Sana Ullah Asif*a,
Abdul Shakoorb,
Bushra Asgharc,
Abdul Waheedd,
Abdullah K. Alanazie,
Muhammad Ehsan Mazhar
*c,
Shahid Atiq
b,
Muhammad Yahya Haroonb,
Sadiab,
Abdul Qayyumf,
Waseem Abbasc,
Zainab Banog and
Farooq Ahmad
*b
aDepartment of Physics, Qilu Institute of Technology, Jinan 250200, Shandong, P. R. China. E-mail: sanaullahasif@gmail.com
bCentre of Excellence in Solid State Physics, University of the Punjab, Lahore, 54590, Pakistan. E-mail: ahmadfarooq1999@gmail.com
cInstitute of Physics, Bahauddin Zakariya University, Multan 60800, Pakistan. E-mail: dr.ehsan@bzu.edu.pk
dCivil Engineering Department NFC-IET, Multan, Pakistan
eDepartment of Chemistry, College of Science, Taif University, Taif, Saudi Arabia
fDepartment of Thermal Engineering and Industrial Facilities, Faculty of Mechanical Engineering, Opole University of Technology, ul. Prószkowska 76, 45-758 Opole, Poland
gDepartment of Physics, University of the Punjab, Lahore, Pakistan
First published on 12th September 2025
Energy crises have prompted researchers to develop new electrode materials for efficient energy storage, leading to the creation of effective energy storage devices. Therefore, this study introduces Ni-doped ZnSe/rGO-based materials fabricated through a hydrothermal synthesis method, which demonstrated enhanced electrical and electrochemical performance. X-ray diffraction (XRD) analysis revealed an increase in the crystallite size from 49.72 nm to 96.74 nm, accompanied by a corresponding growth in the particle size, which can be attributed to the incorporation of Ni and rGO as substituents. The electrochemical characterization of all fabricated electrodes indicated that the best-performing Zn0.90Ni0.10Se/rGO composite achieved a high specific capacitance of 1920.20 F g−1 at 5 mV s−1, significantly surpassing that of pure ZnSe (346.8 F g−1), as determined from CV measurements. Additionally, the Zn0.90Ni0.10Se/rGO electrode demonstrated excellent cycling stability (90.85% capacitance retention after 10000 cycles), a high power density of 3500 W kg−1 at a current density of 7 A g−1, and an energy density of 83.81 Wh kg−1 at a current density of 1 A g−1, with a storage capability of 1058.75 F g−1. The combined effect of Ni and rGO doping in the composites resulted in a notable reduction in series and charge transfer resistances. Under optimal conditions, it exhibited excellent electrochemical performance, as indicated by good ionic conductivity (0.037 S cm−1), the highest transference number for cations (0.90), and a rate constant of 1.42 × 10−8 cm s−1 at an exchange current density of 0.00137 A g−1, as well as a diffusion coefficient of 8.03 × 10−13 m2 s−1, suggesting enhanced ion transport characteristics. These promising attributes of Zn0.90Ni0.10Se/rGO strongly demonstrate it as an ideal electrode material for advanced energy storage applications.
The energy has been significantly affected by globalization, which involves saving extra energy and reusing it when resources are scarce. In this regard, several varieties of energy storage devices are available, including batteries, fuel cells, typical capacitors, and supercapacitors (SCs), which differ in their energy storage capabilities. These can be differentiated by their energy and power densities, which can help mitigate energy storage issues. Usually, SCs are the fundamental parts of energy storage systems. They are a key component for storing electrical energy in the microfarad range. Their anode and cathode remain the same, making them similar to batteries in that aspect. However, batteries provide high energy densities, while SCs deliver high power densities by storing charge at the electrode/electrolyte interface.5 Their high power density, rapid charging and discharging, excellent cycle stability, and environmental friendliness have garnered interest.6,7 SCs are appealing due to their longer lifespan, which surpasses that of batteries, and they deliver more power than both regular capacitors and batteries.8 They are an effective electrochemical energy storage device.9 Their historical background started in the 1950s with the discovery of the electrochemical double-layer capacitor (EDLC), attributed to General Electric engineers. While researching fuel cells in 1961, Standard Oil of Ohio made an accidental discovery. In 1971, NEC named their device the “super-capacitor”. In the meantime, Panasonic created the gold capacitor in 1978, and ELNA produced the Dyna-cap in 1987.10–12 In a nutshell, SCs are categorized into three types: hybrid, double-layer, and pseudo-capacitors. The electrodes in hybrid capacitors are made from two different materials, which increases the capacitance and extends the voltage window.5
Regarding electrode materials, metal selenides with typical two-dimensional structures such as MoSe2, CoSe, and SnSe2 are very promising anode materials for high-performance applications because of their diverse compositions, simple synthesis methods, high theoretical specific capacitance, and good conductivity.13 Transition metal selenides have been identified as potential electrode materials for supercapacitors, offering high energy densities and excellent rate performance.3 ZnSe is considered one of the most promising anode materials owing to its impressive performance in various energy storage devices.14 Ongoing advancements in semiconductors further highlight ZnSe as an important II–VI semiconductor, characterized by a bulk band gap of 2.67 eV and excellent optical transmission properties in the wavelength range from 0.5 to 22 μm,15–20 along with a significant exciton binding energy of 21 meV and a Bohr radius of 3.8 nm.21–23 It appears as a pale-yellow binary solid chemical. Therefore, it is a vital crystalline material for various applications and studies because of its quantum size effects and high surface atom percentage.18 Additionally, it is chemically inert, non-hygroscopic, highly pure, densely structured, and easy to handle, making it an excellent material.24
