Next Article in Journal
Effect of Quercetin and Fingolimod, Alone or in Combination, on the Sphingolipid Metabolism in HepG2 Cells
Next Article in Special Issue
A First-Principles Investigation on the Structural, Optoelectronic, and Thermoelectric Properties of Pyrochlore Oxides (La2Tm2O7 (Tm = Hf, Zr)) for Energy Applications
Previous Article in Journal
The bHLH Transcription Factors in Neural Development and Therapeutic Applications for Neurodegenerative Diseases
Previous Article in Special Issue
Recent Advances in Inverted Perovskite Solar Cells: Designing and Fabrication
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Carbon-Coated ZnS-FeS2 Heterostructure as an Anode Material for Lithium-Ion Battery Applications

1
Metal-Organic Compounds Materials Research Center, Sejong University, 209, Neungdong-ro, Gwangjin-gu, Seoul 05006, Korea
2
Department of Nanotechnology and Advanced Materials Engineering, Sejong University, 209, Neungdong-ro, Gwangjin-gu, Seoul 05006, Korea
3
Department of Metallurgical and Materials Engineering, Inha Technical College, Incheon 22212, Korea
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(22), 13945; https://doi.org/10.3390/ijms232213945
Submission received: 7 October 2022 / Revised: 4 November 2022 / Accepted: 9 November 2022 / Published: 11 November 2022
(This article belongs to the Special Issue Materials for Energy Applications 2.0)

Abstract

:
The construction of carbon-coated heterostructures of bimetallic sulfide is an effective technique to improve the electrochemical activity of anode materials in lithium-ion batteries. In this work, the carbon-coated heterostructured ZnS-FeS2 is prepared by a two-step hydrothermal method. The crystallinity and nature of carbon-coating are confirmed by the investigation of XRD and Raman spectroscopy techniques. The nanoparticle morphology of ZnS and plate-like morphology of FeS2 is established by TEM images. The chemical composition of heterostructure ZnS-FeS2@C is discovered by an XPS study. The CV results have disclosed the charge storage mechanism, which depends on the capacitive and diffusion process. The BET surface area (37.95 m2g−1) and lower Rct value (137 Ω) of ZnS-FeS2@C are beneficial to attain higher lithium-ion storage performance. It delivered a discharge capacity of 821 mAh g−1 in the 500th continuous cycle @ A g−1, with a coulombic efficiency of around 100%, which is higher than the ZnS-FeS2 heterostructure (512 mAh g−1). The proposed strategy can improve the electrochemical performance and stability of lithium-ion batteries, and can be helpful in finding highly effective anode materials for energy storage devices.

1. Introduction

Rechargeable secondary batteries are currently one of the market’s most important energy storage technologies. Particularly, lithium-ion batteries (LIBs) are the most common power sources for portable electronic devices. Most importantly, electrode materials are crucial to the battery technology’s ability to store and convert electrical charge [1]. However, because the most used graphite anode has a low theoretical capacity (372 mAh g−1) and poor rate capability, developing a novel anode material with high energy/power density is a hot topic in LIB research [2]. Because of their high capacity, various anode materials such as transition metal oxides [3], metal sulfides [4], and metal/nonmetal have been widely investigated as promising candidate anodes for LIBs to meet the demand for increased energy and power density for LIBs. In recent years, transition metal sulfides (TMS), which have high theoretical capacities and are cheap, have turned out to be one of these alternatives [5]. Because of its high capacity, abundance, and low cost, ZnS is thought to be the most likely candidate to replace graphite as an anode for LIBs. Nanostructured ZnS could be used for energy storage, environmental remediation, photo electrolysis, and catalysis [6]. Poor structural stability and conductivity have limited ZnS in energy storage applications [7,8]. Much research has gone into improving the structural stability and conductivity of ZnS. Carbonaceous materials are the most stable matrices for ZnS impregnation, which improves ZnS conductivity owing to its porous architecture and structural stability. However, some factors continue to obstruct the use of ZnS. Another method for stabilizing the structure of ZnS-based electrode materials is to create a heterostructure for lithium-ion batteries [9].
Most interestingly, pyrite FeS2, an abundant mineral with low toxicity in the Earth’s crust, has a theoretical specific capacity of 894 mAh g−1 when it undergoes a four-electron electrochemical reaction between 0.01 and 3.0 V (vs. Li+/Li). Sadly, a large volume expansion occurs during the conversion of FeS2 to Fe0 and Li2S during the discharge process, resulting in a severe pulverization issue [10,11]. Consequently, a new solid electrolyte interphase (SEI) film would form on the newly generated nanoparticles’ surface, resulting in a high degree of polarization. In addition, the ground nanoparticles may escape from their original conductive networks and even detach from the anode, resulting in rapid capacity degradation. In addition, low electrical conductivity and slow ionic diffusion kinetic render the FeS2 anode’s rate capability inferior. So, FeS2 anodes need to be changed right away to make them last longer and increase their rate capability [12,13]. Numerous efforts have been made thus far to address the aforementioned issue with the FeS2 anode. In general, the strategy of combining two different metal sulfide nanoparticles with a carbon-based matrix often delivers features of good structural stability with poor volume expansion and boosted electrical conductivity. The uniformly coated carbon layer on the heterostructure significantly improved the Li-ion storage performances.
On the other hand, it was recently found that a heterostructure multi-component TMS can improve the electrochemical reactions and store more energy [14,15]. Herein, heterostructure nanoarrays have several benefits, including rich electroactive sites for redox reactions, an increased electrode–electrolyte contact area, a short electrolyte diffusion path, and attractive synergistic effects between the heterointerface [16,17,18]. Additionally, the junction of the heterosystem could generate the built-in electric charge, which definitely enhanced the electrical conductivity and ion transport, which will lead to better lithium-ion storage properties [17,19]. Herein, the carbon-coated ZnS-FeS2 heterostructure is a novel building block for the enhanced electrical conductivity and structural flexibility of long-life lithium-ion battery applications. In this work, two-step hydrothermal methods were used to fabricate the carbon-coated ZnS-FeS2 heterostructure for lithium-ion battery applications. ZnS-FeS2@C has produced an initial discharge capacity of 1481 mAh g−1 at a current density of 0.1 A g−1. After 500 cycles, it delivered the specific capacity of 821 mAh g−1 in the 500th cycle at a current density of 1 A g−1. This method involves merely fabricating the carbon-coated ZnS-FeS2 heterostructure-based bimetallic sulphides for outstanding Li-ion storage applications.

