Next Article in Journal
Analysis of the Measurement of Transportation Carbon Emissions and the Emission Reduction Path in the Yangtze River Economic Belt under the Background of “Dual Carbon” Goals
Previous Article in Journal
Temperature Variations in Deep Thermal Well LZT-1 in Lądek-Zdrój (Bohemian Massif; SW Poland)—Evidence of Geothermal Anomaly and Paleoclimatic Changes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Superior Electrochemical Performance and Cyclic Stability of WS2@CoMgS//AC Composite on the Nickel-Foam for Asymmetric Supercapacitor Devices

1
Department of Physics, Riphah International University, Campus Lahore, Lahore 54000, Pakistan
2
Department of Electrical and Biological Physics, Kwangwoon University, Seoul 01897, Republic of Korea
3
Department of Physics, College of Science, Princess Nourah Bint Abdulrahman University, P.O. Box 84428, Riyadh 11671, Saudi Arabia
*
Author to whom correspondence should be addressed.
Energies 2024, 17(14), 3363; https://doi.org/10.3390/en17143363
Submission received: 4 June 2024 / Revised: 24 June 2024 / Accepted: 1 July 2024 / Published: 9 July 2024
(This article belongs to the Section F: Electrical Engineering)

Abstract

:
Two-dimensional (2D) sulfide-based transition metal dichalcogenides (TMDs) have shown their crucial importance in energy storage devices. In this study, the tungsten disulfide (WS2) nanosheets were combined with hydrothermally synthesized cobalt magnesium sulfide (CoMgS) nanocomposite for use as efficient electrodes in supercapattery energy storage devices. The characteristics of the WS2@CoMgS nanocomposite were better than those of the WS2 and CoMgS electrodes. XRD, SEM, and BET analyses were performed on the nanocomposite to examine its structure, morphology, and surface area in depth. In three-electrode assemblies, the composite (WS2@CoMgS) electrode showed a high specific capacity of 874.39 C g−1 or 1457.31 F g−1 at 1.5 A g−1. The supercapattery device (WS2@CoMgS//AC) electrode demonstrated a specific capacity of 325 C g−1 with an exceptional rate capability retention of 91% and columbic efficiency of 92% over 7000 cycles, according to electrochemical studies. Additionally, the high energy storage capacity of the WS2@CoMgS composite electrode was proved by structural and morphological investigations.

1. Introduction

The advancement of alternate energy sources that are renewable, power conversion, and storage systems has received a lot of attention due to global concern about the increased use of fossil fuel energy and the environmental issues that go along with it [1,2,3]. Power consumption has skyrocketed because the majority of human appliances are now powered by electricity, which has prompted experts to look into very effective energy storage technologies [4,5,6]. From toys to surgical devices, battery-powered devices have been well-established and extensively utilized. Inferior lifetimes, less rich power densities, longer charging times, heating issues, and a lack of security for the environment are some drawbacks of batteries [7,8,9]. A supercapacitor (SC) is an interesting technology that is one of the many battery alternatives that are accessible. Due to their excellent cycle stability, outstanding power density, and quick charging or discharging execution, supercapacitors (SCs) have drawn a lot of attention [10,11,12,13].
There exist three separate categories of supercapacitors: I. The electrochemical process that retains energies by collecting ions on the exterior of an electrode provides the basis for the charging as well as discharge mechanism of electric double-layer capacitors (EDLCs) [13]. II. The second type of capacitor is a redox capacitor, which utilizes a faradic technique for carrying charge through redox processes on the top of the electrode. Electrolytes and materials used for electrodes undergo quick and recoverable electrochemical alterations in redox capacitors, which is known as the faradic phenomenon. III. To create a hybrid system, a hybrid supercapacitor integrates EDLCs with a faradic electrode substance [14,15,16,17].
It is acknowledged that the morphogenesis and nanostructure characteristics of the 2-D TMD sulfide-based materials used to make supercapacitors have a significant impact on their efficiency [18,19,20,21]. A material’s particular extent of surface, pore-size allocation, and layer thickness are determined by its morphology and nanostructures [22,23]. The electrolyte’s connectivity and its delivery and diffusion pathways, the mechanical properties of the electrodes used for supercapacitor devices, and the active opportunities for electrochemical reactions are all characterized by these characteristics [24,25,26,27]. Therefore, it is highly desired to produce electrodes with enhanced characteristics. According to this perspective, the most promising possibilities are nonporous materials with large surface areas, small pore sizes, and nanoscale particle sizes [28,29]. According to recent research, outstanding specific capacitance along with conductive properties may be achieved by doping or combining one transition metal oxide into another kind of metal composite [22,30]. Fast paths for electrons along with ion transportation, synergistic effects, and higher electron collecting efficiency are only a few of these electrode composite’s considerable competitive benefits [31,32]. Exceptional capacitance accomplishment may be achieved using WO4, MoS2, and nanostructures mixed with metal oxide, which are appropriate for electrochemical capacitors [33,34].
Li et al. synthesized the WS2/NiCo2O4 heterogeneous structures, which have been reinforced by carbon cloth and used to apply supercapacitors with outstanding performance [35]. Tarugu et al. synthesized cauliflower-like ZnWO4/WS2 composite by a simple, cost-effective two-step hydrothermal process, which was created on the top layer of nickel foam and examined as potentially an electrode substrate for supercapacitors with excellent performance [36]. Raghavendra et al. assembled a nanoflower composite WS2/ZnCo2O4 that resembles jasmine petals using a hydrothermal process for improved electrodes in supercapacitors [37].
It is clear from a direct comparison of these important performance indicators that WS2@CoMgS-based electrodes provide a substantial improvement over current materials. The unique properties and promising future of WS2@CoMgS for next-generation supercapatteries are highlighted by its improved specific capacitance, better energy and power densities, superior cycle stability, and outstanding rate capability.
Motivated by the previously mentioned literature, we employed a hydrothermal method to produce a WS2@CoMgS composite in this investigation. Here, we create a simple synthesis method for creating high-surface-area, narrow-pore-size-distribution, nanoporous WS2@CoMgS nanospheres that may be used as efficient electrode material. With its extraordinary specific capacitance (1457.31 F/g at 1.5 A/g), excellent charging capability, and long cycle life, this nanoporous WS2@CoMgS electrode has promise to be a superior performance electrode substance for applications in supercapacitor technology.

2. Experimental Analyses

2.1. Material

Cobalt nitrate (Co (No3)2. 6H2O) with a purity of 99.8%, magnesium chloride (MgCl2) with a purity of 99.9%, sodium sulfide (Na2S) with a purity of99.9%, tungsten disulfide (WS2) with a purity of 99.9%, N-polyvinylidene fluoride (PVDF), activated carbon (AC), and carbon black were purchased from Sigma-Aldrich. These chemicals were utilized without further purification. Distilled water was used during the experiments for synthesizing solutions, cleaning materials, and purging impurities.