Various methods, such as solid-state synthesis, co-precipitation, green synthesis, and chemical reduction, have been used to develop it.13,25 Additionally, other techniques, such as melt quenching, aqueous colloidal methods, sonochemical methods, vapor-phase synthesis, and microwave irradiation, are employed to synthesize ZnSe.26 However, the hydrothermal method is considered the most effective because of its many advantages over the others, including simplicity, phase purity, uniformity, and better control over synthesis conditions.27 It features a crystal structure with an F3m space group, an FCC symmetry, and lattice parameters of a, 1/4 b, 1/4 c, and 5.668 Å. Pyramidal [ZnSe4] units are arranged in separate vertex-sharing chains, forming tunnel structures that are ideal for storing Li, Na, and K ions along the [110] direction.13
To enhance the supercapacitive performance of transition metal selenides, incorporating transition metal impurities (such as Al, Ag, Cu, Ni, and Co) as dopants is more effective than doping with rare earth ions.28 Generally, host materials with a certain atomic weight percentage of transition metal ions replace cation sites in the ZnSe lattice. Nickel is particularly suitable as a dopant because the atomic radius difference between nickel(II) and Zn2+ is only 8.3%, making nickel(II) an ideal choice for doping in Zn-based semiconductor compounds. Consequently, Ni2+ atoms can replace Zn2+ atoms in the lattice. Additionally, since Zn2+ (0.74 Å) and Ni2+ (0.69 Å) have similar valence states and radii, Ni2+ may also substitute Zn2+ in the ZnSe lattice.29 Incorporating transition metals into ZnSe matrices notably alters their electrical and electrochemical properties, leading us to select Ni/ZnSe semiconductors. As a result, Ni-doped ZnSe particles with sizes of 5 nm and 20 nm prepared at different temperatures exhibit a phase transition from cubic zinc blende to hexagonal wurtzite with increasing temperature.23,30 The cubic phase is generally considered more stable than the hexagonal phase.31
Notably, ZnSe, a metal selenide, has been studied as a potential electrode material because of its impressive performance in energy storage devices. Comparing ZnSe to ZnO and ZnS, in particular, shows its higher electrochemical activity and electrical conductivity.32 Regarding the rate performance of batteries and SCs, these materials have attracted considerable attention. As a result, developing efficient electrode materials for SCs and batteries remains a significant challenge.33–35 Furthermore, incorporating Ni into ZnSe may be an optimal choice for enhancing ion transport properties and conductivity, as ZnSe has a wider band gap that can be reduced through Ni doping, which promotes conductivity. Additionally, Ni doping can induce lattice distortions that help improve electron and ion mobility, along with enhancing pseudocapacitive characteristics. This combination is relatively better than NiO and NiSe2 owing to its optimal conductivity and stable matrix with Ni sites, which accommodate redox-active sites.
Primarily, when ZnSe is used as an anode material, it tends to pulverize and amorphize during charge and discharge cycles, leading to poor cycling stability. To address these issues, constructing a ZnSe/carbon hybrid material is considered one of the most effective ways to enhance the electrochemical performance of electrodes. ZnSe exhibits a significant synergistic effect with carbon, which greatly increases the anode's capacity for energy storage. Dispersing ZnSe nanoparticles in reduced graphene oxides to synthesize ZnSe–rGO nanocomposites as anode materials has been explored for all types of SCs.13,14 Owing to its buffer effect and the 2D conducting channel, the unique structure of the three-dimensional porous graphene network offers high conductivity, rapid ion insertion, ample active sites, short diffusion paths for ions, and structural stability. Graphene and its analogs, such as rGO and GO, can be produced. Thanks to its structure, graphene shows excellent mechanical strength, flexibility, and high charge carrier mobility.36–43 A published report has demonstrated that ZnSe–NiSe composites can deliver a specific capacity as high as 651.5 mAh g−1 at 1 A g−1 with an excellent cyclability of 98.7% after 10000 cycles, yielding an energy density of 44.4 Wh kg−1. These results suggest that ZnSe is a promising electrode material because of its relatively high capacity.44
Here, we systematically report the unique composition of ZnSe with 10% Ni and 10% rGO contents. The samples were prepared via a hydrothermal method, and ZnSe composites with 2D material (rGO), were investigated. The prepared composites were used to investigate samples for electron and ion transport properties in high-performance energy storage SC applications. For instance, the ionic conductivity and the cation transference number were separately investigated for the optimized electrode series in a three-electrode assembly, which also demonstrated enhanced charge-transfer kinetics (rate constant), favorable exchange current density, and improved diffusion dynamics.