2. Results and Discussion

2.1. Scheme Diagram and Structural Analysis

The schematic diagram in Figure 1 illustrates the various steps of fabrication of carbon-coated hetero-structured ZnS-FeS2@C by a simple hydrothermal method. Firstly, the addition of sodium hydroxide solution to metal ion (Zn2+ and Fe2+) solution can form the respective metal hydroxides (Zn(OH)2 and Fe(OH)2). The thiourea react with water and ethanol and release S2− ions. Further, the released S2− ions react with Zn(OH)2 and Fe(OH)2 to form corresponding ZnS and FeS2 during the hydrothermal process. The prepared metal sulphide was carbon coated by glucose using the hydrothermal method, which is displayed in Figure 1b. The carbon-coated heterostructure of ZnS-FeS2 was annealed in an argon atmosphere, which improved the crystallinity and completely reduced the unreduced moieties of glucose in the coated-carbon network. This makes the carbon-coated electrode materials more suitable for lithium-ion battery applications. The following reactions take place during the formation of Zn(OH)2, Fe(OH)2 ZnS, and FeS2 [20,21]:
Z n 2 + + 2   ( O H )     Z n ( O H ) 2
F e 2 + + 2   ( O H )     F e ( O H ) 2
Z n ( O H ) 2 + F e ( O H ) 2 + 3   S 2     Z n S + F e S 2 + 2 H 2 O + O 2   ( )
Figure 2a displays the obtained XRD pattern of ZnS, FeS2, ZnS-FeS2, and ZnS-FeS2@C. The obtained XRD patterns have clearly exposed electrode material’s crystallinity and phase purity. The obtained XRD pattern of ZnS has diffraction peaks at 28.9, 33.4, 47.5, 56.6, 69.8, and 77.2°, which are attributed to the (0054), (1037), (1073), (1154), (0273), and (0291) planes of the ZnS rhombohedral crystal system (JCPDS NO: 01-083-1700). As shown in Figure 2a, pyrite FeS2 shows diffraction peaks at 28.7, 33.2, 37.3, 40.8, 47.4, 56.2, 59.2, 61.8, 64.3, 76.6, and 78.9°, which correspond to the (111), (200), (210), (211), (220), (311), (222), (023), (321), (331), and (420) planes with the cubic crystal system (42-1340) [10,22]. However, the heterostructure of ZnS-FeS2 and ZnS-FeS2@C contains the diffraction peaks of both zinc sulphide and iron sulfide. The carbon-coated ZnS-FeS2 has strong diffraction peaks, which exposed a crystalline nature. The average crystalline size of the electrode materials was calculated by the Scherrer formula. The sizes of ZnS, FeS2, ZnS-FeS2, and ZnS-FeS2@C were 28.5 nm, 47.92 nm, 25.74 nm, and 28.90 nm, respectively. No more impurity peaks were observed in the electrode materials, revealing the phase purity of the products. Figure 2b shows the Raman spectra of ZnS-FeS2 and ZnS-FeS2@C. As can be seen, ZnS-FeS2@C shows the two peaks at 1322 cm−1 and 1590.6 cm−1 corresponding to the D band and G bands, respectively. The intensity ratio of the D and G band is 1.03 for carbon coated heterostructure of ZnS-FeS2. This suggests that carbon coating exists in the partially graphitic and disordered carbon network. Further, it may be beneficial to the improvement in conductivity of heterostructure ZnS-FeS2 nanomaterials [8]. Moreover, the surface area and pore structure are a valuable parameter to determine the electrochemical kinetics of electrode materials. The N2 absorption/desorption quantities of carbon-coated ZnS-FeS2 heterostructure is higher than ZnS, FeS2, and ZnS-FeS2, which is shown in Figure 2c. The BET surface area value of ZnS-FeS2@C is 37.95 m2g−1, which is higher than ZnS-FeS2 (29.20 m2g−1), ZnS (4.55 m2g−1), and FeS2 (3.23 m2g−1). The surface area of heterostructure is significantly enhanced by carbon coating using glucose. Additionally, in Figure 2d, the pore size of ZnS, FeS2, ZnS-FeS2, and ZnS-FeS2@C is 1.43 nm, 1.42 nm, 0.87 nm, and 0.87 nm, respectively. The nano porous nature and high surface area of electrode materials are beneficial for the better electrochemical activities in lithium-ion batteries [8,23].
XPS measurements are carried out to understand the elements and their chemical states of electrode materials. As shown in Figure S1, the survey spectrum contained the Zn, Fe, S, and C elements in ZnS-FeS2@C. The high-resolution spectrum of Zn 2p (Figure 3a) had two predominant peaks at 1021.04 eV and 1044.2 eV, corresponding to Zn 2p3/2 and Zn 2p1/2, respectively. This represents the zinc that exists in +2 state in ZnS-FeS2@C [24,25,26]. As for the Fe 2p spectrum in Figure 3b, it has two major peaks around 710 eV and 725 eV, corresponding to FeS2 satellite peaks. The deconvoluted peaks of Fe 2p spectrum at 710.5 eV, 723.7 eV, and 713.8 eV, 726.5 eV are attributed to the Fe 2p3/2 and Fe 2p1/2 orbitals, which correspond to Fe2+/3+ states [10,12,27]. Figure 2c displays the carbon 1s spectrum, which holds the three deconvolution binding energies at 284.5 eV, 285.4 eV, and 288.8 eV, designated to C-C&C=C, C-O/C-S, and C=O/O-C=O, respectively [10,27,28]. However, the binding energies at 161.3 eV and 163.4 eV were attributed to the S 2p3/2 and S 2p1/2 orbitals with S2−state, respectively. The peak at 168.12 eV in S2p spectrum (Figure 2d) revealed the surface oxidation of electrode materials [28,29]. These results suggest that the existence of a carbon layer in the heterostructure of ZnS-FeS2 may be helpful to enhance the structural flexibility with a good conductive carbon network system. It may long establish the electrochemical activities in lithium-ion storage applications.