2.2. Synthesis of WS2 Nanosheets

The tip-sonication approach was utilized for the synthesis of tungsten disulfide (WS2) nanosheet in a solvent. A sonication probe was utilized to provide localized ultrasonic energy and mix WS2 powder with a solvent of NMP. The sonication probe is then immersed into the mixture, and ultrasonic energy is applied for a specified duration, typically ranging from minutes to hours. During sonication, the energy from the probe breaks apart the WS2 layers, leading to the exfoliation and dispersion of individual or few-layered WS2 nanosheets in the solvent, as shown in Figure 1a.

2.3. Synthesis of CoMgS and WS2@CoMgS Composite

The hydrothermal technique was employed for the synthesis of cobalt magnesium sulfide (CoMgS) because of its superior simplicity and efficacy. The solutions of sodium sulfide (0.50 mM), magnesium chloride (0.50 mM), and cobalt nitrate (0.50 mM) in 20 mL were prepared in the distilled water. A solution of Na2S·9H2O was added dropwise into the CoMg solution while continuously stirring at room temperature. The resulting mixture was then transferred into an autoclave, placed in an oven for 12 h, and heated to 180 °C. Afterward, it was gradually cooled down to room temperature. The obtained solution was thoroughly washed with distilled water and ethanol to remove impurities. Afterward, the product was dried for 1.5 h in an oven at 70 °C after washing. Furthermore, to create the composite material, CoMgS was mixed with WS2 nanosheets in a mass ratio of 75% CoMgS to 25% WS2, as depicted in Figure 1b.

2.4. Preparation of Working Electrode

Electrochemical analysis is a tool used to study the electrochemical presentation of nanostructures. The active material was placed on the nickel foam and utilized as a working electrode (WE) during electrochemical studies. The Hg/HgO was utilized as reference electrode (RE) Hg/HgO, whereas platinum wire is a counter electrode (CE). The slurry involves PVDF of 0.002 g as a binding agent, 0.016 g of active material, and carbon black of 0.002 g. The slurry is continuously stirred for 8 h, creating a homogenous solution suitable for depositing on nickel foam. The active material weight loaded on NF was 0.61 mg.
The fabrication of asymmetric supercapacitors requires the positive and negative electrodes to be charged in an ideal manner for maximum performance. Equivalent charge storage is accomplished on both electrodes using the mass balance equation. In this instance, an asymmetric supercapacitor made of WS2@CoMgS is configured as the positive electrode (+Ve) and activated carbon serves as the negative electrode (−Ve). For constructing a mass balance equation, charges on both electrodes should be equal.
q + = q
where q is the change which is given by the following:
q = C ·   m   ·   V
Here, V is the voltage, and Csp is the specific capacitance of electrodes. From Equation (2), charges stored on positive and negative electrodes are [38].
q + = C + ·   m +   ·   V +   ( for   positive   electrode )
q = C ·   m   ·   V   ( for   negative   electrode )
Equation (1) becomes the following:
C + ·   m +   ·   V + = C ·   m   ·   V
Equation (3) is rearranged as follows:
m + m = C · V C + · V +

2.5. Electrochemical Measurements

An electro-analytical methodology that is frequently used to get precise information on electrochemical responses is cyclic voltammetry (CV). Incorporating a translucent carbon working electrode, a platinum countering electrode, and a Hg/HgO referencing electrode, the three-electrode cell frameworks were used for the CV [39]. The electrochemical behavior of the material can be better understood by using the CV data. The material’s redox peaks, or oxidation as well as reduction peaks, are displayed graphically in a cyclic voltammogram and are used to forecast the capacitor-like behavior of an electrode. Consequently, the potential that allows the component to be reduced and oxidized may be found [40,41]. The mechanisms that may be connected to the WS2@CoMgS composite’s capacitive behavior may involve the existence of a redox system.
CV, GCD, and EIS measurements of as-synthesized (CoMgS and WS2@CoMgS) electrodes were achieved in a 1 M KOH solution. The saturated Hg/HgO, Pt wire, and CoMgS or WS2@CoMgS electrodes were used as referencing, counter, and functioning electrodes, respectively. The CV assessments were carried out among an applied voltage range of 0.0 to +0.6 V at different scanning rates (5, 10, 20, 40, and 60 mV/s). The GCD experiments were conducted at different current densities (1.5–2.3 A/g) throughout the potential range of 0 to + 0.6 V. These equations can be employed to determine the specific capacity (Qs) and capacitance (Csp) of investigated nanocomposite electrodes based on CV curves with varying scanning rates [29,30]:
Q s = 1 m v V i V f ( I × V d V )
C sp = V i V f I × V d V 2 m a ( V f V i )
In Equation (5), specific capacity is represented by Qs, where m is the active mass; V is the scan rate; and V i V f I × V d V is the area. In Equation (6), the potential window is represented by (vfvi) [42]. At numerous current densities (1.5–2.3 A/g), GCD observations were conducted. Equation (7) could potentially be used to determine the specific capacity through GCD’s analysis.
Q s = I × t m
The current and discharge times had been indicated by I and t, respectively, in the above case. We divided Equations (5) and (6) by the potential window to calculate the specific capacitance (F/g). CoMgS and WS2@CoMgS composite’s energy and power densities were computed using the following formulas [43]:
E d = C s   ( V ) 2 × 3.6
P d = E d × 3600 t
In Equation (8), Cs stands for specific capacity determined by GCD, ΔV is the applied voltage, and t is the discharge period at a specified current density in Equation (9) [44,45,46].

