Open Access Article
Hafeez Ur Rehmana,
Hamza Khana,
Zeeshan Abbasi
ab,
Lotfi Ben Taharc,
Rafaqat Ali Khana,
Amir Waseem
d and
Ahson Jabbar Shaikh
*a
aDepartment of Chemistry, COMSATS University, Islamabad – Abbottabad Campus, Abbottabad-22060, KPK, Pakistan. E-mail: ahson@cuiatd.edu.pk; ahsonjabbar@hotmail.com; Tel: +92-(0)334-377-0104
bNational Synchrotron Radiation Laboratory, University of Science and Technology of China, Hefei, Anhui-230029, China
cDepartment of Chemistry, College of Science, Northern Border University, P.O. Box 1231, Arar-91431, Saudi Arabia
dDepartment of Chemistry, Quaid-i-Azam University Islamabad, Islamabad-45320, Pakistan
First published on 8th September 2025
Zinc cobalt sulfide (ZnS–CoS) nanoparticles have emerged as promising electrode materials for supercapacitors due to their excellent electrochemical properties. However, achieving both high energy density and power density remains a challenge due to particle agglomeration and instability. In this work, dynamic light scattering-based optimization of ZnS–CoS nanoparticles is carried out by adjusting synthetic parameters including temperature, pH, reagent addition rate, and stabilizer concentration, resulting in significantly smaller particle sizes and improved stability. The optimized ZnS–CoS nanoparticle-based electrode exhibited an exceptional specific capacitance of 1156 F g−1, an energy density of 194 Wh kg−1, and a power density of 7260 W kg−1, which are significantly higher than the values reported in the literature. Electrochemical impedance spectroscopy (EIS) results confirmed lower charge transfer resistance (35.88 Ω), indicating faster ion transport and enhanced conductivity. Moreover, the optimized ZnS–CoS electrode demonstrated remarkable cycling stability, retaining 93.87% capacitance after 10
000 cycles. The charge transfer mechanism was understood by computational studies and four different machine learning models, namely, stacking regressor, TPOT, ANN, and RSM models, which were applied to verify the experimental specific capacitance of ZnS. The accuracy of performance is best for the stacking regression model, followed by ANN, TPOT, and RSM models. These results highlight the critical role of nanoparticle size optimization in enhancing the electrochemical performance and demonstrate DLS-optimized ZnS–CoS as a superior candidate for next-generation supercapacitors.
Transition metal sulfides (TMSs) have emerged as promising candidates for SC electrodes due to their low cost, high electrical conductivity, multiple oxidation states, superior redox activity, and excellent specific capacitance.10–14 Among TMSs, zinc sulfides and cobalt sulfides stand out due to their high theoretical capacitance, cost-effectiveness, and eco-friendly nature.15–17 However, pure ZnS and CoS suffer from poor specific capacitance and significant agglomeration.18,19 To address this challenge, mixed ZnS–CoS nanoparticles have been explored, leveraging the synergistic effect of Zn and Co to enhance the electrical conductivity, stability, and electrochemical performance.20 The combination of Zn and Co in a sulfide matrix enhances the electrochemical activity, conductivity, and cycling stability, making ZnS–CoS highly efficient for energy storage applications. ZnS–CoS offers a balance between high specific capacitance and good electrical conductivity while outperforming many metal oxides and carbon-based materials in energy density.
Synthesis methods are important tools for determining the final structure and characteristics of the resulting materials. To date, various methods including precipitation6,21–23 hydro/solvothermal processes,24 electrodeposition25 and self-template processes25 have been reported for the synthesis of pure and mixed TMSs. Due to the ease of handling, scalability, uniform mixing at the nano level, and control of particle size and shape, precipitation is considered as an effective method for the synthesis of nanoparticles.26–28
Although extensive work has been done on mixed metal sulfide nanoparticles, very limited work has been done for controlling the size and stability of mixed transition metal sulfide nanoparticles. The optimization of nanoparticles involves the systematic adjustment of various parameters to enhance their properties and performance for specific use. The parameters that can be altered are the concentration of constituent sulfides, pH, temperature, reagent addition rate, stabilizer concentration, etc.29,30 This results in control of size, shape, composition, and surface chemistry of the resulting nanoparticles. Furthermore, the interactions between nanoparticles and their surroundings can be analyzed to better understand the impact of their performance.31,32
In the last few decades, machine learning approaches have been broadly used in materials science to support the discovery of new materials and predict their performance with accuracy.33 Machine learning technology accelerates the development and analysis of energy materials, improves accuracy, and reduces research time.34 Key supercapacitor performance metrics including specific capacitance,35 power density, energy density,36 and remaining useful life37 can be predicted using machine learning models. Both modeling-based forecasting techniques and data-driven methodologies can be used to produce these predictions, leading to more precise performance evaluation and productive material optimization.
