Open Access Article
Catherine Sekyerebea
Diko†
a,
Haodong
Shi†
b,
Wang
Lei
a,
Zichen
Zhu
c,
Yining
Liu
c,
Maurice
Abitonze
a,
Wendolina Martina
Micha Obono
d,
Yimin
Zhu
*a,
Yan
Yang
*c,
Zhongshuai
Wu
b and
Jian
Liu
*e
aEnvironmental Science and Engineering, Dalian Maritime University, Environmental Science & Engineering College, 116026, China
bState Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, 457 Zhongshan Road, Dalian, Liaoning, China
cResearch Department 11, SINOPEC (Dalian) Research Institute of Petroleum and Petrochemicals Co., Ltd, No. 96, Nankai Street, Lushunkou District, Dalian, 116045, Liaoning, China
dMaterials Science and Engineering College, Dalian University of Technology, 116024, China
eCollege of Chemistry and Chemical Engineering, Inner Mongolia University, Hohhot, Inner Mongolia, People's Republic of China
First published on 12th September 2025
Lithium–sulfur (Li–S) batteries are promising candidates for future energy storage systems because of their abundant theoretical capacity and low cost. However, challenges such as polysulfide shuttle effects and poor conductivity hinder their practical use. Yolk–shell structured nanocomposites offer a promising avenue for addressing the challenges in Li–S batteries. Herein, one-pot hydrothermal synthesis of yolk–shell SnS2@MoS2@C nanospheres is reported, where the inclusion of the tin precursor plays a pivotal role in tuning these unique nanostructures. The resulting architecture provides enlarged interlayer spacing, internal voids, and robust stability, facilitating efficient ion transport and volume buffering. Electrochemical evaluations reveal a high initial capacity of 1445 mA h g−1 at 0.1C, with excellent rate-performance, retaining 802 mA h g−1 at 3C. Remarkably, at 1C, the capacity increases from 1044.8 to 1114.6 mA h g−1 after 600 cycles. These results highlight the structural and functional advantages of SnS2-driven yolk–shell architectures for next-generation Li–S cathodes.
To address these limitations, the development of advanced electrode materials with tunable electrochemical properties is essential. Transition metal sulfides (TMSs) serve as promising electrode materials for energy storage applications. This is due to their high theoretical capacities, layered structures, and relatively low production costs.12–14 However, the practical use of TMS faces limitations due to its low electronic and ionic conductivities and significant volume expansion during cycling. These issues cause electrode pulverization, detachment of active materials, and, eventually, rapid capacity fading with poor rate performance during repeated charge–discharge cycles.14 To overcome these limitations, one approach involves constructing TMS composites where conductive carbon-based materials are adopted.15,16 Another strategy focuses on engineering the shape and morphology of TMS materials at the nanoscale (nanoparticles, nanosheets, nanorods, nanospheres, etc.).17 By harnessing these two approaches, TMS carbon-based composites can offer significant structural advantages and enhanced functionalities for advanced energy storage applications.
Recently, yolk–shell architectures have served as active materials in battery assembly, boosting energy density, while the void space relieves mechanical stress from volume changes, preventing structural damage.18–20 When applied to Li–S batteries, this design can accommodate and confine soluble polysulfides within the shell. Also, yolk–shell hosts have been shown to accelerate polysulfide conversion by improving sulfur utilization and extending cycle life.21 Despite these advances, the fabrication of yolk–shell TMS-carbon composites remains underexplored for Li–S applications.