A simple solvothermal method was used to produce rGO by reducing GO sheets. First, 45 mL of absolute ethanol was added to an aqueous GO dispersion of 2 mg mL−1. This solution was calcined at 180 °C for 3 hours and then sonicated and stirred continuously for 2 hours and 3 hours at 60 °C, respectively. Additionally, this suspension was placed in a Teflon-coated stainless-steel autoclave, which helped reduce GO sheets, and it was heated at about 180 °C for 3 hours. After cooling the autoclave to room temperature, the product was washed, dried, and obtained as a fine dark powder.
The procedure for preparing Zn0.90Ni0.10Se/rGO, was developed according to the described rationale, based on its stoichiometric composition. For example, ZnSe, which was selected as the electrode material, belonged to the chalcogenide family with a general formula AX, where A is any transition metal and X is a chalcogen such as Se or S. In this formula, doping Zn with Ni acted as a bimetallic to enhance electrical and electrochemical properties. The synthesized phase was physically mixed with rGO to reach an optimal weight percentage. The compound Zn0.90Ni0.10Se/rGO, containing 90% Ni-doped ZnSe (10% Ni and 90% Zn combined with Se in a total ratio of 1:
1) and 10% rGO, was achieved via a hydrothermal synthesis route developed based on this approach.
First, the hydrothermal method was used to synthesize ZnSe. Then, 2.18 g of the selenium powder and 1.422 g of zinc acetate were each mixed in separate 20 mL beakers containing deionized water and stirred with a glass rod for 1 minute. The zinc acetate solution and selenium powder were placed on a magnetic stirrer, and a NaOH solution was added drop-wise while stirring to reach a pH of 8. Subsequently, the mixture was transferred to Teflon cylinders, which were then placed in an autoclave and heated in an oven at 160 °C for 4 hours. After cooling to room temperature, the reaction mixture was centrifuged at 2000 rpm for 10 minutes and washed several times with deionized water, followed by thorough rinsing with absolute ethanol. The washed samples were dried under vacuum at 100 °C for 2 hours and then annealed at 400 °C for approximately 3 hours. The resulting solution was dried to obtain a fine powder. This powder was mixed with 10% nickel nitrate (Ni(NO3)2) to develop the dopant composition and stirred continuously to produce a homogeneous mixture, following the same method and conditions as those used for ZnSe synthesis. A similar process was used to prepare the composite recipes. The same procedure was applied to prepare the composites ZnSe/rGO and Zn0.90Ni0.10Se/rGO, in which rGO was included at 10%. The electrochemical analysis and material synthesis route are illustrated in Fig. 1.
A diffractometer was used to identify the crystal phases of the synthesized compounds with graphite-filtered Cu-Kα radiation (λ = 1.54 Å). Scanning electron microscopy (SEM) was conducted to examine the morphological features of ZnSe. The energy-dispersive X-ray spectroscopy (EDX) analysis of pure ZnSe confirmed the atomic and weight percentages of the elements, validating the stoichiometric composition of the samples.