3. Results and Discussion

3.1. XRD, SEM, and BET Analysis

The nanocomposite (WS2@CoMgS) was examined using XRD, SEM, and BET testing for structure, surface morphology, and elemental information, respectively. The potentiostat (CS-300.corrTest-China) was used for electrochemical performance. X-ray diffraction (XRD) analysis was used to calculate the structural characteristics of the prepared WS2 and WS2@CoMgS samples. The XRD data of WS2 and WS2@CoMgS composite is shown in Figure 2a. Peaks at plans of (100), (101), (102), and (110) correspond to CoS, whereas peaks at plans of (111), (200), (202), and (311) correspond to MnS for JCPDS-01-075-0605 and JCPDS-65-1919 respectively [47]. XRD of WS2 indicates that the exfoliated nanoflakes have a crystalline hexagonal structure [48]. The size of the crystallite (D) was determined using the Scherrer equation [49].
D = K × λ β C o s θ
The average crystalline size of WS2@CoMgS is 46 nm. The lattice parameters were calculated by analyzing the peak positions in the XRD patterns using Bragg’s law and the equation specific to hexagonal crystal systems. The lattice parameter of the samples was determined using the given equation.
L a t t i c e   P a r a m e t e r = d × h 2 + k 2 + l 2
In this context, h, k, and l represent the Miller Indices, and d denotes the inter-planner spacing as determined by Bragg’s law. The estimated value in WS2 and WS2@CoMgS are 3.18 Angstrom and 8.01 Angstrom. The EDAX of the WS2 and WS2@CoMgS nanocomposite is shown in Figure S1a. The SEM envision in Figure 2b shows the surface structure of the WS2@CoMgS material. The investigation displays that CoS nanoflakes covered with MgS nanoparticles form a structured matrix. Furthermore, configuration-induced aggregation of nanoflakes increases strong ionic conductivity, improving electrolyte ion dispersion. Furthermore, SEM of WS2 demonstrates the development of plane and incessant nanosheets composed of crystalline hexagonal flakes with lateral sizes ranging from 50 to 500 nm. It is also clear that the concentration of the exfoliating agent (SDBS) influences the nanosheet’s separation from the bulk sample, hence increasing the exfoliation process [50]. As a result, in the SEM study of WS2@CoMgS, we detect CoS nanoflakes coated with MgS nanomaterials, as well as a smooth and continuous nanosheet composed of crystalline hexagonal WS2 flakes.
The pore size, pore volume, and surface area of CoMgS are approximately 12.07 nm, 0.013 cm3 g−1, and 11.94 m2 g−1, respectively [51]. BET nitrogen adsorption and desorption isotherm studies improve electrode performance by carefully altering pore size, pore volume, and surface area. The BET analysis properly quantifies sample surface areas. Figure 2c depicts the BET analysis for WS2@CoMgS, including its isotherm. The WS2@CoMgS material had the following properties: pore volume of 11.27 cm3 g−1, pore size of 17.33 nm, and surface area of 13.64 m2 g−1. The increased surface area has a direct impact on both specific capacity and ensuring cyclic stability. These outcomes show that the significant surface area and pore size of WS2@CoMgS allow for fast charge retention. Because of the larger surface area of WS2@CoMgS, the BET study results indicate that there is a greater possibility of charge accumulation on the electrode surface.
An analysis of the profiles of CV for WS2, CoMgS, and WS2@CoMgS at a scan rate of 5 mV/s is shown in Figure 3a. It shows that the WS2@CoMgS electrode’s substrate has a superior power retention capability, as indicated by its larger integral surface and improved redox current responsiveness. This might be the consequence of an increase in the intrinsic resistance of the WS2@CoMgS material, making it more challenging for electrons and ions to get absorbed onto the material’s surface throughout the charging procedure. Additionally, this resistance leads electrons and ions to be quickly expelled from the electrode surface, whereas the discharging process starts. Figure 3b,c displays the as-synthesized cyclic voltammetry (CV) patterns of CoMgS, and WS2@CoMgS electrodes throughout the potential region of 0.0–0.6 V at a scan rate of (5–60) mV/s. A 1 M KOH aqueous solution of electrolytes having a potential of 0.6 V was used for the CV testing.
The CV plots at various scan speeds showed prominent redox peaks attributed to rapid kinetic processes, indicating that the samples acted like batteries. The redox peaks of the WS2@CoMgS electrode migrated in extra positive and negative potential zones, respectively, compared to those of WS2 and CoMgS. This happens because the electro-active material WS2@CoMgS, which is uniformly deposited, has ample time to react with the electrolyte ions and active species present on the Ni-foam substrate. The efficiency of WS2@CoMgS is tested using three different aqueous electrolytes (1 M LiNO3), 1 M NaCl, and 1 M KOH). The outstanding performance of the WS2@CoMgS electrode in 1 M KOH is shown by the CV curves of WS2@CoMgS in dissimilar electrolytes shown in Figure S1.
The current investigation examined the samples’ electrochemical characteristics at various current densities while maintaining a constant voltage. In Figure 3d, the WS2, CoMgS, and WS2@CoMgS composite of GCD is compared at a current density of 5 A/g. The GCD curves of CoMgS and WS2@CoMgS are depicted in Figure 3e,f. The GCD composites of WS2, CoMgS, and WS2@CoMgS are compared at 5 A/g in Figure 3d. The CoMgS and WS2@CoMgS GCD curves are shown in Figure 3e,f. The nonlinearity of galvanostatic charge–discharge (GCD) curves deviates from the anticipated triangular shape [52,53]. The divergence, which signifies redox reactions occurring at the electrode surface, implies that the resulting materials are suitable for battery quality as electrode materials. With an increase in current density, the duration of charge/discharge is reduced. Nevertheless, the curves exhibited a consistent pattern, suggesting the durability of the synthetic materials. The decrease in discharge times can be attributed to the reduction in electrode-specific capacities and shorter contact times resulting from higher current densities.
Figure 4a shows the relationship between anodic and cathodic peak currents (mA), and the value of R2 is found to be 0.9877 and 0.9886, respectively, for the WS2@CoMgS composite [54]. Specific capacities can be calculated using the CV with the help of Equation (1), which reveals that WS2 and CoMgS exhibit 318.84 and 389 C/g at a scan rate of 5 mV/s, respectively. A superior specific capacity of 549.74 C/g is found in the WS2@CoMgS, as shown in Figure 4b. The specific capacity of WS2@CoMgS is increased because of the synergistic impact of CoMgS, higher absorbency, and stronger conductivity. As current increases with higher scan rates, the infrared (IR) loss is elevated [55]. The specific capacitance of WS2, CoMgS, and WS2@CoMgS is determined using CV at various scan rates (5–60 mV/s), as shown in the supplementary information (Figure S2a). The comparison between specific capacity and capacitance is described in Table 1.
Figure 4c shows the specific capacity of WS2@CoMgS, which was determined using these GCD curves. It appeared that at a 1.5 A/g current density, it had a higher specific capacity than both WS2 and CoMgS. The reasons for this enhancement in specific capacity included increased porosity, enormous surface area, and better electrical conductivity [56]. The specific capacitance graph of WS2, CoMgS, and WS2@CoMgS is determined using GCD and is shown in the supplementary information (Figure S2b). A comparison of CV and GCD results of specific capacitance is shown in Figure S2c. The sample was examined using the electrochemical impedance spectroscopy (EIS) technique. EIS is an essential instrument for studying electrode-electrolyte interactions and their mechanisms. This gives data on the internal resistance of the electrode material and the resistance between the electrode and electrolyte. [57]. Figure 4d illustrates the experimental findings, demonstrating the high electroactivity of CoMgS and WS2@CoMgS electrodes and its facilitation of enhanced ion and electron mobility. This indicates that, according to supplemental information in Table S1, WS2@CoMgS has the lowest value compared to the other samples. Compared to earlier values, the supercapattery energy storage device (WS2@CoMgS//AC) performance is significantly more significant.