2.2. Morphological Analysis

The surface morphology of ZnS, FeS2, ZnS-FeS2, and ZnS-FeS2@C was investigated by SEM, with images displayed in Figure 4, Figures S2 and S3. Figure S2a–c shows the nanoparticle morphology of ZnS powder sample. The EDS mapping exposed the uniform distribution of zinc and sulphur elements in Figure S2d,e. The elemental map sum spectrum reveals the presence of zinc and sulphur elements in Figure S2f. SEM images of FeS2 display the stacked plate-like morphology with respective iron and sulphur elements in composition in Figure S3a–f. The heterostructure of ZnS-FeS2 depicts the agglomerated nanoparticle morphology in Figure S4a,b. Further clear observation of heterostructure ZnS-FeS2 morphology in HRTEM images has demonstrated FeS2 in the plate-like structure, but ZnS in the uniform nanoparticle structure in Figure S4c,d. The high-resolution images in Figure S4e at 10 nm contain the two different lattice fringes, which are associated with ZnS and FeS2 phases. The SAED pattern ZnS-FeS2 has clear diffraction spots, which confirm the crystalline nature of ZnS-FeS2. Figure S4g–i images were exposed the consistent distribution of Fe, Zn, and S elements.
Moreover, carbon-coated heterostructure ZnS-FeS2 exhibits the same agglomerated nanoparticle morphology for the prepared powder sample in Figure 4a,b. Figure 4c,d shows the HR-TEM images of ZnS-FeS2@C, which exposed that the heterostructured ZnS-FeS2 nanoparticles are coated with a thin layer carbon by the hydrothermal method. The hydrothermally carbon-coated layer is annealed to improve the crystallinity of the final composition. The high magnification image at 5 nm (Figure 4e) comprise different kinds of lattice fringes, which associated with 0.305 nm and 0.263 nm d spacing, corresponding to (0120) and (1037) planes of ZnS, and 0.314 nm d spacing, corresponding to the (111) plane of FeS2 nanoparticles. The selected area electron diffraction image in Figure 4f demonstrates a clear diffraction pattern, which further agrees with the crystalline nature of ZnS-FeS2@C. The uniform distribution C, Zn, Fe, and S elements in ZnS-FeS2@C is confirmed by elemental mapping images in Figure 4g–j. More specifically, the TEM results of ZnS-FeS2 and ZnS-FeS2@C revealed an average particle size of 22 nm and 25 nm, respectively. These results are in good agreement with the calculated average grain size of electrode materials obtained from the XRD results. However, this study reveals the presence of carbon-coated heterostructured nanoparticles, which is beneficial for a higher amount of lithium-ion intercalation/de-intercalation rate with improved structural stability.