In this study, a novel facile one-pot hydrothermal synthesis of yolk–shell SnS2/MoS2/Carbon (SnS2@MoS2@C) nanospheres was designed for Li–S applications. The incorporation of a tin (Sn) precursor was found to play a crucial role in regulating the nanosphere structure, promoting the formation of a porous yolk–shell with expanded interlayer architectures. Notably, in the absence of the Sn precursor, hollow-shell MoS2@C nanospheres were obtained, whereas omitting both Sn and carbon precursors led to the formation of amorphous core–shell MoS2 structures. Comprehensive characterization revealed that yolk–shell SnS2@MoS2@C exhibited uniform morphology, enhanced surface area, and expanded interlayer spacing for improved ion and electron transport. This design significantly enhances electrochemical performance by providing an abundance of active sites for sulfur redox reactions. The yolk–shell SnS2@MoS2@C nanospheres demonstrated superior lithium–sulfur redox kinetics, achieving a high discharge capacity of 1445 mA h g−1 at 0.1C and maintaining 802 mA h g−1 at 3C. Furthermore, the yolk–shell exhibited remarkable cycling stability, achieving a capacity of 1114.6 mA h g−1 at 1C after 600 cycles and a capacity increase rate of 0.01% per cycle. Compared to hollow-shell MoS2@C and amorphous MoS2 nanospheres, the yolk–shell nanostructure also suppressed the polysulfide shuttle more effectively and promoted stable cycling. These results show the efficacy of SnS2-induced yolk–shell engineering as a promising method for designing high-performance cathode materials in lithium–sulfur battery systems.
For carbonization, the SnS2@MoS2@RF powder was heated in a tube furnace under a N2 atmosphere. This involved an initial heating rate of 1 °C min−1 up to 220 °C with a dwell time of 60 minutes, followed by further heating at 1 °C min−1 to 400 °C maintained for 120 minutes to yield the final SnS2@MoS2@C yolk–shell nanospheres. This temperature was selected to ensure carbon formation while avoiding MoS2 and SnS2 decomposition.
:
1) in a molar ratio of 5
:
1 to create a 10 mM Li2S6 solution. The mixture was heated and stirred at 65 °C until complete dissolution. Afterward, 10 mg of the nanomaterials were added to 5 mL of the Li2S6 solution and homogenized for 4 hours. The entire process was conducted within a glove box filled with argon. The color changes were observed by ultraviolet (UV) spectroscopy.
:
20
:
10, and the mixture was homogenized in N-methyl-2-pyrrolidone (NMP) solvent to form a viscous cathode slurry. This slurry was then uniformly coated onto a 0.02 μm aluminum (Al) foil current collector and dried at 60 °C under vacuum for 12 hours.
:
1 v/v mixture of DOL and DME) was added to each cathode. The cyclic voltammetry (CV) behaviors of the symmetric batteries were evaluated at a scan rate of 10 mV s−1, with a voltage window spanning from −0.8 to 0.8 V.
:
1 of sulfur and Li2S dissolved in a 1 M lithium bis(trifluoromethanesulfonyl)imide (LiTFSI) electrolyte with 2 wt% lithium nitrate (LiNO3) in 1,2-dimethoxyethane (DME)
:
1,3-dioxolane (DOL) (1
:
1 v/v). The solution was heated to 70 °C and vigorously stirred to ensure complete sulfur dissolution. Li–S was assembled in an argon-filled glovebox with H2O < 0.01 ppm and O2 < 0.01 ppm in a typical assembly using CR2032 coin-type cells. 12 mm SnS2@MoS2@C coated on aluminum foil served as the cathode, and lithium foil served as the anode. Sulfur loading was maintained at ∼1 mg cm−2 with a complementary electrolyte/sulfur ratio of 15 μL mg−1. Also, a 19 mm Celgard 2400 was used as a separator. Ultimately, the electrochemical performance was evaluated using a Neware BTS 8.0 battery test system, in a voltage window of 1.7–2.8 V. Cyclic voltammetry (CV) and electrochemical impedance spectroscopy (EIS) were conducted on a CHI-760E electrochemical workstation at a scan rate of 0.1 mV s−1 to 0.5 mV s−1 and in the frequency range of 10 kHz to 0.01 Hz, respectively.