3.2. Electrochemical Analysis of Supercapattery Device

A two-electrode supercapattery entitled WS2@CoMgS//AC was synthesized by the association of an AC electrode with a WS2@CoMgS electrode. In this configuration, the activated carbon electrode had a negative potential, and the WS2@CoMgS electrode had a positive one; during these experiments, KOH solution was utilized as the electrolyte in a one-molar ratio. The detailed electrochemical measurement of AC is displaced in Figure S4. Figure 5a depicts a simplified representation of the two-electrode arrangement. Figure 5b shows a CV comparison of the two electrodes employed in the supercapattery application. The steady slope seen during cyclic voltammetry (CV) demonstrates the capacitance of activated carbon. In contrast, the prominent peaks in the cyclic voltammetry (CV) of WS2@CoMgS suggest the presence of a battery-grade electrode. The first step in the characterization process was to use cyclic voltammetry to determine the best potential window, which ranged from 0 V to 1.6 V. It was observed that the device’s optimal potential window was 0 to 1.6 V. Next, GCD and CV evaluated device performance. Before characterization, each electrode was conditioned within its potential window. Figure 5c depicts the CV at 5–60 mV/s over 0–1.6 V. The device’s CV response varies dramatically between battery-grade and carbonaceous electrodes. The hybrid CV form proposes surface control energy storage and diffusion. At lower potentials, there are no peaks showing charge buildup by adsorption. Modest peaks at higher potentials reveal a diffusion-controlled charge storage mechanism. GCD was performed between 0 and 1.6 V at various current densities, as shown in Figure 5d. The absence of hump-shaped or triangular GCD profiles indicates that charge storage is influenced by both surface-controlled and diffusion-controlled mechanisms.
Additionally, Qs was determined using GCD at a varied current density of 1.3 to 2.3 A/g, as illustrated in Figure 6a. Additionally, the comparison of the specific capacity of all samples for supercapattery energy storage devices is shown in the supplementary information (Figure S3). A graph depicting the charge/discharge duration was also plotted versus 7000 cycles and is displayed in Figure 6b. After 7000 GCD cycles, the supercapattery device’s long-term stability and columbic efficiency are shown in Figure 6c. The evaluation revealed a remarkable 91% specific capacity as well as a promising 92% columbic efficiency throughout prolonged cycle operations. This result validated the constructed device’s improved cycle durability.
After 7000 cycles, the cyclic stability results of WS2@CoMgS//AC electrodes show higher stability, which justifies thoroughly examining the degradation mechanisms to account for these data. Numerous variables could be responsible for the WS2@CoMgS composite’s increased stability. First, adding CoMgS to the WS2 matrix may lessen cycling-related stress and strain by adding more active sites and improving structural integrity. Secondly, improved electron and ion transport could minimize the deterioration of electrode materials due to the synergistic effects of WS2 and CoMgS. Additionally, CoMgS may keep WS2 layers from clumping together, preserving their large surface area and reliable electrochemical performance [58].
Equations (4) and (5) were utilized to calculate the energy density (Ed) and power density (Pd) of the WS2@CoMgS//AC device, providing an all-encompassing characterization of it. After that, a Ragone plot was made using these values, as shown in Figure 6d. The supercapattery device demonstrated an amazing power density of 876.89 W/kg and a noteworthy energy density of 23.99 Wh/kg at a current density of 1.5 A/g. Table 2 compares the computed Ed and Pd with previously reported values in the same category.

4. Conclusions

In conclusion, WS2@CoMgS, a hydrothermally produced composite, is an extremely effective supercapacitor electrode material. We employed techniques such as SEM, XRD, and BET to evaluate the material’s morphology, crystallinity, and pore size and area. In the SEM study of WS2@CoMgS, we detected CoS nanoflakes coated with MgS nanomaterials, as well as a smooth and continuous nanosheet composed of crystalline hexagonal WS2 flakes. The WS2 nanosheets combine with hydrothermally synthesized CoMgS nanocomposite for use as efficient electrodes in supercapattery devices. The characteristics of the WS2@CoMgS nanocomposite were better than those of the WS2 and CoMgS electrodes. In three-electrode assemblies, the composite WS2@CoMgS electrode shows a high specific capacity of 874.39 C g−1 or 1457.31 F g−1 at 1.5 A g−1. The supercapattery device (WS2@CoMgS//AC) electrode demonstrated a specific capacity of 325 C g−1 with an exceptional rate capability retention of 91% and a columbic efficiency of 92% over 7000 cycles, according to electrochemical studies. Additionally, the high energy storage capacity of the WS2@CoMgS composite electrode was proven by structural and morphological investigations.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/en17143363/s1, Figure S1: EDAX and electrochemical measurement of electrode material in different electrolytes; Figure S2: Comparison of specific capacitance of the samples; Figure S3: Specific capacity of all samples; Figure S4: Electrochemical measurement of activated carbon; Table S1: BET analysis of all the samples.

Author Contributions

Methodology, M.I., A.M.A., M.W.I., S.M. and A.S.A.; Data curation, M.I., A.M.A., M.W.I., S.M. and A.S.A.; Writing—original draft, A.M.A. All authors have read and agreed to the published version of the manuscript.

Funding

The authors declare that no funds or grants were received during the preparation of this manuscript.

Data Availability Statement

The original contributions presented in the study are included in the article/Supplementary Materials; further inquiries can be directed to the corresponding author.

