Lim, J.; Kim, D.; Park, S.; Yan, B.; Ko, D.; Cho, Y.; Lee, L.Y.S.; Piao, Y. CoFe Prussian blue analogues on 3D porous N-doped carbon nanosheets boost the intercalation kinetics for a high-performance quasi-solid-state hybrid capacitor. J. Mater. Chem. A 2022, 10, 14501–14512. [Google Scholar] [CrossRef]
  18. Srivastava, M.; Das, A.K.; Khanra, P.; Uddin, M.E.; Kim, N.H.; Lee, J.H. Characterizations of in situ grown ceria nanoparticles on reduced graphene oxide as a catalyst for the electrooxidation of hydrazine. J. Mater. Chem. A 2013, 1, 9792–9801. [Google Scholar] [CrossRef]
  19. Zakharova, G.S.; Liu, Y.; Enyashin, A.N.; Yang, X.; Zhou, J.; Jin, W.; Chen, W. Metal cations doped vanadium oxide nanotubes: Synthesis, electronic structure, and gas sensing properties. Sensors Actuators B Chem. 2018, 256, 1021–1029. [Google Scholar] [CrossRef]
  20. Chen, T.W.; Rajaji, U.; Chen, S.M.; Lou, B.S.; Al-Zaqri, N.; Alsalme, A.; Alharthi, F.A.; Lee, S.Y.; Chang, W.H. A sensitive electrochemical determination of chemotherapy agent using graphitic carbon nitride covered vanadium oxide nanocomposite; sonochemical approach. Ultrason. Sonochem. 2019, 58, 104664. [Google Scholar] [CrossRef]
  21. Balasubramaniana, K.; Karuppiahb, C.; Alagarsamyc, S.; Mohandossd, S.; Arunachalame, P.; Govindasamyf, C.; Velmurugang, M.; Yangh, C.-C.; Leeb, H.J.; Ramaraj, S.K. Highly sensitive detection of environmental toxic fenitrothion in fruits and water using a porous graphene oxide nanosheets based disposable sensor. Environ. Res. 2024, 259, 118344. [Google Scholar] [CrossRef] [PubMed]
  22. Zhang, Q.; Xu, Y.; Li, C.; Chen, W.; Zhu, W.; Wang, L. Staggered nickel–vanadium layered double hydroxide nanosheets on reduced graphene oxide via in-situ growth for enhanced supercapacitor performance. J. Alloys Compd. 2023, 935, 168048. [Google Scholar] [CrossRef]
  23. Wu, X.; Han, Z.; Zheng, X.; Yao, S.; Yang, X.; Zhai, T. Core-shell structured Co3O4@NiCo2O4 electrodes grown on flexible carbon fibers with superior electrochemical properties. Nano Energy 2017, 31, 410–417. [Google Scholar] [CrossRef]
  24. Zhang, Z.; Huang, Y.; Yan, J.; Li, C.; Chen, X.; Zhu, Y. A facile synthesis of 3D flower-like NiCo2O4@MnO2 composites as an anode material for Li-ion batteries. Appl. Surf. Sci. 2019, 473, 266–274. [Google Scholar] [CrossRef]
  25. Wang, R.; Xu, C.; Lee, J.M. High performance asymmetric supercapacitors: New NiOOH nanosheet/graphene hydrogels and pure graphene hydrogels. Nano Energy 2016, 19, 210–221. [Google Scholar] [CrossRef]
  26. Mohamed, S.G.; Hussain, I.; Shim, J.J. One-step synthesis of hollow C-NiCo2S4 nanostructures for high-performance supercapacitor electrodes. Nanoscale 2018, 10, 6620–6628. [Google Scholar] [CrossRef] [PubMed]
  27. Zhang, P.; He, H.; Li, Q. Ni–Co–S nanosheets supported by CuCo2O4 nanowires for ultra-high capacitance hybrid supercapacitor electrode. Int. J. Hydrogen Energy 2020, 45, 4784–4792. [Google Scholar] [CrossRef]