![]() | ||
Fig. 2 (a) XRD patterns of ZnSe, (b) Zn0.90Ni0.10Se, (c) ZnSe/rGO, and (d) Zn0.90Ni0.10Se/rGO. (e) Crystal structure of ZnSe. (f) Crystallite size vs. different doping percentages. |
Some additional peaks are observed in the XRD patterns with the doping of cations (Ni). These peaks might indicate the formation of secondary or impurity phases.45 They suggest that excess Zn and Se oxides are present in the synthesized ZnSe powder. However, other peaks in the spectrum indicate the presence of impurities, making it difficult for ZnSe nanoparticles to form correctly and implying low purity. The peaks increase and become sharper as the temperature rises. For example, at 160 °C (a lower temperature), more unidentified peaks are observed, suggesting the poorer purity of the ZnSe at lower temperatures. This indicates that ZnSe's crystallinity significantly improves above 220 °C.15,45 The residual zinc and SeO2 contents in the powder decrease at a longer reaction time and a higher temperature.20 In the spectrum, the (111) plane's XRD peak intensity appears to be higher than that of the others. As the Ni concentration increases, the XRD peak intensity also slightly increases. The peak at 2θ = 27.82° matches the development of ZnSe along the (111) crystal planes and suggests the strong distribution of the Ni dopant in the material.47 The diffraction peaks of Ni-doped ZnSe show a slight shift to higher angles with increasing Ni concentration (10% at the Zn site compared to pure ZnSe). This shift, caused by the (111) peak moving to a higher angle, results from divalent Ni atoms occupying Zn2+ sites, leading to a decrease in the interplanar spacing.29
The difference in cation ionic radii, which causes compressional stress in the host lattice, is responsible for the peak shift at very small angles. As the temperature rises, the powders exhibit a ZnSe hexagonal phase. The increasing temperature of the growing crystallite size enhances the kinetic energy of the metal ions and their rate of dissociation. Large-sized hexagonal phases crystallize as a byproduct in Ni-doped ZnSe due to the high kinetic energy of the reacting metal ions with mobilized Se2− ions. Because of the aging process, the dopant ion (Ni2+) initially attached to the ZnSe layer's surface diffuses into the lattice, resulting in a cubic structure and lattice formation. Initially, ZnSe is rapidly nucleated by rising temperatures; once stable, ZnSe nuclei form, and they act as catalysts for the addition of new ZnSe nuclei. This accelerates crystal growth and leads to the formation of a hexagonal structure.23 Therefore, the substitution of transition metals, based on their atomic radii and infrared (IR) reflectivity, provides deeper insight into the system, allowing us to trace structural distortions caused by the doping element.48 Match software was used to collect various data, such as the peak angles, interplanar spacings, peak heights, peak areas, and FWHMs of both pure and doped ZnSe samples at different percentages. The Scherrer formula was applied to determine the crystallite size.45–50
![]() | ||
Fig. 3 (a–d) SEM images of ZnSe, Zn0.90Ni0.10Se, ZnSe/rGO, and Zn0.90Ni0.10Se/rGO. (e) EDX spectra of all samples. |
Furthermore, to improve the electrochemical performance of ZnSe and Zn0.90Ni0.10Se, two composites were developed, namely, ZnSe/rGO and Zn0.90Ni0.10Se/rGO, respectively, aiming to enhance further ion transport properties in the Ni-doped ZnSe-based material. The primary objective of creating these two composites was to assess their potential for balanced electronic and ionic conduction in energy storage applications. In this context, morphological features play a key role, as shown in Fig. 3(c) and (d). It is evident that adding rGO significantly changes the morphology of pristine ZnSe and its highest Ni-doped version. Specifically, ZnSe/rGO displays different structures, including sheets clustered with nanoparticles that resemble bulk forms. The development of porous networks and wool-like flakes notably reduces these features. During electrochemical testing, the Zn0.90Ni0.10Se/rGO shows markedly improved rate performance, thanks to its accessible conductive networks, compared to all other samples.51–55
Besides this, the porosity factor, which promotes conductive channels, is a key element in enhancing the electrochemical properties of energy storage devices. For example, in brief, the ZnSe exhibits an ice cube-like morphology, offering more redox active sites, but with poor conductivity, which was further improved by doping with Ni and adding rGO. Therefore, it can be concluded that porosity is not the only factor affecting charge carrier mobility; electrical conductivity, which allows for more efficient power transfer, also plays a crucial role, in determining, ion transport and structural stability. Electrochemical testing further confirms that the composite materials, rather than the pristine samples, deliver superior energy storage capabilities.
The data analysis of the EDX patterns validates the stoichiometric composition of the elements in all samples. Zn is found to have three elemental peaks at 1.0, 9.6, and 8.6 keV, and Se has one at 1.6 keV; the O peak is seen at 1.5 keV. Ni has two peaks in doped samples located at 0.9 and 7.5 keV. Also, the essential elements in the composite samples, such as Zn, Se, and C in ZnSe/rGO and Zn, Ni, Se, C, and O in Zn0.90Ni0.10Se/rGO, as evidenced from the EDX spectra of all samples, are displayed in Fig. 3(e). The atomic weight percentage of each constituent is listed in Table 1.
Sample | Wt% Zn | Wt% Se | Wt% Ni | Wt% C |
---|---|---|---|---|
ZnSe | 45.29 | 54.71 | 0 | 0 |
Zn0.90Ni0.10Se | 40.96 | 54.95 | 4.09 | 0 |
ZnSe/rGO | 40.77 | 49.23 | 0 | 10 |
Zn0.90Ni0.10Se/rGO | 36.86 | 49.45 | 3.68 | 10 |
Further justification is provided by the presentation of elemental mapping, which indicates the distribution of specific elements in ZnSe/rGO and Zn0.90Ni0.10Se/rGO, as shown in Fig. 4 and 5, respectively. For example, the distribution of Zn, Se, C, and O in ZnSe/rGO is represented by red, green, magenta, and orange colors, respectively. A similar pattern, including the Ni content in Zn0.90Ni0.10Se/rGO, is shown in Fig. 5 with different colors. These elemental maps confirm the presence and distribution of all essential elements in the composite samples within certain regions when analyzed using this technique.