Simultaneously, carbonization of resorcinol-formaldehyde (RF) enhanced the porosity and structural stability of the nanospheres. Furthermore, the shell acted as a physical barrier to minimize polysulfide dissolution in lithium–sulfur batteries and a conductor for ion and electron transport.24,25
SEM images in Fig. 1b and S1a show the highly exposed composite edges of SnS2@MoS2 nanosheets uniformly and vertically grown onto a spherical yolk–shell matrix.26 The TEM images in Fig. 1c and S1b further illustrate the porous nature of yolk–shell nanospheres. They displayed an average particle size of 410 nm. The average shell thickness and yolk size were measured to be 27 nm and 186 nm, respectively. This offered the necessary void spaces that can effectively confine polysulfides and enhance electrochemical performance.27,28 To contextually understand the yolk–shell formation, SEM and TEM images of the control samples were compared side by side. As shown in Fig. S1c and d, without the addition of the Sn precursor, hollow-shell MoS2@C microspheres were formed. The further absence of the RF precursor resulted in the formation of amorphous MoS2 nanospheres (Fig. S1f and g). The diameter of MoS2@C exhibited an average size of ∼900 nm and a shell thickness of ∼120 nm, while MoS2 nanospheres exhibited ∼600 nm overall. It is evident that the unique features exhibited in SnS2@MoS2@C yolk–shell nanospheres were synergically harnessed by the incorporation of the Sn precursor.
Compared to the control samples (Fig. S1e and i), the high-resolution TEM (HRTEM) image of SnS2@MoS2@C in Fig. 1d further displayed 1T/2H-MoS2 crystalline and ordered nanosheets composed of a cross-linked interplanar spacing of about 0.68 nm and 0.72 nm corresponding to the (002) crystal plane.29–32 A d-spacing of 0.27 nm was attributed to the (001) plane of SnS2 nanosheets, which were displayed through the entire structural matrix.33,34 Similar lattice fringes were observed by Liu et al. in the fabrication of a three-dimensional MoS2/SnS2-RGO anode for advanced sodium batteries and capacitors.35 Also, the associated EDS mapping demonstrated the homogeneous distribution of tin (Sn), molybdenum (Mo), sulfur (S), carbon (C), and oxygen (O) elements within the yolk–shell composite nanospheres (Fig. 1e and S2c). For further clarity and as evidence of the elemental distribution, Fig. S2a and b present the yolk–shell nanostructure with matching cross-sectional morphology of SnS2@MoS2@C nanospheres.
The crystal structures of SnS2@MoS2@C nanospheres and its controls were also subjected to XRD analysis, as shown in Fig. 2a. The samples again displayed mostly 1T/2H-MoS2 (JCPDS no. 37-1492) diffraction peaks indexed to 17.8° (002), 33.3° (100), and 58.5° (110) planes.36,37 The diffraction peaks corresponding to MoS2 in SnS2@MoS2@C were notably sharper and more intense than those in MoS2@C. This indicates enhanced crystallinity and reduced amorphous content in the composite while promoting better ordering of the MoS2 phase.38 Also, the improvement can be attributed to the incorporation and uniform distribution of SnS2, as confirmed by HRTEM and EDS-mapping. In contrast, the weaker peaks observed in bare MoS2 can be ascribed to the poor transition from amorphous or less ordered Mo–S to the oriented MoS2 layer.39,40
These structural modifications are consistent with the observed shifts in the Raman spectra (Fig. 2b). Also, 1T and 2H phases existed in SnS2@MoS2@C and MoS2@C nanocomposites with E2g1 and A1g modes of the hexagonal MoS2 crystal.41–43 MoS2@C showed characteristic peaks at 365 and 400 cm−1, while SnS2@MoS2@C moved to a slightly higher frequency with lower intensity peaks at 376 and 406 cm−1, respectively. This shift, along with the reduction in peak intensity in SnS2@MoS2@C, could be attributed to defect formation and altered interlayer interactions resulting from the integration of SnS2. Moreover, the relative intensity of ID/IG was calculated to be 0.86, suggesting a relatively high degree of graphitization.44 From these observations, it is obvious that MoS2 is the major species in the composite. Although SnS2 plays a critical role in the formation of yolk–shell nanospheres and is uniformly distributed throughout the composite, its presence is almost undetectable by XRD and Raman spectroscopy. This is due to its relatively low concentration of the Sn precursor and strong signals from MoS2. Evidently, these observations are consistent with reports on low-loading heterostructures (e.g., SnS2/Bi2WO6,45 and MoS2/SnS2 (ref. 46 and 47)). Furthermore, the reduced Raman signals in the yolk–shell nanospheres can be an indication that SnS2 and MoS2 exist as few-layer or highly dispersed within the carbon matrix.48,49
Based on the TGA curves (Fig. 2c), between 25 °C and 120 °C, there was minimal mass loss attributed to the residual water adsorbed in the samples.50 After an initial weight loss occurs at around 430 °C, the SnS2@MoS2@C composite retained 61.74% of its original weight, due to sulfur loss occurring between 100 and 430 °C. This weight reduction was primarily due to the combustion of carbon, the oxidation of MoS2 to MoO3, and the transformation of SnS2 into SnO2. An additional 24.66% weight loss occurred after the plateau between 700 and 800 °C, credited to continued MoS2 oxidation and MoO3 sublimation. In comparison, hollow-shell MoS2@C retained 44.21% of its weight at a lower plateau. On the other hand, pure MoS2 reported similar initial weight loss to SnS2@MoS2@C. The final weight drop beyond 800 °C in all samples corresponded to MoO3 volatilization. Consequently, SnS2@MoS2@C retains 21.5% mass at 900 °C, compared to 10.6% for MoS2@C and 1.5% for MoS2.