2.3. Electrochemical Analysis

In order to strengthen the structural features of the carbon-coated ZnS-FeS2 heterostructure by galvanostatic charge/discharge studies, cyclic voltammetry study and electrochemical impedance spectroscopy were studied. The above-mentioned studies were carried out by assembled 2032 half cells. Figure 5a displays the initial galvanostatic charge/discharge cycles of ZnS nanoparticles. It delivered the first cycle discharge/charge capacity of 773/424 mAh g−1 at a current density of 0.1 A g−1, with a coulombic efficiency of 55%. The subsequent cycle curve shape was more intact with each other, which increased to a coulombic efficiency of 94% in the fifth cycle, with a specific capacity of 344 mAh g−1. Figure 5b displays the discharge/charge curve of FeS2, which delivered the specific capacity of 687/469 mAh g−1 at a current density of 0.1 A g−1. The initial coulombic efficiency of FeS2 was 68%; in the continuous cycle, it was increased to reach the maximum (96.3% in the fifth cycle). Figure 5c,d shows the charge/discharge plot of heterostructure ZnS-FeS2 and ZnS-FeS2@C at a current density of 0.1 A g−1. The charge/discharge capacities of the first cycle were 579/839 mAh g−1 and 1039/1481 mAh g−1 for heterostructure ZnS-FeS2 and ZnS-FeS2@C, respectively, with the corresponding coulombic efficiency of 69 and 70%, respectively. The capacity loss in the very first cycle corresponds to solid electrolyte interphase film formation [22,30]. The subsequent cycle plot was overlapped, which represents the good lithium-ion insertion/extraction behaviour of ZnS-FeS2@C.
The rate capability of ZnS, FeS2, ZnS-FeS2, and ZnS-FeS2@C was measured at various current densities, as shown in Figure 5e. The ten continuous cycles were performed at every current density to estimate the average specific capacity of ZnS-FeS2@C, which was 921, 700, 540, 447, and 382 mAh g−1 at current densities of 0.1, 0.2, 0.5, 1, and 2 A g−1, respectively. Again, the current density was switched to 0.1 A g−1, and it delivered the specific capacity of 811 mAh g−1. Additionally, the ZnS, FeS2, and heterostructure ZnS-FeS2 achieved a discharge capacity of 329/129.4 mAh g−1, 443/254 mAh g−1, and 650/328 mAh g−1, respectively, at a current density of 0.1/2 A g−1. Moreover, the rate capability plot expressed an improved specific capacity when the current density reaches 0.1 A g−1. The reactivation of electrode materials leads to an attained maximum specific capacity.
Further, Figure 5f displays the long-term cycle stability of the electrode materials at a current density of 1 A g−1. The ZnS nanoparticle has shown the 1st and 500th cycle discharge capacity of 241 and 68 mAh g−1, respectively. This plot conveyed that ZnS has a decreasing trend of discharge capacity, caused by the poor electrical conductivity of ZnS. The pristine plate like FeS2 shows a discharge capacity of 429 mAh g−1 in the 500th cycle. This plot has revealed the increasing specific capacity of FeS2 from the 100th cycle, achieved by a high amount of reactivation of electrode materials. The heterostructured ZnS-FeS2 delivered a specific capacity of 512 mAh g−1 in the 500th cycle. Additionally, this cyclic stability plot of ZnS-FeS2 has revealed the flatly increasing trend compared with pristine FeS2. Finally, the carbon-coated heterostructured ZnS-FeS2 nanoparticles delivered a specific capacity of 567 mAh g−1 in first cycle, which was increased to 821 mAh g−1 in 500th continuous cycle, with a coulombic efficiency of around 100%. The cyclic graph of ZnS-FeS2@C was depicted the initial cycles it was decreasing capacity which corresponds to the formation of a stable SEI film and activation electrode materials. From the 50 to 250th cycle, the specific capacity was well improved to reach a maximum; after the 250th cycle, it underwent stabilization up to the 435th cycle. After the 435th cycle, it underwent the capacity fading; finally, it reached a specific capacity of 821 mAh g−1 in the 500th cycle. The hydrothermal carbon coating of the heterostructure definitely improved the electrical conductivity and prevented the pulverization of electrode materials, leading to good structural flexibility. The reactivation of electrode materials was a commonly noted behaviour in metal-based nanomaterials and carbon-based materials [31,32]. The obtained specific capacity value was compared with reported ZnS- and FeS2-based anode materials in LIB applications, which are given in Table 1.
Further, the electrochemical activity of the electrode materials was evaluated in 2032 coin-type half cells. The CV of all of the electrode materials was conducted between 0.01 and 3.0 V versus Li+/Li at various scan rates from 0.2 to 1.2 mV s−1. As shown in Figure S5a, at the very first cathodic scan of ZnS, 0.25, 0.55, 0.67, and 1.36 V correspond to the decomposition of ZnS and formation of Li2S, LiZn alloy, and SEI film, respectively. The anodic scan of ZnS shows the peak current at 0.76 and 0.35 V, which were associated with the decomposition of Li2S, LiZn alloy, and ZnS formation [36,37,38].
Z n S + 2 L i + + 2   e     Z n + L i 2 S
Z n + L i + + e     L i Z n
The above redox reactions occurred in the continuous scans, corresponding to the multistep alloying/dealloying process of lithium-zinc alloys [39]. In Figure S6a, the CV curve of plate-like FeS2 is displayed, and the cathodic scan of the FeS2 contained peak current at 1.95 V corresponds to the decomposition of FeS2 and formation of Li2S and Fe. The oxidation peak current at 0.91 and 1.34 V corresponds to conversion of Fe to FeS2. The electrochemical reactions are described below in Equations (4) and (5) [12,22,27].
F e S 2 + 2 L i + + 2   e     Z n + L i 2 F e S 2
L i 2 F e S 2 + 2 L i + + 2   e     F e + L i 2 S  
Furthermore, the heterostructured ZnS-FeS2 in Figure S7a has a strong reduction/oxidation peak at 1.58/0.82 V, corresponding to the decomposition of ZnS and FeS2 as well as formation of Li2S and their reverse reactions. The carbon-coated ZnS-FeS2 has smothered redox peaks in Figure 6a, and the subsequent cycles at different scan rates also have an identical shape. This represents the good reversibility of ZnS-FeS2@C electrode material. Additionally, the dependence of charge storage mechanism was evaluated by power law [40]:
i = a v b
where i is the peak current, ν is the scan rate (mV s−1), and a and b are constant parameters. A b value close 1 indicates that the capacitive behaviour is dominant, while a b value close to 0.5 represents that the charge storage depends on the diffusion process. Figure S5b displays that the b valves of anodic peak 1 to 3 and cathodic peak 1 to 2 are 0.63, 0.66, 0.83, 0.68, and 0.79, respectively, representing that the charge storage mechanism majorly depends on the combined process. The b value of FeS2 in Figure S6b indicates that the charge storage mechanism corresponds to a diffusion-controlled process. The heterostructured ZnS-FeS2 has a b value of 0.72/0.87, attributed to the combined charge/discharge mechanism in Figure S7b. Here, Figure 6b displays that the b value of the anodic and cathodic peak is 0.5/0.69, which characteristics the combined charge storage mechanism. In addition, the exact contribution of the current response was predicted by the following equation:
i = k 1 v + k 2 v 0.5
where k1 and k2 are parameters and k1v and k2ν0.5 are assigned to capacitive and diffusion-controlled contribution, respectively. Figures S5c,d, S6c,d and S7c,d display the charge contributions of ZnS, FeS2, and ZnS-FeS2 at various scan rates, respectively. When increasing the scan rate from 0.2 to 1.6 V, the capacitive contributions also increase. Figure 6c reveals that the carbon-coated heterostructured ZnS-FeS2 have 57% of the diffusion process in the total capacity at a scan rate of 1.2 mV s−1. Figure 6d reveals that the capacitive behaviour increased from 17 to 51%, with a corresponding scan rate of 0.2 to 1.6 mV s−1. The combined charge storage mechanism of ZnS-FeS2@C composition has revealed excellent cycle stability, which agrees with the good cycle stability nature of the carbon-coated ZnS-FeS2 heterostructure.
To further investigate the electrochemical performance of the electrode materials at room temperature, an EIS study was conducted. As depicted in Figure 6f, the Nyquist plot was comprised of a semicircle in the high frequency region attributed to the charge transfer resistance (Rct), where the y-axis intercept in the x-axis represents the solution resistance (Rs) and tilted line at a low frequency corresponds to the Warburg impedance (σ). The fitted equivalent circuit of the Nyquist plot is shown in Figure S8 and the fitted parameters are described below. The solution resistance of the carbon-coated heterostructure is 1.12 Ω, which lower than heterostructured ZnS-FeS2 (1.2 Ω). The charge transfer resistance of ZnS-FeS2@C is 137 Ω, which is lower than that of other electrode materials. The lower charge transfer resistance of ZnS-FeS2@C is beneficial to the high amount of lithium-ion or electron transport between the anode and cathode. It leads to an excellent performance in lithium-ion battery applications. As shown in Figure 6f, the Warburg factor (σ) value of ZnS, FeS2, ZnS-FeS2, and ZnS-FeS2@C is 154.7, 251, 145.6, and 132.3, respectively. A lower Warburg factor value indicates better ionic conductivity of ZnS-FeS2@C. The lower Rs, Rct, and σ values of ZnS-FeS2@C have endorsed the obtained good rate capability and cyclic stability behaviour [41].
On the basis of the preceding analysis, the significantly enhanced electrochemical performance is primarily attributable to the amplified conductivity, enhanced e/Li+ transfer efficiency, and retarded volume expansion during the electrochemical process, which is closely related to the structural advantages of the carbon-coated heterostructured ZnS-FeS2 and can be assumed to be due to the following aspects: (1) the presence of nano-structured materials can shorten the ion diffusion path length, (2) the synergistic effect between the heterostructure interface, (3) the carbon coating can improve the electrical conductivity and structural stability of ZnS-FeS2, (4) the combined charge storage mechanism is helpful for long cyclic performance, and (5) the lower charge transfer and ionic diffusion resistance may lead to an excellent rate capability and improve the capacity of the electrode materials. The as-prepared carbon-coated ZnS-FeS2 heterostructure anode materials exhibit excellent electrochemical performance in higher lithium-ion storage applications owing to the above-mentioned factors.

3. Methods and Materials

3.1. Chemicals

Zinc nitrate hexahydrate (Zn(NO)3 6H2O), iron nitrate nonahydrate (Fe(NO)3 9H2O), sodium hydroxide (NaOH), thiourea ((NH2)2CS), ethanol (C2H5OH), sulphur powder (S) was purchase from Sigma Aldrich (Seoul, Korea) and Millipore water was used in this work. The chemical used these experiments without further purification.