![]() | ||
Fig. 7 CV curves at different scan rates for (a) ZnSe, (b) Zn0.90Ni0.10Se, (c) ZnSe/rGO, and (d) Zn0.90Ni0.10Se/rGO. (e) Combined CV curves at 5 mV s−1, and (f) scan rate vs. specific capacitance. |
The redox peak current and CV area of Zn0.90Ni0.10Se compared to the ZnSe electrode indicate that the addition of Ni leads to faster reaction kinetics, as evident from the integral area of the CV, which measures the specific capacitance values of the electroactive material. ZnSe and Zn0.90Ni0.10Se electrodes exhibit the maximum specific capacitance values of 346.8 F g−1 and 1172.6 F g−1, respectively, at a SR of 5 mV s−1. This indicates that nickel in selenides is more electrochemically active than zinc and has the longest GCD curve discharging time due to the synergistic effect between Ni and ZnSe, which signifies its higher energy storage capacity. The specific capacitance values of ZnSe/rGO and Zn0.90Ni0.10Se/rGO electrodes at 5 mV s−1 are 1843.7 F g−1 and 1920.5 F g−1, respectively, as determined from the CV curves. The composites with rGO show higher specific capacitance values than ZnSe and Zn0.90Ni0.10Se, respectively. The Zn0.90Ni0.10Se/rGO electrode demonstrates the highest specific capacitance among the ZnSe, ZnSe/rGO, and Zn0.90Ni0.10Se electrodes. This electrode exhibits more optimal capacitive behavior due to its high SSA and superior electrochemical activity. The combined voltammograms for all samples, recorded at the lowest SR of 5 mV s−1, are shown in Fig. 7(e). The purpose of this plot is to provide better insights into the electron transfer kinetics. Eqn (1) allows us to calculate each sample's capacitance (Csp) based on the previous discussion of the observed redox peaks.
![]() | (1) |
The integral of the CV curve corresponds to the charge stored (area under the curve), while product mv is the active mass, and SR, in this instance. Table 2 provides the Csp values for all samples, and Fig. 7(f) illustrates them. Compared to ZnSe-based electrode materials, this table shows that Zn0.90Ni0.10Se/rGO-based electrode materials possess the highest Csp value ever achieved, which is approximately 1920.20 F g−1. This reflects the superior potential of these electrode materials for hybrid capacitor technology.
Specific capacitances (F g−1) | ||||
---|---|---|---|---|
Scan rates (mV s−1) | ZnSe | Zn0.90Ni0.10Se | ZnSe/rGO | Zn0.90Ni0.10Se/rGO |
5 | 346.82 | 1172.60 | 1843.75 | 1920.20 |
10 | 121.53 | 893.17 | 623.28 | 1702.91 |
20 | 136.97 | 645.43 | 415.78 | 1085.98 |
30 | 120.90 | 517.17 | 331.72 | 1042.27 |
50 | 99.81 | 374.67 | 248.18 | 750.53 |
70 | 102.96 | 268.84 | 201.18 | 677.49 |
100 | 94.62 | 200.43 | 151.90 | 122.32 |
Capacitive and diffusive contributions using Cottrell's equations and Dunn's equations are computed by employing the equations below.
ipeak = avb or log(ipeak) = b![]() | (2) |
I(v) = k1v + k2v1/2 | (3) |
![]() | (4) |
The values of anodic and cathodic peak currents and capacitive and diffusive contributions of each sample were computed using eqn (2)–(4), and their graphical representation is illustrated in Fig. 8(a–f). These graphs suggest the mixed contribution of surface-controlled and diffusion-controlled processes.
GCD tests were conducted to evaluate the electrochemical performance of all fabricated electrodes within the potential range of 0–0.6 V at current densities of 1–7 A g−1. Fig. 9(a–d) displays the GCD profiles for each material. The redox behavior and pseudo-capacitance of the materials were identified using non-linear, plateau-like features in the GCD curves that matched the CV results. A higher accumulation of charges in both cases is indicated by a significant increase in the area under the GCD curve, which occurs because ions do not have enough time to enter the active material during high current density charging, as clearly shown in Fig. 9(a–d). Table 3 provides details about the changes in the observed specific capacitance for ZnSe/rGO electrodes. A clear positive trend is observed for the capacitance of the ZnSe/rGO electrodes with increasing rGO content and the impact of Ni doping. This improvement primarily results from the larger surface area provided by rGO, enabling better ion transport and increased exposure to active sites. Because rGO acts as a conducting scaffold, enhancing electrolyte penetration and facilitating fast redox reactions, key aspects of the pseudocapacitive behavior are enhanced. The discharging time of the GCD curves gradually increases until it reaches the maximum. As a result, the rGO–Zn0.90Ni0.10Se electrode exhibits the most significant area under the CV curve and the longest discharge time of the GCD curve, indicating superior energy storage capacity. Using eqn (5)–(7), the specific capacity was calculated at various current densities.