As calculated by ICP (Table S2), the contents of SnS2 and MoS2 in SnS2@MoS2@C were estimated to be 9.77% and 64.48%, respectively. Similarly, MoS2@C contained 46.8% of MoS2. These findings are in close agreement with the TGA results and suggest that the weight loss was mainly due to the MoS2 fraction. The additional 10.9% residue in SnS2@MoS2@C relative to MoS2@C also aligns well with the SnS2 fraction, confirming that the formation of SnO2 accounted for the greater thermal stability.
The specific surface area of the nanospheres was calculated by BET analysis to ascertain the influence of SnS2 on the ternary nanocomposite (Fig. 2d). Owing to the H3 hysteresis loop at a relative pressure of 0.4–0.9, all three samples exhibited type IV isotherms with mesoporous structures.51–53 SnS2@MoS2@C exhibited the highest specific surface area of 124.34 m2 g−1 and total pore volume (0.251 cm3 g−1). This was greater than those of both MoS2@C (75.95 m2 g−1; 0.114 cm3 g−1) and MoS2 (61.86 m2 g−1; 0.062 cm3 g−1). As shown in Fig. 2e, SnS2@MoS2@C also exhibited a rich mesoporous distribution, with an average pore diameter of 8.08 nm, compared to MoS2@C and MoS2. These characteristics highlight the effectiveness of the yolk–shell structure of SnS2@MoS2@C to provide more accessible active sites and improved pore connectivity. Moreover, the large surface area could facilitate efficient adsorption of polysulfide intermediates, reduce the shuttle effect, and enhance the overall performance and lifespan of Li–S batteries.54 This further suggests that the yolk–shell nanospheres stoichiometrically balance SnS2/MoS2 to enhance their surface properties for electrochemical applications.51
Further investigations were carried out to evaluate the efficacy of SnS2@MoS2@C in adsorbing Li polysulfides (LiPs) before electrochemical assessment. Fig. S3 details the presence of the synthesized nanomaterials in Li2S6 solution at time zero, and the inset of Fig. 2f depicts the adsorption performance. With the introduction of SnS2@MoS2@C and MoS2@C, the Li2S6 solution was almost completely decolorized, while MoS2 exhibited subtle alterations in color after standing for four hours inside a glovebox. The adsorption capability of SnS2@MoS2@C was further corroborated by ultraviolet-visible (UV-vis) spectroscopy, revealing the absence of absorbance peaks. This shows that SnS2@MoS2@C nanospheres have abundant specific surface area to facilitate the physical adsorption of Li polysulfides.