  28. Zhang, J.; Liu, F.; Cheng, J.P.; Zhang, X.B. Binary Nickel-Cobalt Oxides Electrode Materials for High-Performance Supercapacitors: Influence of its Composition and Porous Nature. ACS Appl. Mater. Interfaces 2015, 7, 17630–17640. [Google Scholar] [CrossRef]
  29. Zhang, Y.; Xu, X.; Gao, Y.; Gao, D.; Wei, Z.; Wu, H. In-situ synthesis of three-dimensionally flower-like Ni3V2O8@carbon nanotubes composite through self-assembling for high performance asymmetric supercapacitors. J. Power Sources 2020, 455, 227985. [Google Scholar] [CrossRef]
  30. Maitra, A.; Das, A.K.; Karan, S.K.; Paria, S.; Bera, R.; Khatua, B.B. A Mesoporous High-Performance Supercapacitor Electrode Based on Polypyrrole Wrapped Iron Oxide Decorated Nanostructured Cobalt Vanadium Oxide Hydrate with Enhanced Electrochemical Capacitance. Ind. Eng. Chem. Res. 2017, 56, 2444–2457. [Google Scholar] [CrossRef]
  31. Sahoo, R.; Roy, A.; Dutta, S.; Ray, C.; Aditya, T.; Pal, A.; Pal, T. Liquor ammonia mediated V(v) insertion in thin Co3O4 sheets for improved pseudocapacitors with high energy density and high specific capacitance value. Chem. Commun. 2015, 51, 15986–15989. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, H.; Lei, L.; Li, Z.; Huang, L.; Wei, T.; Wang, C.; Wei, X. Surface Defect Modulation with Intercalation Ion Doping Vanadium Oxide to Enhance Zinc Storage Performance. Energy Fuels 2022, 36, 2872–2879. [Google Scholar] [CrossRef]
  33. Kim, K.H.; Yoshihara, Y.; Abe, Y.; Kawamura, M.; Kiba, T. Preparation of porous zinc oxide-nickel oxide nanocomposites by facile one-pot solution process. Ceram. Int. 2017, 43, 1318–1322. [Google Scholar] [CrossRef]
  34. Yang, M.; Fu, X.; Zhang, J.; Wang, Z.; Wang, B.; He, L.; Wu, Z.; Cheng, H.; Pan, H.; Lu, Z. Hierarchical Ultrafine Ni3V2O8 Nanoparticles Anchored on rGO as High-Performance Anode Materials for Lithium-Ion Batteries. Energy Technol. 2019, 7, 1800784. [Google Scholar] [CrossRef]
  35. Yewale, M.A.; Bharadwaj, A.V.S.L.S.; Kadam, R.A.; Shelke, N.T.; Teli, A.M.; Beknalkar, S.A.; Kumar, V.; Alam, M.W.; Shin, D.K. Electrochemical study of CoV2O6 prepared by hydrothermal approach at different molar concentration of vanadium source. Mater. Sci. Eng. B 2024, 307, 117464. [Google Scholar] [CrossRef]
  36. Liu, T.-C.; Pell, W.G.; Conway, B.E.; Roberson, S.L. Behavior of Molybdenum Nitrides as Materials for Electrochemical Capacitors: Comparison with Ruthenium Oxide. J. Electrochem. Soc. 1998, 145, 1882–1888. [Google Scholar] [CrossRef]
  37. Patil, A.M.; Moon, S.; Jadhav, A.A.; Hong, J.; Kang, K.; Jun, S.C. Modifying Electronic Structure of Cation-Exchanged Bimetallic Sulfide/Metal Oxide Heterostructure through In Situ Inclusion of Silver (Ag) Nanoparticles for Extrinsic Pseudocapacitor. Adv. Funct. Mater. 2023, 33, 2305264. [Google Scholar] [CrossRef]
  38. Dhas, S.D.; Thonge, P.N.; Waghmare, S.D.; Kulkarni, G.K.; Shinde, S.K.; Kim, D.Y.; Patil, T.M.; Yewale, M.A.; Moholkar, A.V.; Kim, D. Hierarchical spinel NiMn2O4 nanostructures anchored on 3-D nickel foam as stable and high-performance supercapacitor electrode material. J. Energy Storage 2023, 71, 108168. [Google Scholar] [CrossRef]