We believe that particle size and morphology control38 is an effective strategy for improving the electrochemical performance of TMS-based electrode materials. Over the years, significant efforts have been devoted to exploring the applications of TMSs as electrode materials. However, to the best of our knowledge, no strategies have been defined yet to further improve the specific capacitance, power density, energy density, and other parameters of already good-performing materials. In the present study, we trained stacking regressor and TPOT machine learning models and then successfully applied them to verify the specific capacitance calculated experimentally.
Although ANN and RSM machine learning models are frequently used in the synthesis and optimization of nanoparticles, our work presents a distinct novelty because specific capacitance prediction has mostly been documented for carbon-based39–41 and metal oxide-based nanoparticles.42,43 To the best of our knowledge, the specific capacitance of metal sulfides has not been predicted yet. Additionally, we predicted specific capacitance at different scan rates, which are not documented in the literature. Although machine learning models are significant tools, but they face limitations such as overfitting, high computational demand, and limited interpretability.44 In the present work, we handle these issues by using linearized data and advanced machine learning (ML) approaches to obtain accurate and robust predictions.
:
50) mixed metal sulfide nanoparticles were synthesized via the simplest co-precipitation process. A conical flask containing 10 mL of PVA solution and 5 mL of each 0.1 M zinc acetate and cobalt acetate solution was placed on a hot plate magnetic stirrer at 40 °C and stirred at 1300 rpm. Then, 20 mL of 0.2 M Na2S·3H2O solution was quickly added to this mixture, while the temperature was maintained at 40 °C with constant stirring at 1300 rpm. After 90 minutes, a black solution of mixed cobalt sulfide and zinc sulfide was produced. The prepared solution was centrifuged, washed three times with distilled water and then kept in an oven at 110 °C for further drying.
:
75), (50
:
50), and (75
:
25), using different amounts of zinc acetate and cobalt acetate, i.e., 2.5 mL
:
7.5 mL, 5 mL
:
5 mL, and 7.5 mL
:
2.5 mL, respectively. Depending on the stability, polydispersity index, particle size, and optical properties, further optimization was carried out on 50
:
50 mixed metal sulfide nanoparticles.
:
1 ratio of metal salts. The synthesis procedure was followed as mentioned above. Based on the smaller size and reasonable stability, further improvements were made at pH 7.
:
1 ratio of salts, 80 °C, and pH 7, the effect of the sulfide source (Na2S·3H2O) addition rate on mixed metal sulfide nanoparticles was investigated. Using the above procedure, sodium sulfide was added rapidly, and slow additions over time period of 5, 10, 15, and 20 minutes respectively, with the best results obtained from rapid addition of the sulfide source.
:
1 ratio of salts, pH 7, 80 °C, and quick addition of reagent. The optimal concentration of PVA was found to be 2%.
Specific capacitance was calculated from the CV curve using the following equation:
| Cp = A/(2mνdV) | (1) |
The following equation was used to calculate the specific capacitance from the GCD curve:
| Cs = Im × Δt/ΔV | (2) |
| E = Cs × ΔV/7.2 | (3) |
| P = 3600 × E/Δt | (4) |
In this study, feed-forward artificial neural networks (ANNs) that are well established for regression problems involving complex non-linear relationships between input variables and the target output were used. Operating parameters such as scan rate have a significant impact on the specific capacitance of electrochemical materials, and feed-forward ANNs are the best choice to learn these non-linear dependencies compared to conventional linear models. Two hidden layers were chosen to maintain a balance between computational simplicity and predictive accuracy. Our initial trials revealed that one hidden layer was not sufficient to fully capture the non-linear response, and that more than two layers increased the risk of overfitting. A total of 64 neurons were assigned to each hidden layer, which provides sufficient function approximation capacity while maintaining a manageable model complexity. This design choice aligns with material modeling, which is used for small-to-medium networks to avoid overparameterization when working with small experimental datasets.