3.2. Synthesis of ZnS, FeS2, and ZnS-FeS2 Heterostructure

In a typical synthesis of ZnS-FeS2, 0.744 g Zn(NO)3 6H2O and 1.01 g of Fe(NO)3 9H2O were dissolved in 40 mL of ethanol. Then, 2.0 g of NaOH in 40 mL water was added to the above metal ion solution dropwise. After the complete addition of NaOH, it was continuously stirred for an hour. To the above mixture, 0.762 g of thiourea was added and continuously stirred for an hour. After that, it was transferred into a 120 mL Teflon-lined stainless-steel autoclave. It was kept in an oven at 180 °C for 12 h. The final products were centrifuged and washed with water and ethanol to remove impurities. The hydrothermally prepared ZnS-FeS2 was annealed at 600 °C for 2 h in the Ar atmosphere. The bare ZnS and FeS2 were synthesized using the same procedure using corresponding precursors of 1.487 g Zn(NO)3 6H2O and 2.02 g of Fe(NO)3 9H2O, respectively.

3.3. Carbon Coating of ZnS-FeS2

Here, 200 mg of ZnS-FeS2 was dispersed in 40 mL water. Then, 400 mg of glucose in 40 mL water was added dropwise and continuously stirred for an hour. It was transferred to 120 mL of Teflon-lined stainless-steel autoclave and kept in an oven at 180 °C for 24 h. The final product of carbon-coated ZnS-FeS2 was collected by centrifuge process, and it was washed with ethanol and water several times. It was annealed at 600 °C for 2 h in an Ar atmosphere. Further, it was used for the analysis.

3.4. Material Characterizations

The crystal structure and phase purity of ZnS, FeS2, ZnS-FeS2, and ZnS-FeS2@C were investigated by X-ray diffraction analysis using PANalytical instrument with Cu Kα radiation. Raman spectra of ZnS-FeS2 and ZnS-FeS2@C were recorded using Renishaw Inc. (Wotton-under-Edge, UK), Raman instrument. N2 adsorption/desorption isotherm spectrum was recorded using nano POROSITY-HQ Mirae Instruments (Seoul, Korea). The chemical state of elements in the final composition was examined by X-ray photoelectron spectroscopy equipped with Al Kα radiation (Thermo Scientific Inc., Waltham, MA, USA). The surface morphology of the materials is noted by FE-SEM (SU08010) Tokyo, Japan and HR-TEM system equipped with an energy-dispersive X-ray spectroscopy system operated at an accelerating voltage of 200 kV (JEM-ARM200F). Electrodes are made up using a 70:10:20 ratio of active material, sodium carboxy methyl cellulose, and super P with 1 mL of water as solvent. The slurry was coated on the surface of copper foil, which is dried in a vacuum oven overnight. The Li-ion storage capability was investigated by an assembled 2032-type coin cell. It was fabricated in an Ar-filled glove box using Li metal and polypropylene as counter/reference electrodes and a separator with 1.2 M LiPF6 in ethylene carbonate and dimethyl carbonate with 3 wt% of vinyl carbonate as an electrolyte additive. Cyclic voltammetry was carried out with a potential window of 0.01 to 3.0 V using different scan rates 0.2 to 1.2 mVs−1. Electrochemical impedance spectroscopy was analysed using the frequency range from 0.1 kHz to 100 mHz with an applied amplitude of 5 mVs−1 in biologic instrument. Cycle stability and rate capability of (lithium-ion storage and conversion) were measured using a WonAtech battery tester (WBCS3000S WonAtech Co., Seoul, Korea).

4. Conclusions

In conclusion, a facile composition of the ZnS-FeS2@C heterostructure is prepared by the hydrothermal method for high-performance LIB anode materials. The crystalline carbon-coated ZnS-FeS2 heterostructure has a high surface area of 37.95 m2g−1, which is higher than that of the heterostructured ZnS-FeS2. The nanoparticle morphology of ZnS and the plate-like morphology of FeS2 have their own advantages of higher hetero interface as well as shorter ion diffusion paths. During the charge–discharge process, ZnS-FeS2@C has followed both capacitive and diffusion mechanisms, which leads to good rate capability behaviour. These features have definitely improved the Li-ion storage-ability, in terms of the ZnS-FeS2@C, which delivered an initial discharge capacity of 1481 mAh g−1 at a current density of 0.1 A g−1. Additionally, ZnS-FeS2@C nanoparticles delivered a specific capacity of 567 mAh g-1 in the first cycle, which is increased to 821 mAh g−1 in the 500th cycle at a current density of 1 A g−1. Overall, this work offers more insights that novel heterostructures will have broad application prospects in lithium-ion storage applications.

Supplementary Materials

The supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms232213945/s1.

Author Contributions

P.N.: conceptualization, data curation, investigation, methodology, visualization, writing—original draft; M.M.: methodology, investigation, and formal analysis; N.K. and H.-W.Y.: data curation and validation; W.-S.K.: methodology and validation; S.-J.K.: supervision, writing, and review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Research Foundation of Korea (NRF) grant funded by the ministry of Science, ICT, and Future Planning [NRF-2018K1A4A3A01064272] and [NRF-2020R1A6A1A03038540]. This research work was partly supported by Korea Basic Science Institute (National research Facilities and Equipment Centre) grant funded by the Ministry of Education (2022R1A6C101A774).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

There is no more conflict of financial interest in this research work.