  39. Frenzel, N.; Hartley, J.; Frisch, G. Voltammetric and spectroscopic study of ferrocene and hexacyanoferrate and the suitability of their redox couples as internal standards in ionic liquids. Phys. Chem. Chem. Phys. 2017, 19, 28841–28852. [Google Scholar] [CrossRef]
  40. Auti, P.S.; Yewale, M.A.; Kadam, R.A.; Kumar, R.; Nakate, U.T.; Teli, A.M.; Jadhavar, A.A.; Kumar, V.; Warule, S.S.; Shin, D.K. Materials Science & Engineering B Hydrothermally synthesized MnCo2O4 nanoparticles for advanced energy storage applications. Mater. Sci. Eng. B 2024, 301, 117198. [Google Scholar] [CrossRef]
  41. Conway, B.E. Electrochemical supercapacitors. In Advances in Lithium-Ion Batteries; Springer: Boston, MA, USA, 2002; pp. 481–505. [Google Scholar] [CrossRef]
  42. Lindström, H.; Södergren, S.; Solbrand, A.; Rensmo, H.; Hjelm, J.; Hagfeldt, A.; Lindquist, S.E. Li+ ion insertion in TiO2 (anatase). 1. Chronoamperometry on CVD films and nanoporous films. J. Phys. Chem. B 1997, 101, 7710–7716. [Google Scholar] [CrossRef]
  43. Yewale, M.A.; Teli, A.M.; Beknalkar, S.A.; Kumar, V.; Shin, D.K. Smart Textile Flexible MnCo2O4 Electrodes: Urea Surface Modification for Improved Electrochemical Functionality. Materials 2024, 17, 1866. [Google Scholar] [CrossRef]
  44. Yewale, M.A.; Kumar, V.; Teli, A.M.; Beknalkar, S.A.; Nakate, U.T.; Shin, D.K. Elevating Supercapacitor Performance of Co3O4-g-C3N4 Nanocomposites Fabricated via the Hydrothermal Method. Micromachines 2024, 15, 414. [Google Scholar] [CrossRef]
  45. Noh, S.H.; Lee, H.B.; Lee, K.S.; Lee, H.; Han, T.H. Sub-Second Joule-Heated RuO2-Decorated Nitrogen- and Sulfur-Doped Graphene Fibers for Flexible Fiber-type Supercapacitors. ACS Appl. Mater. Interfaces 2022, 14, 29867–29877. [Google Scholar] [CrossRef]
  46. Zhang, Y.; Lin, B.; Sun, Y.; Han, P.; Wang, J.; Ding, X.; Zhang, X.; Yang, H. MoO2@Cu@C Composites Prepared by Using Polyoxometalates@Metal-Organic Frameworks as Template for All-Solid-State Flexible Supercapacitor. Electrochim. Acta 2016, 188, 490–498. [Google Scholar] [CrossRef]
  47. Lakra, R.; Kumar, R.; Sahoo, P.K.; Sharma, D.; Thatoi, D.; Soam, A. Facile synthesis of cobalt oxide and graphene nanosheets nanocomposite for aqueous supercapacitor application. Carbon. Trends 2022, 7, 100144. [Google Scholar] [CrossRef]
  48. Kedir, C.N.; Salinas-Torres, D.; Quintero-Jaime, A.F.; Benyoucef, A.; Morallon, E. Hydrogels obtained from aniline and piperazine: Synthesis, characterization and their application in hybrid supercapacitors. J. Mol. Struct. 2022, 1248, 131445. [Google Scholar] [CrossRef]
Figure 1. XRD spectra of Ni3V2O8 and Ni3V2O8-rGO nanoparticles.
Figure 1. XRD spectra of Ni3V2O8 and Ni3V2O8-rGO nanoparticles.
Micromachines 15 00930 g001
Figure 2. FESEM micrograph of Ni3V2O8 (ad) and Ni3V2O8-rGO (fi) nanoparticles, EDS spectra with elemental composition of Ni3V2O8 (e) and Ni3V2O8-rGO (j).