![]() | (5) |
![]() | (6) |
![]() | (7) |
![]() | ||
Fig. 9 (a–d) GCD curves of all samples, (e) current density vs. specific capacitance, and (f) Ragone plot. |
Sample | Current density (A g−1) | Discharge time (s) | Specific capacitance (F g−1) | Energy density (Wh kg−1) | Power density (W kg−1) |
---|---|---|---|---|---|
ZnSe | 1 | 163.99 | 446.39 | 31.61 | 500 |
2 | 73.88 | 312.78 | 22.15 | 1000 | |
3 | 9.69 | 273.70 | 19.38 | 1500 | |
5 | 44.64 | 207.74 | 14.71 | 2500 | |
7 | 19.11 | 166.07 | 11.76 | 3500 | |
Zn0.90Ni0.10Se | 1 | 226.60 | 162.96 | 13.35 | 295 |
2 | 79.37 | 247.76 | 20.30 | 1000 | |
3 | 46.52 | 59.18 | 4.85 | 1500 | |
5 | 21.88 | 385.50 | 31.59 | 2500 | |
7 | 12.64 | 251.88 | 20.64 | 3500 | |
ZnSe/rGO | 1 | 466.89 | 852.50 | 65.12 | 500 |
2 | 200.42 | 732.47 | 55.95 | 1000 | |
3 | 107.69 | 593.23 | 45.31 | 1500 | |
5 | 44.50 | 432.18 | 33.01 | 2500 | |
7 | 29.02 | 348.21 | 26.60 | 3500 | |
Zn0.90Ni0.10Se/rGO | 1 | 639.93 | 1058.75 | 83.81 | 500 |
2 | 256.75 | 855.47 | 67.72 | 1000 | |
3 | 144.79 | 728.26 | 57.65 | 1500 | |
5 | 77.10 | 561.05 | 44.41 | 2500 | |
7 | 32.28 | 471.08 | 37.29 | 3500 |
The active mass of the material is represented by m, the potential window by ΔV, the discharge time by Δt, and the current applied by I. The specific capacitance values of the Zn0.90Ni0.10Se/rGO electrode decrease steadily with increasing current density, dropping from 446.39, 162.96, 852.50, and 1058.75 F g−1 at 1 A g−1 to 166.07, 251.88, 348.21, and 471.08 at 7 A g−1, respectively, supporting the previous discussion. The trends in the capacitance, energy density, and power density of each sample are shown in Fig. 9(e) and (f). Additionally, the non-linear charge/discharge curves of the rGO–Zn0.90Ni0.10Se electrode align well with the CV redox peaks, showing 90.67% capacity retention after 10000 cycles, which indicates excellent rate performance, improved cycling stability, and enhanced electrochemical activity, as shown in Fig. 10, illustrating how specific capacitance varies with the current density in the rGO-based Ni/ZnSe composite.
The specific capacitance for each sample decreases rapidly as the current density increases, likely due to the inherent resistance of the active materials and the inevitable rise in IR losses at higher current densities.59,60 The power and energy densities of the rGO-based Ni/ZnSe composite were calculated via GCD testing using eqn (4) and (5). The highest energy densities for ZnSe, Zn0.90Ni0.10Se, ZnSe/rGO, and Zn0.90Ni0.10Se/rGO are 31.68, 13.35, 65.12, and 83.81 Wh kg−1, respectively, at a current density of 1 A g−1, while at 7 A g−1, the energy densities are 11.76, 20.64, 26.60, and 37.29 Wh kg−1, respectively. To distinguish the outcomes of this work, a comparison of different materials from a literature survey is reported in Table 4.36,44,56–60 Interpreting the literature survey results, different electrode materials show varying lifespans and performance parameters. For example, the best ZnSe composite (ZnSe/MoSe2) among all others demonstrates a cyclability of 99.96% after 5000 cycles, and ZnSe/FeSe has the highest capacitance (1419.8 W kg−1) and energy density (63.20 Wh kg−1) among all the reviewed materials reported in Table 4, but it has the lowest energy density (83.81 Wh kg−1) compared to this study. Our material exhibits the longest cycling period with an excellent stability of 90.86% after 10000 GCD cycles, and it also shows outstanding energy density compared to others.