Acknowledgments

The authors appreciate Princess Nourah bint Abdulrahman University Researchers Supporting Project number (PNURSP2024R479), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bose, R.; Patil, B.; Jothi, V.R.; Kim, T.H.; Arunkumar, P.; Ahn, H.; Yi, S.C. Co3Se4 nanosheets embedded on N-CNT as an efficient electroactive material for hydrogen evolution and supercapacitor applications. J. Ind. Eng. Chem. 2018, 65, 62–71. [Google Scholar] [CrossRef]
  2. Jayaraman, T.; Murthy, A.P.; Elakkiya, V.; Chandrasekaran, S.; Nithyadharseni, P.; Khan, Z.; Senthil, R.A.; Shanker, R.; Raghavender, M.; Kuppusami, P.; et al. Recent development on carbon based heterostructures for their applications in energy and environment: A review. J. Ind. Eng. Chem. 2018, 64, 16–59. [Google Scholar] [CrossRef]
  3. Theerthagiri, J.; Karuppasamy, K.; Durai, G.; Rana AU, H.S.; Arunachalam, P.; Sangeetha, K.; Kuppusami, P.; Kim, H.-S. Recent advances in metal chalcogenides (MX; X = S, Se) nanostructures for electrochemical supercapacitor applications: A brief review. Nanomaterials 2018, 8, 256. [Google Scholar] [CrossRef] [PubMed]
  4. Durai, G.; Kuppusami, P.; Theerthagiri, J. Microstructural and supercapacitive properties of reactive magnetron co-sputtered Mo3N2 electrodes: Effects of Cu doping. Mater. Lett. 2018, 220, 201–204. [Google Scholar] [CrossRef]
  5. Thiagarajan, K.; Theerthagiri, J.; Senthil, R.; Madhavan, J. Simple and low cost electrode material based on Ca2V2O7/PANI nanoplatelets for supercapacitor applications. J. Mater. Sci. Mater. Electron. 2017, 28, 17354–17362. [Google Scholar] [CrossRef]
  6. Zhang, X.; Tang, Y.; Zhang, F.; Lee, C.-S. A novel aluminum–graphite dual-ion battery. Adv. Energy Mater. 2016, 6, 1502588. [Google Scholar] [CrossRef]
  7. Sousa, R.E.; Costa, C.M.; Lanceros-Méndez, S. Advances and future challenges in printed batteries. ChemSusChem 2015, 8, 3539–3555. [Google Scholar] [CrossRef] [PubMed]
  8. Dehghani-Sanij, A.; Tharumalingam, E.; Dusseault, M.; Fraser, R. Study of energy storage systems and environmental challenges of batteries. Renew. Sustain. Energy Rev. 2019, 104, 192–208. [Google Scholar] [CrossRef]
  9. Wang, M.; Jiang, C.; Zhang, S.; Song, X.; Tang, Y.; Cheng, H.-M. Reversible calcium alloying enables a practical room-temperature rechargeable calcium-ion battery with a high discharge voltage. Nat. Chem. 2018, 10, 667–672. [Google Scholar] [CrossRef]
  10. Ates, M.; Chebil, A.; Yoruk, O.; Dridi, C.; Turkyilmaz, M. Reliability of electrode materials for supercapacitors and batteries in energy storage applications: A review. Ionics 2022, 28, 27–52. [Google Scholar] [CrossRef]
  11. Raghavendra, K.V.G.; Vinoth, R.; Zeb, K.; Gopi, C.V.M.; Sambasivam, S.; Kummara, M.R.; Obaidat, I.M.; Kim, H.J. An intuitive review of supercapacitors with recent progress and novel device applications. J. Energy Storage 2020, 31, 101652. [Google Scholar] [CrossRef]
  12. Khan, K.; Tareen, A.K.; Aslam, M.; Mahmood, A.; Zhang, Y.; Ouyang, Z.; Guo, Z.; Zhang, H. Going green with batteries and supercapacitor: Two dimensional materials and their nanocomposites based energy storage applications. Prog. Solid State Chem. 2020, 58, 100254. [Google Scholar] [CrossRef]
  13. Mirzaei-Saatlo, M.; Asghari, E.; Shekaari, H.; Pollet, B.G.; Vinodh, R. Performance of ethanolamine-based ionic liquids as novel green electrolytes for the electrochemical energy storage applications. Electrochim. Acta 2024, 474, 143499. [Google Scholar] [CrossRef]
  14. Oyedotun, K.; Madito, M.; Bello, A.; Momodu, D.; Mirghni, A.; Manyala, N. Investigation of graphene oxide nanogel and carbon nanorods as electrode for electrochemical supercapacitor. Electrochim. Acta 2017, 245, 268–278. [Google Scholar] [CrossRef]
  15. Zhu, J.; Zhou, W.; Zhou, Y.; Cheng, X.; Yang, J. Cobalt sulfide/reduced graphene oxide nanocomposite with enhanced performance for supercapacitors. J. Electron. Mater. 2019, 48, 1531–1539. [Google Scholar] [CrossRef]
  16. Li, Y.; Wang, G.; Wei, T.; Fan, Z.; Yan, P. Nitrogen and sulfur co-doped porous carbon nanosheets derived from willow catkin for supercapacitors. Nano Energy 2016, 19, 165–175. [Google Scholar] [CrossRef]
  17. Chen, Q.; Cai, D.; Zhan, H. Construction of reduced graphene oxide nanofibers and cobalt sulfide nanocomposite for pseudocapacitors with enhanced performance. J. Alloys Compd. 2017, 706, 126–132. [Google Scholar] [CrossRef]
  18. Yu, Z.; Tetard, L.; Zhai, L.; Thomas, J. Supercapacitor electrode materials: Nanostructures from 0 to 3 dimensions. Energy Environ. Sci. 2015, 8, 702–730. [Google Scholar] [CrossRef]
  19. Ali, M.; Afzal, A.M.; Iqbal, M.W.; Mumtaz, S.; Imran, M.; Ashraf, F.; Ur Rehman, A.; Muhammad, F. 2D-TMDs based electrode material for supercapacitor applications. Int. J. Energy Res. 2022, 46, 22336–22364. [Google Scholar] [CrossRef]
  20. Imran, M.; Afzal, A.M.; Iqbal, M.W.; Hegazy, H.H.; Iqbal, M.Z.; Mumtaz, S.; Qureshi, R. Manganese (Sulfide/Oxide) based electrode materials advancement in supercapattery devices. Mater. Sci. Semicond. Process. 2023, 158, 107366. [Google Scholar] [CrossRef]
  21. Alsaiari, M.; Imran, M.; Afzal, A.M.; Iqbal, M.W.; Algethami, J.S.; Harraz, F.A. Effect of MoS2 and electrolyte temperature on the electrochemical performance of NiCoS@rGO-based electrode material for energy storage, oxygen reduction reaction and electrochemical glucose sensor. Mater. Today Chem. 2024, 35, 101909. [Google Scholar] [CrossRef]
  22. Yang, J.; Shang, J.; Liu, Q.; Yang, X.; Tan, Y.; Zhao, Y.; Liu, C.; Tang, Y. Variant-Localized High-Concentration Electrolyte without Phase Separation for Low-Temperature Batteries. Angew. Chem. 2024, e202406182. [Google Scholar] [CrossRef]