The hidden layers use the ReLU activation function, which is commonly known as the default activation, is computationally efficient, and avoids vanishing gradient issues used in modern ANN applications. Since specific capacitance is a continuous-valued property and the task is a regression problem, a single linear output neuron was used. The Adam optimizer, a popular algorithm that adaptively modifies learning rates and generally ensures faster convergence than conventional stochastic gradient descent, was used to train the model. Since the mean squared error (MSE) loss function is the default option for regression models and assesses larger deviations more significantly, it was selected to accurately predict capacitance values.
Training was carried out with a batch size of 16 over 100 epochs. These values were selected after initial testing, which demonstrated that this configuration offered a favorable trade-off between convergence speed, training stability, and computational efficiency. For this relatively small dataset, smaller batch sizes give more stable weight updates, but a greater number of epochs did not significantly improve validation metrics. The scan rate was selected as the input feature because it is a physically valuable predictor and has a strong correlation with specific capacitance, as observed in experiments. To ensure that all values were on a similar scale and to enhance numerical stability during optimization, the data were standardized using scikit-learn's StandardScaler before training. By examining the trained ANN models using mean squared error and coefficient of determination (R2), the accuracy of these parameters was verified. Both metrics indicated that the network architecture and training parameters were suitable for predicting the specific capacitance of ZnS nanoparticles.
000× working at an accelerating voltage of 1 kV to 30 kV was used to study the surface morphology of the synthesized materials.
:
CoS = 75
:
25, 50
:
50, and 25
:
75) nanoparticles were synthesized and characterized by UV-vis spectroscopy, DLS, and zeta stability measurements. The UV-visible analysis was performed in the range of 200–800 nm for all (mixed) metal sulfide nanoparticles, and the graphs were normalized at 216 nm, as shown in Fig. 1a. The absorbance peaks of ZnS and CoS appear at 221 nm and 214 nm, respectively, which correspond to the literature values and confirm the formation of ZnS and CoS NPs.47,48 In the case of mixed metal sulfide (50
:
50), the characteristic peaks of both ZnS and CoS are visible at 221 nm and 214 nm, respectively, which proves the synthesis of mixed metal sulfide nanoparticles. Fig. 1b describes the zeta potential of mixed ZnS–CoS nanoparticles prepared with different concentrations of salts. It is clear that the stability of mixed metal sulfides is less than that of pure metal sulfides except for 50
:
50, which is close to the stability of pure ZnS. Among all, pure cobalt sulfide nanoparticles are most stable followed by pure ZnS. The zeta potential values can be influenced by many factors such as pH, ionic strength, concentration of different components, and specific conditions used.48 The hydrodynamic backscattered particle size of different ratios of metal sulfides is shown in Fig. 1c, where the hydrodynamic backscattered size of pure ZnS and CoS was found to be 406 nm and 245 nm, respectively. It is clear that the hydrodynamic backscattered size gradually decreases with the increase in the concentration of cobalt sulfide. The same trend was observed for the forward scattering angle except for cobalt sulfide, which shows 707 nm size, as shown in Fig. 1d. Fig. 1e expresses the PDI values that have no regular trend; however, the increasing concentration of cobalt in mixed sulfides results in a lower PDI, hence better control in size.
:
50) was chosen based on higher stability and synthesized at 20 °C, 40 °C, 60 °C, and 80 °C. Fig. S3a shows the UV-vis spectra of mixed ZnS and CoS nanoparticles prepared at these different temperatures. The graphs are normalized at 212 nm. The characteristic peaks as observed previously are also observed here; however, no significant difference is observed in the UV-vis peaks. Fig. S3b represents the zeta potential values of nanoparticles synthesized at these temperatures. It is clear that with the increase in temperature, there is a substantial decrease (negatively) in the zeta potential values, representing higher stability, hence mixed nanoparticles are more stable when synthesized at 80 °C. Fig. S3c and d depict the backscatter and forward scatter hydrodynamic size distribution of the mixed ZnS–CoS nanoparticles synthesized at these temperatures. It is observed that backscattered size distribution decreases with the enhancement in the temperature that corresponds with the literature,29 as the reaction is relatively slow at lower temperatures causing the particles to grow and agglomerate, whereas faster kinetics play an important role at higher temperatures, causing smaller and uniform size of nanoparticles.49 Backscattering is sensitive to smaller particles, while larger particles prefer to be scattered in a forward