Material | Specific capacitance (F g−1) | Current density (A g−1) | No. of cycles | Capacitance retention (%) | Energy density (Wh kg−1) | Power density (W kg−1) | Ref. |
---|---|---|---|---|---|---|---|
ZnSe/CoSe2 | 645 | 1.0 | 5k | 99.6 | 57 | 743 | 36 |
ZnSe/NiSe | 651.5 mAh g−1 | 1.0 | 10k | 98.7 | 44.4 | — | 44 |
MnSe/ZnSe | 1439.89 | 1 | — | — | 56.17 | 265 | 56 |
ZnSe/FeSe | 1419.8 | — | — | — | 63.20 | 283.20 | 57 |
ZnSe/MoSe2 | 450 | 1 | 2000 | 99.96% | 43 | 740 | 58 |
NiSe | 763.65 C g−1 | 1.5 | 3000 | 83% | 32.04 | 1112.4 | 59 |
MnSe2/CoSe2/rGO | 1138 C g−1 | 1 | 5000 | 80% | 45.80 | 853.10 | 60 |
Zn0.90Ni0.10Se/rGO | 1058.75 | 1 | 10![]() |
90.85% | 83.81 | 500 | This study |
EIS was performed in a three-electrode setup to investigate the capacitive nature and ion kinetics of electrode materials. The EIS plots of rGO-based Ni/ZnSe composites are shown in Fig. 11(a) at an applied AC signal with an amplitude of 10 mA and in the frequency range from 10−2 Hz to 105 Hz. For obtaining quantitative measurements, the fitting of Nyquist plots was performed using the equivalent circuit. With the help of similar circuits, one can study the electrical behavior of the circuit. A tiny semicircle is observed for the materials. Smaller semicircles indicate a significant improvement in conductivity, while larger semicircles indicate lower conductivity. The EIS plots of ZnSe-based nanocomposites show a semicircle with a smaller diameter. The diameter of the semicircle indicates the charge transfer resistance, which is directly correlated with conductivity and the contact developed at the electrode/electrolyte interface. The smaller the semicircle, the lower the charge transfer resistance.59
![]() | ||
Fig. 11 (a) Combined EIS spectra of all samples, (b–e) fitting spectra of all samples, and (f) trend of solution and charge transfer resistances from sample to sample. |
The composite electrodes ZnSe, Zn0.90Ni0.10Se, ZnSe/rGO, and Zn0.90Ni0.10Se/rGO exhibit the smallest semicircle diameter, indicating better charge transfer at the electrode/electrolyte interface. A similar circuit, shown in Fig. 11(b–e), illustrates how the internal resistance of the material, the binder resistance, and the ionic resistance of the electrolyte all contribute to the overall solution resistance (Rs). Charge transfer resistance (Rct) is another value that affects at the electrode/electrolyte interface. Lower Rs and Rct values indicate that rGO-based Ni/ZnSe electrodes, owing to their high SSAs and good electrochemical activity, exhibit more ideal capacitive behavior.3 The impedance values of ZnSe, ZnSe/rGO, Zn0.90Ni0.10Se, and Zn0.90Ni0.10Se/rGO electrodes are listed in Tables 5 and 6, calculated using EC Lab Biologic Lab Software, yielding a 1% error margin estimated from the chi-square value. Here, the Rct values of ZnSe and Zn0.90Ni0.10Se are 802 Ω and 435 Ω, respectively. ZnSe has a much better Rct than Zn0.90Ni0.10Se. Notably, the composite electrode Zn0.90Ni0.10Se shows the lowest charge transfer resistance among the other electrodes, which can be attributed to its superior surface qualities and electrochemical activity. The addition of two-dimensional graphene improves the electronic conductivity of the material compared to the composite electrodes ZnSe/rGO and Zn0.90Ni0.10Se/rGO, whose resistance to the intercalation and de-intercalation of electrolyte ions is lower (Rct = 23.89, 18.67 Ω), resulting in a higher rate of electrolyte ion transport to the active electrode surface. As a result, we observe a decrease in the solution and charge transfer resistances, as evidenced from the graph shown in Fig. 11(f). Consistent with the CV and GCD results, this increase enhances the number of conductive channels, thereby facilitating the movement of electrolyte ions. Therefore, graphene-based composites significantly boost the electrochemical performance of the electrode.