  23. Xu, X.; Dong, Y.; Hu, Q.; Si, N.; Zhang, C. Electrochemical hydrogen storage materials: State-of-the-art and future perspectives. Energy Fuels 2024, 38, 7579–7613. [Google Scholar] [CrossRef]
  24. Xu, J.; Wang, X.; Wang, X.; Chen, D.; Chen, X.; Li, D.; Shen, G. Three-dimensional structural engineering for energy-storage devices: From microscope to macroscope. ChemElectroChem 2014, 1, 975–1002. [Google Scholar] [CrossRef]
  25. Moosavifard, S.E.; El-Kady, M.F.; Rahmanifar, M.S.; Kaner, R.B.; Mousavi, M.F. Designing 3D highly ordered nanoporous CuO electrodes for high-performance asymmetric supercapacitors. ACS Appl. Mater. Interfaces 2015, 7, 4851–4860. [Google Scholar] [CrossRef]
  26. He, C.; Han, X.; Kong, X.; Jiang, M.; Lei, D.; Lei, X. Fe-doped Co3O4@ C nanoparticles derived from layered double hydroxide used as efficient electrocatalyst for oxygen evolution reaction. J. Energy Chem. 2019, 32, 63–70. [Google Scholar] [CrossRef]
  27. Moosavifard, S.E.; Shamsi, J.; Fani, S.; Kadkhodazade, S. 3D ordered nanoporous NiMoO 4 for high-performance supercapacitor electrode materials. RSC Adv. 2014, 4, 52555–52561. [Google Scholar] [CrossRef]
  28. Wu, X.; He, G.; Ding, Y. Dealloyed nanoporous materials for rechargeable lithium batteries. Electrochem. Energy Rev. 2020, 3, 541–580. [Google Scholar] [CrossRef]
  29. Huang, H.; Shi, H.; Das, P.; Qin, J.; Li, Y.; Wang, X.; Su, F.; Wen, P.; Li, S.; Lu, P. The chemistry and promising applications of graphene and porous graphene materials. Adv. Funct. Mater. 2020, 30, 1909035. [Google Scholar] [CrossRef]
  30. Ikkurthi, K.D.; Rao, S.S.; Jagadeesh, M.; Reddy, A.E.; Anitha, T.; Kim, H.-J. Synthesis of nanostructured metal sulfides via a hydrothermal method and their use as an electrode material for supercapacitors. New J. Chem. 2018, 42, 19183–19192. [Google Scholar] [CrossRef]
  31. Wang, X.; Luo, Z.; Huang, J.; Chen, Z.; Xiang, T.; Feng, Z.; Wang, J.; Wang, S.; Ma, Y.; Yang, H. S/N-co-doped graphite nanosheets exfoliated via three-roll milling for high-performance sodium/potassium ion batteries. J. Mater. Sci. Technol. 2023, 147, 47–55. [Google Scholar] [CrossRef]
  32. Bai, X.; Xu, M.; Li, Q.; Yu, L. Trajectory-battery integrated design and its application to orbital maneuvers with electric pump-fed engines. Adv. Space Res. 2022, 70, 825–841. [Google Scholar] [CrossRef]
  33. Xu, X.; Gao, J.; Huang, G.; Qiu, H.; Wang, Z.; Wu, J.; Pan, Z.; Xing, F. Fabrication of CoWO4@NiWO4 nanocomposites with good supercapacitve performances. Electrochim. Acta 2015, 174, 837–845. [Google Scholar] [CrossRef]
  34. Asad Ur, R.; Afzal, A.M.; Iqbal, M.W.; Imran, M.; Ali, M. Transition Metal Dichalcogenides, Conducting Polymers, and Their Nanocomposites as Supercapacitor Electrode Materials. Polym. Sci. Ser. A 2023, 65, 447–471. [Google Scholar] [CrossRef]
  35. Li, L.; Gao, J.; Cecen, V.; Fan, J.; Shi, P.; Xu, Q.; Min, Y. Hierarchical WS2@NiCo2O4 core–shell heterostructure arrays supported on carbon cloth as high-performance electrodes for symmetric flexible supercapacitors. ACS Omega 2020, 5, 4657–4667. [Google Scholar] [CrossRef]
  36. Anitha, T.; Reddy, A.E.; Durga, I.K.; Rao, S.S.; Nam, H.W.; Kim, H.-J. Facile synthesis of ZnWO4@WS2 cauliflower-like structures for supercapacitors with enhanced electrochemical performance. J. Electroanal. Chem. 2019, 841, 86–93. [Google Scholar] [CrossRef]
  37. Raghavendra, K.V.G.; Sreekanth, T.; Kim, J.; Yoo, K. Novel hydrothermal synthesis of jasmine petal-like nanoflower WS2/ZnCo2O4 as efficient electrode material for high-performance supercapacitors. Mater. Lett. 2021, 285, 129133. [Google Scholar] [CrossRef]
  38. Krishnan, S.G.; Harilal, M.; Pal, B.; Misnon, I.I.; Karuppiah, C.; Yang, C.-C.; Jose, R. Improving the symmetry of asymmetric supercapacitors using battery-type positive electrodes and activated carbon negative electrodes by mass and charge balance. J. Electroanal. Chem. 2017, 805, 126–132. [Google Scholar] [CrossRef]
  39. Francis, S.; Koshy, E.P.; Mathew, B. Electroanalytical techniques: A tool for nanomaterial characterization. In Design, Fabrication, and Characterization of Multifunctional Nanomaterials; Elsevier: Amsterdam, The Netherlands, 2022; pp. 163–175. [Google Scholar]
  40. Nnamchi, P.S.; Obayi, C.S. Electrochemical characterization of nanomaterials. In Characterization of Nanomaterials; Elsevier: Amsterdam, The Netherlands, 2018; pp. 103–127. [Google Scholar]
  41. Imran, M.; Khan, M.S.; Afzal, A.M.; Iqbal, M.W.; Mumtaz, S.; Habila, M.A.; Ahmad, Z. Binary metallic sulphide-based nanocomposites with ZnO additives: A dual-functioning electrode material for energy storage and glucose sensing. Can. J. Chem. Eng. 2024, 102, 2184–2196. [Google Scholar] [CrossRef]
  42. Alsaiari, M.; Imran, M.; Afzal, A.M.; Mumtaz, S.; Algethami, J.S.; Harraz, F.A. Designing high-performance polyaniline@MoS2@AC hybrid electrode for electrochemical–based Next-generation battery-supercapacitor hybrid energy storage device and hydrogen evolution reaction. Phys. Scr. 2024, 99, 035938. [Google Scholar] [CrossRef]
  43. Hegazy, H.; Afzal, A.M.; Shaaban, E.; Iqbal, M.W.; Muhammad, S.; Alahmari, A. Synthesis of MXene and design the high-performance energy harvesting devices with multifunctional applications. Ceram. Int. 2023, 49, 1710–1719. [Google Scholar] [CrossRef]
  44. Yasmeen, A.; Afzal, A.M.; Iqbal, M.W.; Mumtaz, S.; Ouladsmane, M.; Usman, M.; Wabaidur, S.M. Cauliflower like porous structure of cobalt niobium sulfide@carbon nanotubes for electrode materials to enhance the redox activity in the battery-supercapacitor hybrid device. Appl. Nanosci. 2023, 13, 6471–6487. [Google Scholar] [CrossRef]
  45. Iqbal, M.Z.; Faisal, M.M.; Ali, S.R.; Afzal, A.M.; Karim, M.R.A.; Kamran, M.A.; Alharbi, T. Strontium phosphide-polyaniline composites for high performance supercapattery devices. Ceram. Int. 2020, 46, 10203–10214. [Google Scholar] [CrossRef]