References

  1. Huang, S.; Wang, M.; Jia, P.; Wang, B.; Zhang, J.; Zhao, Y. N-Graphene Motivated SnO2@SnS2 Heterostructure Quantum Dots for High Performance Lithium/Sodium Storage. Energy Storage Mater. 2019, 20, 225–233. [Google Scholar] [CrossRef]
  2. Wang, C.; Zhang, Y.; Li, Y.; Zhang, Y.; Dong, Y.; Li, D.; Zhang, J. Construction of Uniform SnS2/ZnS Heterostructure Nanosheets Embedded in Graphene for Advanced Lithium-Ion Batteries. J. Alloys Compd. 2020, 820, 153147. [Google Scholar] [CrossRef]
  3. Chen, Y.; Zong, W.; Chen, H.; Li, Z.; Pang, H.; Yuan, A.; Yang, H.; Shen, X. Cyanide-Metal Framework Derived Porous MoO3-Fe2O3 Hybrid Micro- Octahedrons as Superior Anode for Lithium-Ion Batteries. Chem. Eng. J. 2021, 426, 130347. [Google Scholar] [CrossRef]
  4. Tang, Q.; Su, H.; Cui, Y.; Baker, A.P.; Liu, Y.; Lu, J.; Song, X.; Zhang, H.; Wu, J.; Yu, H.; et al. Ternary Tin-Based Chalcogenide Nanoplates as a Promising Anode Material for Lithium-Ion Batteries. J. Power Sources 2018, 379, 182–190. [Google Scholar] [CrossRef]
  5. Hu, P.; Jia, Z.; Wang, Y.; Zhou, Q.; Liu, N.; Li, F.; Wang, J. Interface Engineering of Hierarchical MoS2/ZnS/C Heterostructures as Anode Materials for Highly Improved Lithium Storage Capability. ACS Appl. Energy Mater. 2020, 3, 7856–7864. [Google Scholar] [CrossRef]
  6. Huang, L.; Zhang, Y.; Shang, C.; Wang, X.; Zhou, G.; Ou, J.Z.; Wang, Y. ZnS Nanotubes/Carbon Cloth as a Reversible and High-Capacity Anode Material for Lithium-Ion Batteries. ChemElectroChem 2019, 6, 461–466. [Google Scholar] [CrossRef]
  7. Sung, K.W.; Koo, B.R.; Ahn, H.J. Hybrid Nanocomposites of Tunneled-Mesoporous Sulfur-Doped Carbon Nanofibers Embedded with Zinc Sulfide Nanoparticles for Ultrafast Lithium Storage Capability. J. Alloys Compd. 2021, 854, 157206. [Google Scholar] [CrossRef]
  8. Ren, H.; Wen, Z.; Wu, G.; Chen, S.; Joo, S.W.; Huang, J. Preparation of Zinc Sulfide@reduced Graphene Oxide Nanocomposites with Enhanced Energy Storage Performance. J. Phys. Chem. Solids 2019, 134, 43–51. [Google Scholar] [CrossRef]
  9. Zhang, Y.; Wang, P.; Yin, Y.; Zhang, X.; Fan, L.; Zhang, N.; Sun, K. Heterostructured SnS-ZnS@C Hollow Nanoboxes Embedded in Graphene for High Performance Lithium and Sodium Ion Batteries. Chem. Eng. J. 2019, 356, 1042–1051. [Google Scholar] [CrossRef]
  10. Hu, Z.; Cui, H.; Zhu, Y.; Lei, G.; Li, Z. Holey Reduced Graphene Oxide Nanosheets Wrapped Hollow FeS2@C Spheres as a High-Performance Anode Material for Sodium-Ion Batteries. J. Power Sources 2022, 536, 231438. [Google Scholar] [CrossRef]
  11. Zhang, Y.; Zhang, Z.; Zhang, Y.; Wang, P.; Han, P.; Li, K.; Liu, W. Preparation of a Sulfur-Doped Graphene-Wrapped FeS2 Microsphere Composite Material for Lithium-Ion Batteries. Energy Fuels 2021, 35, 20330–20338. [Google Scholar] [CrossRef]
  12. Yang, R.; Wang, C.; Li, Y.; Chen, Z.; Wei, M. Construction of FeS2@C Coated with Reduced Graphene Oxide as High-Performance Anode for Lithium-Ion Batteries. J. Electroanal. Chem. 2022, 918, 116467. [Google Scholar] [CrossRef]
  13. Huang, Y.; Zhao, H.; Bao, S.; Yin, Y.; Zhang, Y.; Lu, J. Hollow FeS2 Nanospheres Encapsulated in N/S Co-Doped Carbon Nanofibers as Electrode Material for Electrochemical Energy Storage. J. Alloys Compd. 2022, 905, 164184. [Google Scholar] [CrossRef]
  14. Zhao, Y.; Wang, W.; Chen, M.; Wang, R.; Fang, Z. The Synthesis of ZnS@MoS2 Hollow Polyhedrons for Enhanced Lithium Storage Performance. CrystEngComm 2018, 20, 7266–7274. [Google Scholar] [CrossRef]
  15. Zhu, X.; Li, J.; Ali, R.N.; Huang, M.; Liu, P.; Xiang, B. Toward a High-Performance Li-Ion Battery: Constructing a Co1−xS/ZnS@C Composite Derived from Metal-Organic Framework @3D Disordered Polystyrene Sphere Template. Mater. Des. 2018, 160, 636–641. [Google Scholar] [CrossRef]
  16. Li, B.; Wang, R.; Chen, Z.; Sun, D.; Fang, F.; Wu, R. Embedding Heterostructured MnS/Co1-xS Nanoparticles in Porous Carbon/Graphene for Superior Lithium Storage. J. Mater. Chem. A 2019, 7, 1260–1266. [Google Scholar] [CrossRef]
  17. Chen, F.; Shi, D.; Yang, M.; Jiang, H.; Shao, Y.; Wang, S.; Zhang, B.; Shen, J.; Wu, Y.; Hao, X. Novel Designed MnS-MoS2 Heterostructure for Fast and Stable Li/Na Storage:Insights into the Advanced Mechanism Attributed to Phase Engineering. Adv. Funct. Mater. 2021, 31, 2007132. [Google Scholar] [CrossRef]
  18. Man, X.; Liang, P.; Shu, H.; Zhang, L.; Wang, D.; Chao, D.; Liu, Z.; Du, X.; Wan, H.; Wang, H. Interface Synergistic Effect from Layered Metal Sulfides of MoS 2/SnS 2 van Der Waals Heterojunction with Enhanced Li-Ion Storage Performance. J. Phys. Chem. C 2018, 122, 24600–24608. [Google Scholar] [CrossRef]
  19. Wang, Z.; Jia, H.; Cai, Y.; Li, C.; Zheng, X.; Liang, H.; Qi, J.; Cao, J.; Feng, J. Highly Conductive Mn3O4/MnS Heterostructures Building Multi-Shelled Hollow Microspheres for High-Performance Supercapacitors. Chem. Eng. J. 2020, 392, 123890. [Google Scholar] [CrossRef]