The chemical composition and states of the individual elements in the composite structures were also examined using XPS spectra. In Fig. 3a, the survey scan of SnS2@MoS2@C identified Sn, Mo, S, and C in the composite. The survey spectra of MoS2@C and MoS2 (Fig. S4a) showed Mo, S, and C signals. The C 1s signal in MoS2 arose from residual organic matter. In Fig. 3b, the two peaks at 494.8 and 486.4 eV correspond to Sn 3d3/2 and Sn 3d5/2 binding energies (BE) of the SnS2 state, respectively.55,56Fig. 3c also shows 1T and 2H phases of MoS2 at around 229, 232, and 235 eV assigned to Mo 3d5/2, Mo 3d3/2 orbitals, and Mo6+, similar to those of MoS2@C and MoS2 shown in Fig. S4b.42,57 The peak at Mo6+ was ascribed to Mo–N bonding and matches the Mo2N (3p3/2) orbital, which may be caused by the Mo and N (Mo–N bond) coordination.58
![]() | ||
| Fig. 3 (a) XPS survey scan and (b–e) corresponding high-resolution XPS spectra of Sn, Mo, S, and C elements of SnS2@MoS2@C. | ||
Compared to MoS2@C and MoS2 in Fig. S4c, the yolk–shell was identified at peaks 162.5 and 163.6 eV, which correspond to S 2p3/2 and S 2p1/2 bonding, respectively. These were mainly ascribed to Mo–S 2p3/2 and Mo–S 2p1/2 of MoS2 as shown in Fig. 3d.59 The additional peak that appears at 161.58 eV resulted from the SnS2 in the ternary yolk–shell nanospheres and was ascribed to Sn–S 2p1/2.60Fig. 3e identified three deconvoluted peaks in the C 1s spectrum. The peak at C–C (284.8 eV) was assigned to graphite carbon as the major species, and a weak peak of C–O resulted at 285.8 eV.61,62 Additionally, the peak at 290.3 eV was assigned to O–C
C, inherited from the oxygen-containing precursors.63,64 From the XPS elemental ratios (Table S3), SnS2@MoS2@C shows an S/Mo value of 2.66 and a Sn/Mo value of 0.20, indicating a sulfur-rich dual-sulfide system with clear Sn incorporation. In contrast, MoS2@C exhibits a near-stoichiometric S/Mo ratio of 1.98, confirming the compositional differences that distinguish the two materials. The wt% also confirmed that SnS2@MoS2@C contained 7.6% Sn, 31.0% Mo, and 27.6% S, while MoS2@C mainly has 32.6% Mo and 21.5% S without Sn, close to the ICP findings.
To gain insight into the electrochemical performance of the synthesized nanomaterials, the electrocatalytic performance of yolk–shell SnS2@MoS2@C and its controls was evaluated. These were carried out using Li2S6 in symmetric battery cells in the voltage range of −0.8 to 0.8 V (Fig. 4a). SnS2@MoS2@C exhibited the highest redox current of 0.14 V and the lowest polarization of −0.12 V, showing its superior polysulfide conversion and ion transport capabilities.65,66 In contrast, MpmoS2@C (a = 0.20 V, b = −0.21 V) and MoS2 (a = 0.36 V, b = −0.26 V) displayed lower redox currents and higher polarization, indicating significantly reduced catalytic activity. Furthermore, the SnS2@MoS2@C cathode displayed the highest exchange current density in both reduction and oxidation processes that were fitted between (−0.2 and 0.2) V of the Tafel plot (Fig. S5a). It also showed the smallest Tafel slope extrapolated to obtain the I0 value according to the Butler–Volmer equation.67,68 Additionally, in the absence of Li2S6 in the electrolyte, a blunt-shaped CV curve was yielded (Fig. S5b). This suggests that the process is a chemical reaction rather than a capacitive behavior, making Li2S6 a unique electrochemically active species in the system.69
Cyclic voltammetry (CV) analysis conducted between 1.7 and 2.8 V revealed two cathodic peaks responsible for the multiple-step sulfur reduction process (Fig. 4b). The first cathodic peak 1 (C1) was assigned to the transformation of S8 into soluble long-chain lithium polysulfides (Li2Sx, 4 ≤ x ≤ 8). Again, further reduction to insoluble Li2S2/Li2S was evidenced by the second cathodic peak 2 (C2).70 The anodic peak (A) was attributed to the reverse oxidation reaction of the short-chain sulfides to lithium polysulfides and then to sulfur. The SnS2@MoS2@C electrode displayed the most heightened peak currents and a larger peak area, an indication of the smaller polarization in the SnS2@MoS2@C battery.71 The multiple redox pathways and enhanced reaction kinetics in SnS2@MoS2@C make it the optimal catalyst for battery applications. After four stable cycles of the CV test at 0.1 mV s−1 with good redox reversibility (Fig. S6a–c), SnS2@MoS2@C demonstrated the lowest charge transfer resistance (Rct), as evidenced by the smallest semicircle in the Nyquist plot (Fig. 4c). The simulated interfacial impedance of the Li–S battery demonstrated a substantial reduction from 97.91 Ω to 45.47 Ω as the electrode material was changed from MoS2 to SnS2@MoS2@C (Table S4). This decrease highlights the superior charge transfer efficiency and interfacial electrochemical activity of the SnS2@MoS2@C composite, outperforming MoS2@C and MoS2.22 The results also significantly contribute to enhancing the battery performance of the SnS2@MoS2@C composite.