  46. Lannelongue, P.; Bouchal, R.; Mourad, E.; Bodin, C.; Olarte, M.; Le Vot, S.; Favier, F.; Fontaine, O. “Water-in-Salt” for supercapacitors: A compromise between voltage, power density, energy density and stability. J. Electrochem. Soc. 2018, 165, A657. [Google Scholar] [CrossRef]
  47. Aftab, U.; Tahira, A.; Mazzaro, R.; Morandi, V.; Abro, M.I.; Baloch, M.M.; Yu, C.; Ibupoto, Z.H. Nickel–cobalt bimetallic sulfide NiCo2S4 nanostructures for a robust hydrogen evolution reaction in acidic media. RSC Adv. 2020, 10, 22196–22203. [Google Scholar] [CrossRef] [PubMed]
  48. Imran, M.; Muhammad, Z.; Muzafar, N.; Afzal, A.M.; Iqbal, M.W.; Mumtaz, S.; Munnaf, S.A.; Albaqami, M.D.; Ahmad, Z. Enhanced the electrochemical performance of CoMgS nanocomposite electrode with the doping of ZnO for supercapacitor-battery hybrid device and photochemical activity. J. Appl. Electrochem. 2024, 54, 1501–1515. [Google Scholar] [CrossRef]
  49. Miranda, M.; Sasaki, J.M. The limit of application of the Scherrer equation. Acta Crystallogr. Sect. A Found. Adv. 2018, 74, 54–65. [Google Scholar] [CrossRef] [PubMed]
  50. Chen, W.; Yu, X.; Zhao, Z.; Ji, S.; Feng, L. Hierarchical architecture of coupling graphene and 2D WS2 for high-performance supercapacitor. Electrochim. Acta 2019, 298, 313–320. [Google Scholar] [CrossRef]
  51. Afzal, A.M.; Awais, M.; Yasmeen, A.; Iqbal, M.W.; Mumtaz, S.; Ouladsmane, M.; Usman, M. Exploring the redox characteristics of porous ZnCoS@ rGO grown on nickel foam as a high-performance electrode for energy storage applications. RSC Adv. 2023, 13, 21236–21248. [Google Scholar] [CrossRef]
  52. Philip, A.; Kumar, A.R. Recent advancements and developments employing 2D-materials in enhancing the performance of electrochemical supercapacitors: A review. Renew. Sustain. Energy Rev. 2023, 182, 113423. [Google Scholar] [CrossRef]
  53. Abdullah, M.; Khan, S.; Jabbour, K.; Imran, M.; Ashiq, M.F.; John, P.; Manzoor, S.; Munawar, T.; Ashiq, M.N. Development of binder-free MoTe2/rGO electrode via hydrothermal route for supercapacitor application. Electrochim. Acta 2023, 466, 143020. [Google Scholar] [CrossRef]
  54. Cui, Y.; Shi, Q.; Liu, Z.; Lv, J.; Wang, C.; Xie, X.; Zhang, S. MXene/biomass/chitosan carbon aerogel (MBC) with shared cathode and anode for the construction of high-efficiency asymmetric supercapacitor. Chem. Eng. J. 2023, 472, 144701. [Google Scholar] [CrossRef]
  55. ur Rehman, A.; Batool, Z.; Ahmad, M.; Iqbal, M.W.; ul Haq, A.; Hegazy, H. Impact of ZnO on structural and electrochemical properties of silver spinel ferrites for asymmetric supercapacitors. J. Electroanal. Chem. 2023, 931, 117206. [Google Scholar] [CrossRef]
  56. Baraneedharan, P.; Reddy, B. Redox-based supercapacitor materials: A review. Mater. Res. Innov. 2024, 28, 49–69. [Google Scholar] [CrossRef]
  57. Elkholy, A.E.; Heakal, F.E.-T.; Allam, N.K. A facile electrosynthesis approach of amorphous Mn-Co-Fe ternary hydroxides as binder-free active electrode materials for high-performance supercapacitors. Electrochim. Acta 2019, 296, 59–68. [Google Scholar] [CrossRef]
  58. Prasanna, A.L.; Raghavendra, K.V.G.; Himasree, P.; Durga, I.K.; Gopi, C.V.M.; Rao, S.S.; Kim, H.-J. One-pot facile synthesis of nanorice-like structured CuS@WS2 as an advanced electroactive material for high-performance supercapacitors. SN Appl. Sci. 2020, 2, 409. [Google Scholar] [CrossRef]
  59. Hussain, S.; Rabani, I.; Vikraman, D.; Feroze, A.; Ali, M.; Seo, Y.-S.; Kim, H.-S.; Chun, S.-H.; Jung, J. One-pot synthesis of W2C/WS2 hybrid nanostructures for improved hydrogen evolution reactions and supercapacitors. Nanomaterials 2020, 10, 1597. [Google Scholar] [CrossRef] [PubMed]
  60. Wei, W.; Liu, B.; Gan, Y.; Ma, H.; Chen, D.; Qi, J.; Li, S. One-step hydrothermal synthesis of Ni3S2/MoS2 nanocomposites on rGO: Structural evolution and supercapacitor performance. Surf. Coat. Technol. 2020, 403, 126442. [Google Scholar] [CrossRef]
  61. Shrivastav, V.; Sundriyal, S.; Goel, P.; Shrivastav, V.; Tiwari, U.K.; Deep, A. ZIF-67 derived Co3S4 hollow microspheres and WS2 nanorods as a hybrid electrode material for flexible 2V solid-state supercapacitor. Electrochim. Acta 2020, 345, 136194. [Google Scholar] [CrossRef]
  62. Iqbal, M.Z.; Amjad, N.; Ali, R.; Khan, M.W.; Siddique, S.; Aftab, S.; Wabaidur, S.M. Elucidating binder-free magnetron sputtered molybdenum-tungsten-disulfide thin films for battery-supercapacitor devices. J. Alloys Compd. 2023, 942, 168929. [Google Scholar] [CrossRef]
  63. Wan, L.; Liu, J.; Li, X.; Zhang, Y.; Chen, J.; Du, C.; Xie, M. Fabrication of core-shell NiMoO4@MoS2 nanorods for high-performance asymmetric hybrid supercapacitors. Int. J. Hydrog. Energy 2020, 45, 4521–4533. [Google Scholar] [CrossRef]
  64. Iqbal, M.Z.; Alzaid, M.; Abbasi, U.; Alam, S.; Ali, R.; Afzal, A.M.; Aftab, S. Investigation of magnetron sputtered Ni@ Cu/WS2 as an electrode material for potential supercapattery devices. Int. J. Energy Res. 2022, 46, 7334–7347. [Google Scholar] [CrossRef]
  65. Lin, T.W.; Sadhasivam, T.; Wang, A.Y.; Chen, T.Y.; Lin, J.Y.; Shao, L.D. Ternary composite nanosheets with MoS2/WS2/graphene heterostructures as high-performance cathode materials for supercapacitors. ChemElectroChem 2018, 5, 1024–1031. [Google Scholar] [CrossRef]
  66. Elkatlawy, S.M.; Sakr, A.A.; Wang, J.; Elshahawy, A.M. Constructive Electroactive 2D/2D MoS2-N-rGO and 1D/2D Bi2S3-N-rGO Heterostructure for Excellent Mo-Bi Supercapattery Applications. J. Inorg. Organomet. Polym. Mater. 2023, 33, 1741–1754. [Google Scholar] [CrossRef]