Figure 2. FESEM micrograph of Ni3V2O8 (ad) and Ni3V2O8-rGO (fi) nanoparticles, EDS spectra with elemental composition of Ni3V2O8 (e) and Ni3V2O8-rGO (j).
Micromachines 15 00930 g002
Figure 3. XPS spectra Ni3V2O8-rGO, survey scan (a), Ni 2p (b), C 1s (c), O 1s (d), V 2p (e).
Figure 3. XPS spectra Ni3V2O8-rGO, survey scan (a), Ni 2p (b), C 1s (c), O 1s (d), V 2p (e).
Micromachines 15 00930 g003
Figure 4. Comparative CV (a), GCD (b) profile, Ni3V2O8 and Ni3V2O8-rGO, CV (c,e) profile at different scan rate, GCD (d,f) at different current density of Ni3V2O8 and Ni3V2O8-rGO electrode.
Figure 4. Comparative CV (a), GCD (b) profile, Ni3V2O8 and Ni3V2O8-rGO, CV (c,e) profile at different scan rate, GCD (d,f) at different current density of Ni3V2O8 and Ni3V2O8-rGO electrode.
Micromachines 15 00930 g004
Figure 5. Electrochemical impedance spectra of Ni3V2O8 and Ni3V2O8-rGO electrode.
Figure 5. Electrochemical impedance spectra of Ni3V2O8 and Ni3V2O8-rGO electrode.
Micromachines 15 00930 g005
Figure 6. Log (ip) vs. Log (v) (a,b), ip vs. v (c,d) and ln (ip) vs. Epc E0 of Ni3V2O8 and Ni3V2O8-Rgo.
Figure 6. Log (ip) vs. Log (v) (a,b), ip vs. v (c,d) and ln (ip) vs. Epc E0 of Ni3V2O8 and Ni3V2O8-Rgo.
Micromachines 15 00930 g006
Figure 7. CV (a) profile at different potentials, GCD (b) profile at different potentials, CV (c) profile at different scan rates, GCD (d) profile at different current densities, EIS (e) spectra before and after stability, (f) stability and columbic efficiency over 10,000 GCD cycles, (g) Ragone plot of asymmetric supercapacitor.
Figure 7. CV (a) profile at different potentials, GCD (b) profile at different potentials, CV (c) profile at different scan rates, GCD (d) profile at different current densities, EIS (e) spectra before and after stability, (f) stability and columbic efficiency over 10,000 GCD cycles, (g) Ragone plot of asymmetric supercapacitor.
Micromachines 15 00930 g007
Table 1. Comparative study of the ASC.
Table 1. Comparative study of the ASC.
Sr. No.MaterialEnergy Density
(Wh/kg)
Power Density
(W/kg)
Reference
1.Ni3V2O8-rGO/AC3.56225Present work
2.NS-GF@RuO2//NS-GF@RuO22.931428[45]
3.MoO2@Cu@C2.5886[46]
4.Co3O4/graphene2.4300[47]
5.poly(Ani-co-Pip)/Vu@PSS1.190.018[48]
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

Yewale, M.A.; Morankar, P.J.; Kumar, V.; Teli., A.M.; Beknalkar, S.A.; Dhas, S.D.; Shin, D.-K. Ni3V2O8 Marigold Structures with rGO Coating for Enhanced Supercapacitor Performance. Micromachines 2024, 15, 930. https://doi.org/10.3390/mi15070930

AMA Style

Yewale MA, Morankar PJ, Kumar V, Teli. AM, Beknalkar SA, Dhas SD, Shin D-K. Ni3V2O8 Marigold Structures with rGO Coating for Enhanced Supercapacitor Performance. Micromachines. 2024; 15(7):930. https://doi.org/10.3390/mi15070930

Chicago/Turabian Style

Yewale, Manesh A., Pritam J. Morankar, Vineet Kumar, Aviraj M. Teli., Sonali A. Beknalkar, Suprimkumar D. Dhas, and Dong-Kil Shin. 2024. "Ni3V2O8 Marigold Structures with rGO Coating for Enhanced Supercapacitor Performance" Micromachines 15, no. 7: 930. https://doi.org/10.3390/mi15070930

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