direction;29 therefore, when the size is measured at the forward angle, it is always larger and generally a reversed sequence is observed as compared to the size measured at the backscatter angle. As a result, forward scattering shows an increase in size with the increase in temperature. Overall, better control of size is observed at 80 °C. The polydispersity index of mixed ZnS–CoS nanoparticles at these temperatures is shown in Fig. S3e, where it is clearly observed that polydispersity continuously decreases as the temperature increases, with the lowest polydispersity observed at 60 °C, and after that, there is a slight increase in polydispersity at 80 °C. Hence, with the increase in the synthesis temperature of mixed metal sulfide nanoparticles, there is a shift from broad size distribution to uniform size distribution, and the best uniform size distribution is observed at 60 °C.The backscatter hydrodynamic size of mixed metal sulfides prepared at different pH values is shown in Fig. S4c. The hydrodynamic radius (Rhyd) decreases first at pH 5, and then steadily increases in size with the increase in pH till pH 11, after which again there is a decreasing trend in size. The hydrodynamic size at pH 13 is observed to be 262 nm. Therefore, a general trend could be explained, as a higher pH value causes more stability and relatively smaller size, hence potential agglomeration is prevented at higher pH values. Fig. S4d shows the forward scatter hydrodynamic size of the mixed sulfide nanoparticles. There is no regular trend of forward scatter hydrodynamic size, but with the variation in pH, a minimum size is observed at pH 5. Fig. S4e depicts the polydispersity index of mixed sulfide nanoparticles synthesized at different pH values, where we do not find any regular trend of polydispersity index with the variation in the pH values. The highest polydispersity index (0.93) is observed at acidic pH 3, while the lowest PdI value (0.33) is observed at neutral pH. It means that at neutral pH, uniform size distribution of particles occurs.
:
CoS. The UV-visible spectra displayed in Fig. S5a shows no significant difference. Fig. S5b shows the zeta potential of the synthesized mixed ZnS–CoS nanoparticles prepared at different times of addition of sulfide source. We can see that 5 minute addition has more negative values (−15.77 mV) with higher repulsion between the charged particles that prevents them from agglomeration,50 hence higher stability. Fig. S5c shows the hydrodynamic backscattered size of mixed ZnS–CoS nanoparticles synthesized at different times of addition of sulfide source. The back scattered hydrodynamic size increased with the increase in the time of addition till 10 minutes, and then a decrease in Rhyd was observed at 20 minutes of addition. The best control of size is depicted for quick addition followed by 5-minute addition. Fig. S5d shows the forward scattered hydrodynamic size. An increase in Rhyd is observed for 5 minutes, then a decrease for 10 minutes, and again increase for 20 minutes of addition. Fig. S5e shows that the polydispersity indexes of mixed metal sulfide nanoparticles synthesized at different times of addition of sulfide source are lowest at quick and 20-minute addition and highest at 5- and 10-minute addition. Overall, the size and PdI analyses show that the best reaction kinetics occur at quick addition, yielding the lowest size and PDI value.
:
50 ratio, pH 7, 80 °C, and quick addition of sodium sulfide solution. The UV-visible spectra are shown in Fig. S6a.Fig. S6b depicts the zeta potential of mixed metal sulfides synthesized with different concentrations of PVA. A steady decrease (negatively) in zeta potential is observed with the increase in the concentration of stabilizer. At 2% PVA concentration, the zeta potential is more negative, i.e. −17 mV. Hence, decreasing the zeta potential values with the increase in stabilizer concentration indicates that stability increases by increasing the stabilizer concentration. This is because stabilizers cap the nanoparticle's surface and prevent it from agglomeration.51 In this way, mixed sulfide nanoparticles are highly stabilized by the addition of a capping agent. Fig. S6c demonstrates the backscattered hydrodynamic size of mixed metal sulfides synthesized with different concentrations of PVA. The backscattered hydrodynamic size steadily decreases with the increase in the amount of PVA. For a maximum 2% concentration of stabilizer, a minimum size of 214 nm is achieved. Adding a higher concentration of stabilizer prevents nanoparticles from agglomeration and leads to a reduction in the size of nanoparticles.52 Fig. S6d demonstrates the forward scattered hydrodynamic size of the mixed metal sulfides. It is obvious that Rhyd decreases with the increase in the concentration of PVA, and with 2% PVA concentration, the size is only 185 nm. The polydispersity of mixed sulfide nanoparticles is shown in Fig. S6e. The PDI of mixed sulfide nanoparticles synthesized without any stabilizer is lowest at 0.22, whereas an increase in the percentage of stabilizers (PVA) leads to an increase in the PDI value, reaching a maximum of 0.53 at 2%.