  67. Imran, M.; Waris, M.H.; Khan, R.; Afzal, A.M.; Iqbal, M.W.; Mumtaz, M.A.; Ghfar, A.A.; Ali, A.; Mumtaz, S.; Hussain, Z. High-performance energy storage hybrid supercapacitor device based on NiCoS@CNT@graphene composite electrode material. Phys. Scr. 2023, 98, 115981. [Google Scholar] [CrossRef]
  68. He, G.; Qiao, M.; Li, W.; Lu, Y.; Zhao, T.; Zou, R.; Li, B.; Darr, J.A.; Hu, J.; Titirici, M.M.; et al. S, N-Co-doped graphene-nickel cobalt sulfide aerogel: Improved energy storage and electrocatalytic performance. Adv. Sci. 2017, 4, 1600214. [Google Scholar] [CrossRef]
  69. Li, J.; Liao, K.; Wang, X.; Shi, P.; Fan, J.; Xu, Q.; Min, Y. High-Performance Flexible All-Solid-State Supercapacitors Based on Ultralarge Graphene Nanosheets and Solvent-Exfoliated Tungsten Disulfide Nanoflakes. Adv. Mater. Interfaces 2017, 4, 1700419. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the synthesis of (a) WS2 nanosheets and (b) WS2@CoMgS composite.
Figure 1. Schematic diagram of the synthesis of (a) WS2 nanosheets and (b) WS2@CoMgS composite.
Energies 17 03363 g001
Figure 2. Illustration of the results of the (a) XRD of WS2 and composite WS2@CoMgS and (b) SEM (WS2@CoMgS). The marked area shows the presence of WS2 and (c) shows the BET analyses of the WS2@CoMgS.
Figure 2. Illustration of the results of the (a) XRD of WS2 and composite WS2@CoMgS and (b) SEM (WS2@CoMgS). The marked area shows the presence of WS2 and (c) shows the BET analyses of the WS2@CoMgS.
Energies 17 03363 g002
Figure 3. (a) Comparison of CV curves f or WS2, CoMgS, and WS2@CoMgS at a scan rate of 5 mV/s; (b) CV analysis of CoMgS (c) CV analysis WS2@CoMgS at a scan rate of 5–60 mV/s; (d) comparison of GCD curves for WS2, CoMgS, and WS2@CoMgS at 5 A/g current density; (e) GCD curves of CoMgS; and (f) WS2@CoMgS at 1.5 A/g current density.
Figure 3. (a) Comparison of CV curves f or WS2, CoMgS, and WS2@CoMgS at a scan rate of 5 mV/s; (b) CV analysis of CoMgS (c) CV analysis WS2@CoMgS at a scan rate of 5–60 mV/s; (d) comparison of GCD curves for WS2, CoMgS, and WS2@CoMgS at 5 A/g current density; (e) GCD curves of CoMgS; and (f) WS2@CoMgS at 1.5 A/g current density.
Energies 17 03363 g003
Figure 4. (a) WS2@CoMgS//AC hybrid device peak current vs. square root of scan rate; (b) specific capacity measured by CV at 5–60 mV/s with a 0.6 V potential measurement slot; (c) the specific capacity was determined using GCD at 1.5 to 2.3 A/g; (d) EIS curves of CoMgS and WS2@CoMgS electrodes.
Figure 4. (a) WS2@CoMgS//AC hybrid device peak current vs. square root of scan rate; (b) specific capacity measured by CV at 5–60 mV/s with a 0.6 V potential measurement slot; (c) the specific capacity was determined using GCD at 1.5 to 2.3 A/g; (d) EIS curves of CoMgS and WS2@CoMgS electrodes.
Energies 17 03363 g004
Figure 5. (a) Schematic of asymmetry device. (b) CV plots for AC and WS2@CoMgS at 5 mV/s. (c) The CV curves illustrate the WS2@CoMgS//AC supercapattery device at potential (0.0–1.6 V) with a 60 mV/sec scan rate. (d) GCD curves of WS2@CoMgS//AC supercapattery device under numerous current densities (1.3–2.3 A/g).
Figure 5. (a) Schematic of asymmetry device. (b) CV plots for AC and WS2@CoMgS at 5 mV/s. (c) The CV curves illustrate the WS2@CoMgS//AC supercapattery device at potential (0.0–1.6 V) with a 60 mV/sec scan rate. (d) GCD curves of WS2@CoMgS//AC supercapattery device under numerous current densities (1.3–2.3 A/g).
Energies 17 03363 g005
Figure 6. (a) Specific capacity vs. current density (1.3–2.3 A/g) for WS2@CoMgS//AC; (b) cshows charge–discharge (sec) vs. number. of cycle (7000 cycles); (c) after 7000 cycles, particular capacity vs. columbic efficiency; and (d) Energy density vs. power density of WS2@CoMgS//AC.
Figure 6. (a) Specific capacity vs. current density (1.3–2.3 A/g) for WS2@CoMgS//AC; (b) cshows charge–discharge (sec) vs. number. of cycle (7000 cycles); (c) after 7000 cycles, particular capacity vs. columbic efficiency; and (d) Energy density vs. power density of WS2@CoMgS//AC.
Energies 17 03363 g006
Table 1. Comparison of specific capacity and capacitance of all the samples.
Table 1. Comparison of specific capacity and capacitance of all the samples.
Samples5 mV/s (C/g)1.5 A/g (C/g)5 mV/s (F/g)1.5 A/g (F/g)
WS2318.84501.40531.40833.33
CoMgS389.64766.28649.401277.13
WS2@CoMgS549.74874.39916.241457.31
Table 2. Our findings compare with previous results.
Table 2. Our findings compare with previous results.
MATERIALSQS (F G−1)ED. (WH KG−1)PD. (W KG−1)REFERENCE
W2C@WS2101845.5500[59]
RGO/NI3S2/MOS26.45132.6399.8[60]
CO3S4/WS2412.747.3512[61]
WS2-MOS2511325100[62]
NIMNO4 @ MOS22246.747.544[63]
NI/CU/WS2116.1243.9425[64]
RGO@MOS2-WS236515373[65]
MOS2-N-RGO43833.4850[66]
NICOS@MOS2@RGO1896.6665.441267.18[21]
NICOS@CNT@GR19040.52000[67]
NI-CO-S/SNG-17.51.95[68]
GR/WS2312.423.183.2[69]
WS2@NICOS@ZNS//AC1457.3149.471212.30This 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

Imran, M.; Afzal, A.M.; Iqbal, M.W.; Mumtaz, S.; Alqarni, A.S. Superior Electrochemical Performance and Cyclic Stability of WS2@CoMgS//AC Composite on the Nickel-Foam for Asymmetric Supercapacitor Devices. Energies 2024, 17, 3363. https://doi.org/10.3390/en17143363

AMA Style

Imran M, Afzal AM, Iqbal MW, Mumtaz S, Alqarni AS. Superior Electrochemical Performance and Cyclic Stability of WS2@CoMgS//AC Composite on the Nickel-Foam for Asymmetric Supercapacitor Devices. Energies. 2024; 17(14):3363. https://doi.org/10.3390/en17143363

Chicago/Turabian Style

Imran, Muhammad, Amir Muhammad Afzal, Muhammad Waqas Iqbal, Sohail Mumtaz, and Areej S. Alqarni. 2024. "Superior Electrochemical Performance and Cyclic Stability of WS2@CoMgS//AC Composite on the Nickel-Foam for Asymmetric Supercapacitor Devices" Energies 17, no. 14: 3363. https://doi.org/10.3390/en17143363

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