  20. Venkatachalam, V.; Alsalme, A.; Alswieleh, A.; Jayavel, R. Double Hydroxide Mediated Synthesis of Nanostructured ZnCo2O4 as High Performance Electrode Material for Supercapacitor Applications. Chem. Eng. J. 2017, 321, 474–483. [Google Scholar] [CrossRef]
  21. Naveenkumar, P.; Munisamy, M.; Yesuraj, J.; Kim, K.; Kalaignan, G.P.; Kang, W.S.; Kim, S. Fabrication of Nanoneedle and Nanograss Array of Ni-Mixed CuCo2S4 @Ni-Foam as Binder-Free Electrode Materials for High-Performance Supercapacitor Applications. Energy Fuels 2022, 36, 3272–3282. [Google Scholar] [CrossRef]
  22. Zhang, Y.; Zhang, Z.; Zhu, Y.; Wang, R.; Suo, K.; Lin, G.; Zhang, N. Core-Shell FeS2@NSC Grown on Graphene for High Performance Lithium-Ion Storage. J. Electroanal. Chem. 2022, 918, 116510. [Google Scholar] [CrossRef]
  23. Feng, Y.; Zhang, Y.; Wei, Y.; Song, X.; Fu, Y.; Battaglia, V.S. A ZnS Nanocrystal/Reduced Graphene Oxide Composite Anode with Enhanced Electrochemical Performances for Lithium-Ion Batteries. Phys. Chem. Chem. Phys. 2016, 18, 30630–30642. [Google Scholar] [CrossRef] [PubMed]
  24. Mohamed, H.S.H.; Li, C.-F.; Wu, L.; Shi, W.-H.; Dong, W.-D.; Liu, J.; Hu, Z.-Y.; Chen, L.-H.; Li, Y.; Su, B.-L. Growing Ordered CuO Nanorods on 2D Cu/g-C3N4 Nanosheets as Stable Freestanding Anode for Outstanding Lithium Storage. Chem. Eng. J. 2021, 407, 126941. [Google Scholar] [CrossRef]
  25. Mao, M.; Jiang, L.; Wu, L.; Zhang, M.; Wang, T. The Structure Control of ZnS/Graphene Composites and Their Excellent Properties for Lithium-Ion Batteries. J. Mater. Chem. A 2015, 3, 13384–13389. [Google Scholar] [CrossRef]
  26. Xiao, X.; Zhang, Z.; Yang, K.; Mei, T.; Yan, D.; Wang, X. Design and Synthesize Hollow Spindle Ni-Doped Co9S8@ZnS Composites and Their Enhanced Cycle Performance. J. Alloys Compd. 2021, 853, 157118. [Google Scholar] [CrossRef]
  27. Teng, Y.; Xu, Y.; Cheng, X.; Gao, S.; Zhang, X.; Zhao, H.; Huo, L. Lonicerae Flos-Derived N, S Co-Doped Graphitized Carbon Uniformly Embedded with FeS2 Nanoparticles as Anode Materials for High Performance Lithium Ion Batteries. J. Alloys Compd. 2022, 909, 164707. [Google Scholar] [CrossRef]
  28. Chen, S.; Li, G.; Yang, M.; Xiong, J.; Akter, S.; Mi, L.; Li, Y. Nanotube Assembled Coral-like ZnS@N, S Co-Doped Carbon: A Sodium-Ion Batteries Anode Material with Outstanding Stability and Rate Performance. Appl. Surf. Sci. 2021, 535, 147748. [Google Scholar] [CrossRef]
  29. Zhang, H.; Zhao, L.; Ye, L.; Li, G. Capacity and Cycle Performance of Lithium Ion Batteries Employing CoxZn1-XS/Co9S8@N-Doped Reduced Graphene Oxide as Anode Material. Chem. Eng. J. 2021, 409, 127372. [Google Scholar] [CrossRef]
  30. Zhang, Z.; Huang, Y.; Liu, X.; Chen, C.; Xu, Z.; Liu, P. Zeolitic Imidazolate Frameworks Derived ZnS/Co3S4 Composite Nanoparticles Doping on Polyhedral Carbon Framework for Efficient Lithium/Sodium Storage Anode Materials. Carbon 2020, 157, 244–254. [Google Scholar] [CrossRef]
  31. Sun, H.; Xin, G.; Hu, T.; Yu, M.; Shao, D.; Sun, X.; Lian, J. High-Rate Lithiation-Induced Reactivation of Mesoporous Hollow Spheres for Long-Lived Lithium-Ion Batteries. Nat. Commun. 2014, 5, 4526. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Chen, Z.; Wu, R.; Wang, H.; Jiang, Y.; Jin, L.; Guo, Y.; Song, Y.; Fang, F.; Sun, D. Construction of Hybrid Hollow Architectures by In-Situ Rooting Ultrafine ZnS Nanorods within Porous Carbon Polyhedra for Enhanced Lithium Storage Properties. Chem. Eng. J. 2017, 326, 680–690. [Google Scholar] [CrossRef]
  33. Guo, C.; Wang, Q.; He, J.; Wu, C.; Xie, K.; Liu, Y.; Zhang, W.; Cheng, H.; Hu, H.; Wang, C. Rational Design of Unique ZnO/ZnS@N-C Heterostructures for High-Performance Lithium-Ion Batteries. J. Phys. Chem. Lett. 2020, 11, 905–912. [Google Scholar] [CrossRef] [PubMed]
  34. Duan, J.; Wang, Y.; Li, H.; Wei, D.; Wen, F.; Zhang, G.; Liu, P.; Li, L.; Zhang, W.; Chen, Z. Bimetal-Organic Framework-Derived Co9S8/ZnS@NC Heterostructures for Superior Lithium-Ion Storage. Chem. Asian J. 2020, 15, 1613–1620. [Google Scholar] [CrossRef] [PubMed]
  35. He, C.; Li, X.; Zheng, J.; Tang, B.; Rui, Y. A Variety of Carbon-Coated FeS2 Anodes: FeS2@CNT with Excellent Lithium-Ion Storage Performance. Colloids Surfaces A Physicochem. Eng. Asp. 2022, 637, 128226. [Google Scholar] [CrossRef]
  36. Li, J.; Fu, Y.; Shi, X.; Xu, Z.; Zhang, Z. Urchinlike ZnS Microspheres Decorated with Nitrogen-Doped Carbon: A Superior Anode Material for Lithium and Sodium Storage. Chem. A Eur. J. 2017, 23, 157–166. [Google Scholar] [CrossRef]
  37. He, L.; Liao, X.Z.; Yang, K.; He, Y.S.; Wen, W.; Ma, Z.F. Electrochemical Characteristics and Intercalation Mechanism of ZnS/C Composite as Anode Active Material for Lithium-Ion Batteries. Electrochim. Acta 2011, 56, 1213–1218. [Google Scholar] [CrossRef]
  38. Du, X.; Zhao, H.; Lu, Y.; Zhang, Z.; Kulka, A.; Świerczek, K. Synthesis of Core-Shell-like ZnS/C Nanocomposite as Improved Anode Material for Lithium Ion Batteries. Electrochim. Acta 2017, 228, 100–106. [Google Scholar] [CrossRef]