As a result of these investigations, large-scale synthesis was performed for 50
:
50 ratios, pH 7, 80 °C, quick addition of sodium sulfide solution, and 2% concentration of PVA. The optimized nanoparticles were further characterized and utilized for electrochemical studies.
:
50) metal sulfide nanoparticles by X-ray diffraction (XRD). The 2θ values in the XRD spectra correspond to angles at which the crystal lattice deflects X-rays. Each peak indicates a distinct set of crystal planes. The compound's crystallinity and crystallite size can be determined by these peaks. The crystalline structure of ZnS NPs was determined by analyzing their diffraction pattern, as shown in Fig. 3a, which indicates phase purity. Three significant peaks are observed at 2θ = 28.08°, 47.66°, and 56.85° assigned to the 111, 220, and 311 planes, respectively. All the crystallite peaks have solid alignment with standard JCPDS card no #01-080-0020, which verifies the blende cubic phase of ZnS nanoparticles.43 Fig. 3b shows the XRD pattern of cobalt sulfide nanoparticles. Prominent peaks at 2θ = 29.91°, 36.62°, 47.32°, and 52.29° corresponding to the 100, 101, 102, and 110 lattice planes, respectively, are observed to have strong agreement with the JCPDS card no #65-3418 that validates hexagonal CoS.58 Fig. 3c depicts the XRD pattern of ZnS–CoS, with the characteristic peak at 2θ = 56.85° confirming the presence of ZnS in the optimized ZnS–CoS mixed metal sulfides, while the peaks observed at 2θ = 29.91°, 36.62°, 47.32°, and 52.29° confirm the presence of CoS in mixed metal sulfides.
000× magnification with 200 nm scale bar, which reveals its nano-morphology. Cobalt sulfide's morphological analysis is illustrated in Fig. 4b, with a rod-like morphology. Mixed ZnS–CoS (50
:
50) nanoparticles are shown in Fig. 4c with a particle-like morphology.
The specific capacitance was calculated using eqn (1) for ZnS, CoS, and optimized ZnS–CoS at different scan rates from 5 mV s−1 to 100 mV s−1 in the potential window of −0.4 to 0.6 V. Excellent results were obtained at 5 mV s−1 for each material. For ZnS, it is 315 F g−1 and for CoS, it is 199 F g−1. For mixed metal sulfides, which were optimized under different conditions, extraordinary results were obtained, as presented in Fig. 6d. Specific capacitance was obtained to be 868 F g−1 for optimized mixed metal sulfides, which is more than four times that of CoS and nearly 3 times that of ZnS. With the increase in scan rate, there is an exponential decrease in the specific capacitance due to limited time availability among the electrode and electrolyte intercalation.59
The CV curve maintained at a higher scan revealed excellent stability of the electrode materials. In all samples, the peak current increases with the increase in scan rate that is associated with the redox reaction. In Fig. 6c, the large integration area of optimized ZnS–CoS is attributed to the synergetic effect of ZnS and CoS in mixed metal sulfides. From cyclic voltammetry results, the following conclusion can be drawn:
(1) The voltammogram's peaks seem to be well defined despite some broadening, particularly at higher scan rates. This broadening may indicate quasi-reversibility since it implies a deviation from perfect reversibility.
(2) At higher scan rates (100 mV s−1), the anodic and cathodic peak separation appears to be greater than the ideal 59 mV predicted for a fully reversible process. A quasi-reversible process is characterized by increasing ΔEp with the scan rate.
(3) Peak separation increases with the increase in scan rate (from 5 mV s−1 to 100 mV s−1), which is characteristic of quasi-reversible reactions. The peak separation is constant for reversible processes regardless of the scan rate.