The corresponding Tafel slope analysis of the CV plot revealed that SnS2@MoS2@C exhibited significantly enhanced catalytic activity in both reduction and oxidation processes compared to MoS2@C and MoS2, as shown in Fig. 4d–f and the fitted Tafel slope in Fig. S7. The lower cathodic Tafel slope of SnS2@MoS2@C showed more efficient polysulfide conversion with reduced energy barriers. These improvements are attributed to its yolk–shell structure and the existence of SnS2, and the thin, porous shell. Similarly, in the anodic process, SnS2@MoS2@C demonstrated superior charge transfer kinetics, further emphasizing the critical role of SnS2 and yolk–shell engineering in facilitating redox reactions.
At different scan rates (0.1 to 0.5 mV s−1), the peak currents of SnS2@MoS2@C showed increased slopes compared to MoS2@C and MoS2 (Fig. 4g–i). Furthermore, the linear fitting (Fig. 4j–l) and diffusion coefficient of lithium (DLi+) reflect the enhanced lithium polysulfide conversion kinetics of SnS2@MoS2@C compared to MoS2@C and MoS2 as calculated by using the Randles–Sevcik equation in Table S5. Thus, this suggests a faster Li+ diffusion on the surface of yolk–shell morphology with the ability to synergistically modify the nano-structural properties.72
The rate performance of the yolk–shell electrode was evaluated at different current rates (0.1–3C). As shown in Fig. 5a and S8a, SnS2@MoS2@C delivered the highest discharge capacities of 1445, 1340, 1184, 1092, 942, and 802 mA h g−1 at 0.1, 0.2, 0.5,1, 2, and 3C, respectively. When the current density returns to 0.1C, and then 0.2C, the discharge capacity bounced up to 1311 mA h g−1, revealing a good stability rate.73 In contrast, MoS2@C and MoS2 showed lower capacities at these current densities. MoS2 especially revealed poor discharge capacities of 239 and 51 mA h g−1 at 2 and 3C, respectively. However, when returned to 0.1 and 0.2C, MoS2@C and MoS2 also showed discharge capacities with recoveries similar to their initial current densities. The corresponding charge/discharge curves of the batteries at different current rates are shown in Fig. S8b and c. In an overall observation, the discharge curves displayed two discharge plateaus with an increase in current density. However, in MoS2, the discharge plateau became extremely short, and the voltage differences between charge and discharge plateaus widened. This is an indication of polarization at higher current densities, which is consistent with rate cycling performance. SnS2@MoS2@C, on the other hand, showed complete and wider plateaus of three discharge stages, thus providing discharge capacities with no deterioration even at higher current densities.