Sample | R1 (Ω) ± 1% | R2 (Ω) ± 1% | C1 (F) ± 1% | Q1 (Fs(a−1)) ± 1% | W1 (Ω) ± 1% |
---|---|---|---|---|---|
ZnSe | 0.26 | 802 | 0.0086 | 0.056 | 35.60 |
Zn0.90Ni0.10Se | 0.96 | 435 | 0.76 | 0.008 | 9.77 |
ZnSe/rGO | 0.71 | 23.89 | 0.45 | 0.001 | 0.68 |
Zn0.90Ni0.10Se/rGO | 1.94 | 18.67 | 1.98 | 0.307 | 0.21 |
Sample | Ionic conductivity (S cm−1) ± 1% | Transference number ± 1% | Rate constant (cm s−1) ± 1% | Exchange current density (A g−1) ± 1% | Diffusion coefficient (m2 s−1) ± 1% |
---|---|---|---|---|---|
ZnSe | 0.346 | 0.007 | 3.31 × 10−10 | 3.2 × 10−5 | 2.794 × 10−17 |
Zn0.90Ni0.10Se | 0.093 | 0.089 | 6.11 × 10−10 | 5.9 × 10−5 | 3.710 × 10−16 |
ZnSe/rGO | 0.126 | 0.510 | 1.11 × 10−8 | 0.00107 | 7.658 × 10−14 |
Zn0.90Ni0.10Se/rGO | 0.037 | 0.902 | 1.42 × 10−8 | 0.00137 | 8.030 × 10−13 |
The values of diffusion coefficients (DK+) of the fabricated electrodes with KOH as the electrolyte have been calculated using eqn (8).
![]() | (8) |
For instance, the diffusion coefficients are 2.79 × 10−17, 3.71 × 10−16, 7.65 × 10−14, and 8.03 × 10−13 m2 s−1 for ZnSe, Zn0.90Ni0.10Se, rGO–ZnSe, and rGO–Zn0.90Ni0.10Se, respectively. Zn0.90Ni0.10Se exhibits the highest ion mobility among the samples, indicating greater diffusion kinetics and improved electrochemical performance. Table 6 illustrates the effect of the rGO concentration on ion transport. A lower Warburg factor indicates faster ion mobility, while a higher value suggests increased resistance and reduced diffusion. The diffusion coefficient reaches a value as high as 8.03 × 10−13 m2 s−1 with a low Warburg value for the Zn0.90Ni0.10Se sample, indicating efficient ion transport and potential for supercapacitor applications.
![]() | (9) |
![]() | (10) |
Eqn (10) determines the cation transference number (t+), in which (Zd) is the Warburg impedance in the low-frequency region and (Rb) is the bulk electrolyte resistance. The anion transference number (t−) is determined as t+ (t− = 1 − t+). The t+ values of ZnSe, Zn0.90Ni0.10Se, rGO–ZnSe, and rGO–Zn0.90Ni0.10Se are 0.007, 0.089, 0.510, and 0.902, respectively, according to this equation. The findings unequivocally indicate that the incorporation of rGO reduces the anion transference number but increases the cation transference number. Such a shift signifies an enhancement in cation mobility within the electrolyte system. Table 6 indicates that the rGO–Zn0.90Ni0.10Se-based electrode maintains the maximum cation contribution to the total current. This enhanced anion mobility provides evidence of more effective ionic transport, likely due to structural or interfacial modifications introduced by the incorporation of rGO.
![]() | (11) |
![]() | (12) |
The Nernst equation can be generalized and written in a more extended form as eqn (13).
![]() | (13) |
In summary, the ion transport properties discussed in this study have been thoroughly compared with the existing literature to support our findings. The variation in each transport parameter's fluctuation demonstrates how these characteristics differ across materials at their optimal concentrations. Therefore, this study, with the addition of these features, is much more comprehensive than simply reporting the electrochemical analysis limited to energy and power densities from CV and GCD analysis. The related studies from the literature are listed in Table 7.61–65
Sample | Ionic conductivity (S cm−1) | Transference number | Rate constant (cm s−1) | Exchange current density (mA g−1) | Diffusion coefficient (m2 g−1) | Ref. |
---|---|---|---|---|---|---|
LaNiO3/MXene | 6.3 × 10−3 | 0.3 | — | — | 9.5 × 10−13 | 61 |
BiMnO3/CNTs | 0.978 × 10−3 | 0.31 | 3.91 × 10−7 | — | 3.45 × 10−18 | 62 |
ZnO/CNTs | 0.046 | 0.01 | 6.49 × 10−7 | — | 3.69 × 10−12 | 63 |
MoS2/Se/CNTs | 16.0 × 10−3 | 0.35 | 0.0029 | 0.74 | — | 64 |
BaCoO3/rGO | 0.128 | 0.2 | — | — | 4.51 × 10−13 | 65 |
Zn0.90Ni0.10/rGO | 0.037 | 0.90 | 1.42 × 10−8 | 0.000137 | 8.03 × 10−13 | This work |
This journal is © The Royal Society of Chemistry 2025 |