EIS was applied to achieve the kinetics of the charge transfer. Common impedance plots comprise three primary regions. A high-frequency semicircle whose diameter reflects the charge-transfer resistance (Rct), a low-frequency region, and an intercept on the real axis that indicates the solution resistance. Under similar conditions, we performed electrochemical impedance spectroscopy for ZnS, CoS, and optimized ZnS–CoS, and the resulting impedance plots are illustrated in Fig. 7a. Fig. 7b shows that optimized ZnS–CoS demonstrates the lowest Rct (35.88) among these three materials, as also shown in Table 1. The rapid movement of charge carriers in an electrochemical process is attributed to the synergistic effect of ZnS and CoS, which provide superior conductive support. It suggests that the maximum number of charged ions reaches the surface of the electrode to support the electrochemical redox process. Therefore, the impedance study verifies that the optimized ZnS–CoS is an outstanding material with excellent specific capacitance and superior rate performance.60
| Metal sulfides | Charge transfer resistance (Rct) (Ω) |
|---|---|
| ZnS | 74.20 |
| CoS | 102.06 |
| ZnS–CoS | 35.88 |
To assess the charge storage capacity of ZnS, CoS, and ZnS–CoS, a galvanostatic charge–discharge study was conducted at different current densities of 2, 4, 5, and 6 A g−1 at a fixed potential of −0.4 to 0.6, as depicted in Fig. 8a–c. The GCD curves depict the nonlinear behavior that demonstrates the faradaic reactions between the electrode materials and the electrolyte solution that has strong agreement with the CV results. The specific capacitance calculated from the GCD profile at different current densities is slightly higher than the specific capacitance calculated from CV. A decrease in the specific capacitance with the increase in current densities is observed due to the fact that electrolyte ions do not have enough time to migrate into the pores of materials.61 It is clear from Table 2 that DLS-optimized ZnS–CoS depicts an extraordinary specific capacitance of 1156 F g−1 at 2 A g−1, which is four times higher than that of ZnS and CoS, due to the synergetic effect of Zn and Co.
| Current densities (A g−1) | 2 | 4 | 5 | 6 |
|---|---|---|---|---|
| ZnS | 264 | 208 | 190 | 168 |
| CoS | 248 | 152 | 130 | 7.44 |
| ZnS–CoS | 1156 | 568 | 540 | 432 |
The specific capacitance of cobalt sulfide drops drastically from 130 F g−1 at 5 A g−1 to 7.44 F g−1 at 6 A g−1 (Table 2). This rapid drop is explained by slow charge-transfer kinetics and limited ion diffusion at high current densities, which significantly reduce the effective contact time between electrolyte ions and the active electrode surface.62 Cobalt sulfide depicts a pronounced decrease in specific capacitance as the current density increases, and it is well documented in the literature. For instance, Sathiskumar et al. reported a specific capacitance of 348 F g−1 at 1 A g−1. At higher current densities, this capacitance dropped significantly to 100 F g−1 at 20 A g−1, or a 71% reduction.63 By contrast, composites often retain specific capacitance even at high current density.64 Additionally, our EIS results are also consistent with the specific capacitance values listed in Table 2. The significant loss in capacitance at higher current densities can be explained by the comparatively high charge-transfer resistance of cobalt sulfide (102.66 Ω). Higher surface charge resistance indicates the sluggish charge transfer kinetics that results in poor electrochemical performance.65
Energy density and power density were also calculated using eqn (3) and (4) provided in the experimental section. As shown in Table 3, DLS-optimized ZnS–CoS exhibits an excellent power density of 7260 (W kg−1) at an energy density of 72 (Wh kg−1) at 6 A g−1, which is decreased to 2420 (W kg−1) at an energy density of 194 (Wh kg−1) at 2 A g−1, indicating a higher rate capability at a higher current density. The relation between power density and energy density is illustrated in the Ragone plot (Fig. 9a). Cycling stability is the key parameter to decide the durability and efficiency of materials. The cycling stability of optimized ZnS–CoS was determined by performing repeated charge–discharge tests at a very high current density of 20 A g−1 over 10
000 cycles, as expressed in Fig. 9b. Over 10
000 cycles, 93.87% capacitance is retained. Moreover, from start to end, no structural collapse and stability were observed resulting in high conductivity of DLS-optimized ZnS–CoS mixed metal sulfides. This can be confirmed by the GCD curves of first and last ten cycles, as shown in Fig. 9c and d, respectively.
| Current density (A g−1) | 2 | 4 | 5 | 6 |
|---|---|---|---|---|
| Specific capacitance (F g−1) | 1156 | 568 | 540 | 432 |
| Energy density (Wh kg−1) | 194 | 95 | 90 | 72 |
| Power density (W kg−1) | 2420 | 4840 | 6050 | 7260 |
To support the experimental part in qualitative manner, mathematical Dunn's model was applied. The CV curves were further investigated to explore the charge storage mechanism. For this purpose, power law equation was used to describe the relationship between the specific current and scan rate applied. The results are presented in Fig. 10.