  39. Jiang, H.; Peng, H.; Guo, H.; Zeng, Y.; Li, L.; Zhang, Y.; Chen, Y.; Chen, X.; Zhang, J.; Chu, R. Interfacial Mechanical Strength Enhancement for High-Performance ZnS Thin-Film Anodes. ACS Appl. Mater. Interfaces 2020, 12, 51344–51356. [Google Scholar] [CrossRef]
  40. Tian, G.; Zhao, Z.; Sarapulova, A.; Das, C.; Zhu, L.; Liu, S.; Missiul, A.; Welter, E.; Maibach, J.; Dsoke, S. Understanding the Li-Ion Storage Mechanism in a Carbon Composited Zinc Sulfide Electrode. J. Mater. Chem. A 2019, 7, 15640–15653. [Google Scholar] [CrossRef]
  41. Ding, H.; Huang, H.C.; Zhang, X.K.; Xie, L.; Fan, J.Q.; Jiang, T.; Shi, D.; Ma, N.; Tsai, F.C. Zinc Sulfide Decorated on Nitrogen-Doped Carbon Derived from Metal-Organic Framework Composites for Highly Reversible Lithium-Ion Battery Anode. ChemElectroChem 2019, 6, 5617–5626. [Google Scholar] [CrossRef]
Figure 1. Schematic representation for the (a) synthesis of ZnS-FeS2 and (b) carbon coating of ZnS-FeS2 by the hydrothermal method.
Figure 1. Schematic representation for the (a) synthesis of ZnS-FeS2 and (b) carbon coating of ZnS-FeS2 by the hydrothermal method.
Ijms 23 13945 g001
Figure 2. (a) XRD pattern; (b) Raman spectra of ZnS-FeS2 and ZnS-FeS2@C; (c) N2 adsorption/desorption isotherm plot; and (d) pore size distributions of ZnS, FeS2, ZnS-FeS2, and ZnS-FeS2@C.
Figure 2. (a) XRD pattern; (b) Raman spectra of ZnS-FeS2 and ZnS-FeS2@C; (c) N2 adsorption/desorption isotherm plot; and (d) pore size distributions of ZnS, FeS2, ZnS-FeS2, and ZnS-FeS2@C.
Ijms 23 13945 g002
Figure 3. XPS spectra of ZnS-FeS2@C: (a) Zn2p, (b) Fe2p, (c) C1s, and (d) S2p.
Figure 3. XPS spectra of ZnS-FeS2@C: (a) Zn2p, (b) Fe2p, (c) C1s, and (d) S2p.
Ijms 23 13945 g003
Figure 4. (a,b) FE-SEM images and HRTEM images with (c) low magnification, (d,e) high magnification, (f) SAED pattern, and (gj) EDS mapping images of ZnS-FeS2@C.
Figure 4. (a,b) FE-SEM images and HRTEM images with (c) low magnification, (d,e) high magnification, (f) SAED pattern, and (gj) EDS mapping images of ZnS-FeS2@C.
Ijms 23 13945 g004
Figure 5. Charge–discharge voltage profiles of (a) ZnS, (b) FeS2, (c) ZnS-FeS2, and (d) ZnS-FeS2@C, as well as (e) Rrte capability study of all the electrode materials and (f) cyclic performance of ZnS, FeS2, ZnS-FeS2, and ZnS-FeS2@C at a current density of 1 A g−1 over 500 cycles.
Figure 5. Charge–discharge voltage profiles of (a) ZnS, (b) FeS2, (c) ZnS-FeS2, and (d) ZnS-FeS2@C, as well as (e) Rrte capability study of all the electrode materials and (f) cyclic performance of ZnS, FeS2, ZnS-FeS2, and ZnS-FeS2@C at a current density of 1 A g−1 over 500 cycles.
Ijms 23 13945 g005
Figure 6. (a) CV curves of the ZnS-FeS2@C electrode at different scan rates from 0.2 to 1.6 mV s−1. (b) Linear fitting of log (peak current) versus log (scan rate) plot of ZnS-FeS2@C. (c) Capacitive and diffusion contribution of ZnS-FeS2@C electrode @ 1.2 mV s−1. (d) Percentage of capacitive and diffusion contribution ratio of ZnS-FeS2@C electrode at different rates. (e) Nyquist plot of ZnS, FeS2, ZnS-FeS2, and ZnS-FeS2@C. (f) Straight line fitting of Z’ versus ω−1/2.
Figure 6. (a) CV curves of the ZnS-FeS2@C electrode at different scan rates from 0.2 to 1.6 mV s−1. (b) Linear fitting of log (peak current) versus log (scan rate) plot of ZnS-FeS2@C. (c) Capacitive and diffusion contribution of ZnS-FeS2@C electrode @ 1.2 mV s−1. (d) Percentage of capacitive and diffusion contribution ratio of ZnS-FeS2@C electrode at different rates. (e) Nyquist plot of ZnS, FeS2, ZnS-FeS2, and ZnS-FeS2@C. (f) Straight line fitting of Z’ versus ω−1/2.
Ijms 23 13945 g006
Table 1. Comparison of Li-ion battery performance of ZnS-FeS2@C with other reports.
Table 1. Comparison of Li-ion battery performance of ZnS-FeS2@C with other reports.
MaterialsReversible Capacity—(mAh g−1)Current Density—(A g−1)CyclesReference
ZnS-FeS2@C8211500This work
FeS2/SG400.11400 [11]
rGO@FeS2@C820.71300 [12]
FeS2@NSC/SG3922.5400 [22]
Ni doped Co9S8@ZnS7581500 [26]
FeS2@N/S-C52811000 [27]
ZnS@NSC-800571.411000 [28]
CoxZn1-xS/Co9S8@ rGO78611000 [29]
ZnO/ZnS@N-C/CNT386.61400 [33]
Co9S8/ZnS@NC411.21300 [34]
FeS2@CNT7501200 [35]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Naveenkumar, P.; Maniyazagan, M.; Kang, N.; Yang, H.-W.; Kang, W.-S.; Kim, S.-J. Carbon-Coated ZnS-FeS2 Heterostructure as an Anode Material for Lithium-Ion Battery Applications. Int. J. Mol. Sci. 2022, 23, 13945. https://doi.org/10.3390/ijms232213945

AMA Style

Naveenkumar P, Maniyazagan M, Kang N, Yang H-W, Kang W-S, Kim S-J. Carbon-Coated ZnS-FeS2 Heterostructure as an Anode Material for Lithium-Ion Battery Applications. International Journal of Molecular Sciences. 2022; 23(22):13945. https://doi.org/10.3390/ijms232213945

Chicago/Turabian Style

Naveenkumar, Perumal, Munisamy Maniyazagan, Nayoung Kang, Hyeon-Woo Yang, Woo-Seung Kang, and Sun-Jae Kim. 2022. "Carbon-Coated ZnS-FeS2 Heterostructure as an Anode Material for Lithium-Ion Battery Applications" International Journal of Molecular Sciences 23, no. 22: 13945. https://doi.org/10.3390/ijms232213945

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