| i = aνb | (5) |
log(i) = log a + b log ν
| (6) |
In eqn (5) and (6), i, a, and b represent the current, scan rate, and constants, respectively. Using this equation, we can calculate the factor b for which the values range from 0.5 to 1. The charge storage mechanism is completely surface controlled when the value of b is close to 1. A value of b approaching 0.5 indicates a totally diffusion-controlled process. Plotting the log of a specific current against the log of scan rate in eqn (6) provides the value of factor b calculated from the slope of the straight line. As depicted in Fig. 10a, the value of b is 0.91 that is close to 1 describing that the mechanism is surface controlled. The percentage contribution can be calculated using eqn (7):
| i(ν) = k1ν + k2ν1/2 | (7) |
In Table 4, the analysis of specific capacitance, energy density, power density, and cycling stability is presented for similar materials, as compared to our optimized ZnS–CoS mixed metal sulfide electrodes, which show a significant enhancement in energy density and power density values for our optimized material.
| Materials | Specific capacitance (F g−1) | Energy density (Wh kg−1) | Power density (W kg−1) | Cyclic stability (%) | Ref. |
|---|---|---|---|---|---|
| Zn–Co–S | 1269.1 | 45.4 | 805.0 | 91.6% | 20 |
| Multilayer dodecahedrons Zn–Co sulfide | 971.01 | 31.63 | 8.09 | 90.3% | 66 |
| WS2@NiCoS | 784.38 | 49.47 | 1212.30 | 93.23% | 67 |
| Z ZCS/GO/CN | 811 | 33.3 | 624 | 81.08% | 68 |
| Co–Mn–S | 537 | 73.52 | 10 200 |
90.03% | 69 |
| Zn–Co–S nanostrip | 830 | 19.0 | 514 | 76% | 70 |
| Co–Zn–C–S | 815 | — | — | 83.9% | 71 |
| Zn–Co bimetal sulfides | 6.22 F cm−2 | 0.124 mWh | 21 mW cm−2 | 68.4% | 68 |
| Nanostructured CoS | 1056 | 31.4 | 200 | 90% | 70 |
| ZnS nanowires | 781 | 51 | 200 | 87% | 71 |
| rGO/ZnCoS | 891 | 29 | 2215 | 99% | 68 |
| DLS-optimized ZnS–CoS | 1156 | 194 | 7260 | 93.87% | Our work |
Zinc and cobalt mixed metal sulfides have porous structures that give large gaps and open channels, which increase the ion diffusion and charge transport and act as promising materials for supercapacitors.72 Hydroxide ions from KOH interact with mixed metal sulfides that promote the reversible faradaic reaction.73 The synergistic effect of zinc, cobalt, and sulfur increases the electrical conductivity, which leads to a fast and reversible charge–discharge cycle.20 The synergistic effect also provides an enhanced surface area,74 resulting in better charge storage. All these features make our optimized mixed metal sulfides promising materials for supercapacitor applications. The proposed mechanism for charge transfer at electrodes is shown in Scheme 1.
| Material | EHOMO (eV) | ELUMO (eV) | EL–H gap (eV) |
|---|---|---|---|
| ZnS | −6.72039 | −4.20143 | 2.51896 |
| CoS | −11.6110 | −8.85866 | 2.75234 |
| ZnS–CoS | −10.8105 | −8.28042 | 2.53008 |
The reduced band gap suggests better electron mobility for ZnS–CoS and correlates with the improved electrochemical performance and enhanced charge storage capability as observed in the experimental results. Specifically, the cyclic voltammetry (CV) analysis revealed that the specific capacitance values were 264 F g−1 for ZnS, 248 F g−1 for CoS, and significantly higher 1156 F g−1 for the ZnS–CoS mixed sulfide, measured at a scan rate of 2 mV s−1. The results also complement electrochemical impedance spectroscopy (EIS) analysis, which demonstrated a lower charge transfer resistance for the ZnS–CoS mixed sulfide, confirming its superior electrical conductivity.
The density of states (DOS) analysis is shown in Fig. 12a–c, indicating that electron transfer between the valence and conduction bands is crucial for determining their conductivity. The Fermi level lies between these two bands, affecting the overall electron transfer mechanism. The DOS analysis provides insights into energy level formation, particularly the lowest unoccupied molecular orbital (LUMO) values. For example, in Fig. 12, the LUMO of ZnS is observed at −4.20 eV and for CoS, it is observed at −8.85 eV. When mixed sulfides are formed, i.e., ZnS–CoS, the LUMO value is observed at −8.28 eV. These findings, also supported by computational and electrochemical analyses, illustrate that the synergistic effects are observed after mixing, on the electronic structure and capacitive performance. Hence, ZnS–CoS achieves better electrochemical performance due to its improved conductivity, as predicted by computational calculations.
![]() | ||
| Fig. 13 Comparison between the experimental results and the stacking regressor model predictions at different scan rates. | ||
Furthermore, it is clear from the above discussion that the artificial neural network (ANN) performed better in terms of prediction than the RSM regression model. The accuracy of the ANN model is attributed to its flexibility in capturing nonlinear behavior,75,76 which is well documented in the literature. For instance, Khoshraftar et al. used both RSM and an ANN to model CO2 adsorption on Fe–Ni/AC-modified carbon nanotubes. The ANN model achieved R2 = 0.9990 (MSE = 4.247 × 10−4), whereas the RSM model had predicted R2 = 0.8899.77 In another study, Luan et al. optimized PVB synthesis using both RSM and ANN. Their results show RSM: R2 = 0.9865, RMSE = 0.4754; versus ANN: R2 = 0.9962, RMSE = 0.2551.78 In short, ANN models frequently achieve superior predictive accuracy and robustness than RSM because they can fit the interacting nonlinear behavior more flexibly.
Although extensive work has been reported on mixed transition metal sulfide nanoparticles, very limited attention has been paid to controlling their particle size and stability. Since parameters such as precursor concentration, pH, temperature, reagent addition rate, and stabilizer concentration28,29 directly influence the size, shape, and surface chemistry, their optimization is essential for tuning the electrochemical performance. Previous studies have suggested that the particle size and morphology control36 can significantly enhance the properties of TMS-based electrodes; however, systematic strategies for improving specific capacitance, power density, and energy density remain lacking. In this work, we addressed this gap by optimizing key synthesis parameters to achieve controlled nanoparticle characteristics. The improved size and morphology directly contributed to enhanced electrochemical performance, demonstrating the importance of controlled synthesis in advancing TMS-based energy storage materials. For instance, Muhammad Sufyan Javed et al. reported a specific capacitance of 110 F g−1 for ZnS@FeSe2 nanostructures,79 A. Shah et al. achieved 95 F g−1 for ZnS/CuSe2//AC,80 while Hongxia Wang et al. reported 200 F g−1 for Cu5Sn2S7/ZnS.81 These earlier studies did not involve the optimization of nanoparticle characteristics. In contrast, our optimized nanoparticles achieved a much higher capacitance of 1156 F g−1, along with improved power density and energy density (Table 4), demonstrating that careful optimization of synthesis parameters is a highly effective strategy for advancing the performance of TMS-based energy storage materials.
In addition, our study introduces novelty by predicting the specific capacitance of ZnS by using machine learning models, for the first time, at different scan rates, which has not been reported before. Unlike conventional ML approaches that face issues like overfitting and poor interpretability, we applied linearized data with advanced ML to achieve more robust and reliable predictions. Furthermore, DFT calculations revealed a reduced band gap in Zn–CO–S, which is directly correlated with enhanced electrochemical properties, thereby reinforcing the experimental outcomes.
000 continuous charge–discharge cycles. This highlights its superior stability and durability, making it a highly promising material for long-term energy storage applications. To further understand the charge storage mechanism, Dunn's power law model was applied, revealing that the ZnS–CoS electrode exhibited a dominant surface-controlled capacitive contribution of 52.30% at a scan rate of 5 mV s−1, further supporting its excellent charge storage capability. Computational calculations, including density of states (DOS) and HOMO–LUMO energy gap analysis, confirmed the synergistic effect of Zn and Co, resulting in enhanced electronic conductivity and improved charge transport properties. Stacking regressor, TPOT, ANN, and RSM models were applied to verify the experimental specific capacitance of ZnS. The performance metrics of these models follow the ranking of stacking regressor > ANN > TPOT > RSM. This study demonstrates how well the stacking regression model can be used to improve the electrical and electrochemical behavior.
Overall, this study demonstrated that precise nanoparticle size and morphology optimization using DLS play a crucial role in enhancing the electrochemical performance of ZnS–CoS electrodes. The optimized ZnS–CoS nanoparticles not only surpass many previously reported transition metal sulfide-based electrodes but also establish a new benchmark for next-generation supercapacitor materials. The combination of high capacitance, superior energy and power density, excellent stability, and favorable charge storage kinetics makes ZnS–CoS a highly promising candidate for high-performance energy storage devices. Future research can explore further material modifications, hybrid structures, and scalable synthesis approaches to enable the widespread adoption of ZnS–CoS for commercial supercapacitor applications.
| This journal is © The Royal Society of Chemistry 2025 |