Fig. 5b further displayed the optimal performance of SnS2@MoS2@C yolk–shell nanospheres compared to MoS2@C and MoS2 through the initial charge/discharge evaluation of the electrodes at a current density of 0.1C. The enhanced polarization in SnS2@MoS2@C can be attributed to the effective synergistic integration of SnS2 and MoS2, which created additional active sites for redox reactions to accelerate lithium polysulfide conversion kinetics.62 Furthermore, in Fig. 5c, the phase conversion coefficient (QH/QL) profile of SnS2@MoS2@C produced a higher coefficient of 2.46 than that of MoS2@C (2.42) and MoS2 (2.21). Here, QH is associated with the conversion process of S8 to Li2S4, and QL with the conversion process of Li2S4 to Li2S2/Li2S.74 While all values are below the theoretical maximum of 3, they align with expected trends due to the complex multi-electron transfer processes and the LiPS shuttle effect.75,76 Additionally, their corresponding overpotentials (ΔE) revealed a linear decrease from 0.12 V, 0.15 V, and 0.20 V corresponding to SnS2@MoS2@C, MoS2@C, and MoS2, respectively. Ultimately, the differential capacity (dQ/dV) curves (Fig. S6d–f) and the corresponding charge/discharge profiles (Fig. S6g–i) of the cathode materials corroborated the cycling stability. Furthermore, the optimal electrochemical performance of SnS2@MoS2@C compared to the control sample during the first three cycles at 0.1C was consistent with the CV profile.
In terms of constant stability assessment, the SnS2@MoS2@C electrode displayed exceptional cycling stability and a coulombic efficiency (CE) of 98.75% at a current density of 0.2C, as shown in Fig. 5d and Table S6. It achieved an initial capacity of 1355.4 mA h g−1, retaining 1224.6 mA h g−1 after 100 cycles, with a capacity fading rate of 0.097% per cycle. This performance highlights the robust structural integrity of the yolk–shell and its electrochemical stability. In contrast, the MoS2@C electrode exhibited an initial capacity of 1108.0 mA h g−1, which declined to 1089.0 mA h g−1 after 100 cycles. It also presented a higher fading rate of 0.26% per cycle. Meanwhile, the bare MoS2 electrode, with an initial capacity of 1066.3 mA h g−1, retained 897.2 mA h g−1, displaying a fading rate of 0.15% per cycle. These results emphasize the superior performance of the SnS2@MoS2@C yolk–shell nanospheres, attributed to the synergistic interaction between SnS2 and MoS2 layers, which enhanced the interlayer spacing that aided ion transport while mitigating structural degradation.62
Furthermore, the yolk–shell SnS2@MoS2@C electrode demonstrated remarkable long-term cycle stability. The capacity increased from 1044.8 mA h g−1 to 1114.6 mA h g−1 over 600 cycles with a slight capacity increase of 0.01% and CE of 95.22% (Fig. 5e). This capacity increase could be attributed to an activation process, where the electrode structure becomes more accessible to ions during extended cycling.77,78 This characteristic could have led to improved utilization of active materials. In comparison, the MoS2@C electrode showed a capacity improvement from 613 mA h g−1 to 724.6 mA h g−1, which is a moderate structural improvement. The bare MoS2 electrode suffered some capacity decay, decreasing from 680.71 mA h g−1 to 358.34 mA h g−1, with a fading rate of 0.079% per cycle. To better demonstrate the performance advantages of the SnS2@MoS2@C yolk–shell nanospheres, the synthesis steps and electrochemical performance were compared with those of other highly studied TMS cathodes. As shown in Table S7, the SnS2@MoS2@C yolk–shell nanospheres outperformed most other electrodes in terms of rate capacities.
The remarkable cycling stability of SnS2@MoS2@C yolk–shell nanospheres stems from their structural and chemical features. The hollow interior provides sufficient space to buffer polysulfides during volume expansion, while the robust and porous shell preserves the framework. Elemental mapping confirmed that Sn, Mo, S, and C are uniformly distributed in both yolk and the shell, suggesting a cooperative role in structural integrity and electrochemical activity. Moreover, SnS2 and MoS2 possess a strong affinity for polysulfides, enabling strong adsorption that could confine the active material within the cathode. At the same time, the carbon layer physically confines them, thereby suppressing dissolution and shuttle effects over long-term cycling.
Supplementary information: additional structural characterizations (SEM, TEM, HRTEM, EDX, XPS), electrochemical analyses (CV, Tafel, EIS, charge/discharge), BET & ICP data, and comparative performance tables. See DOI: https://doi.org/10.1039/d5na00772k.
Footnote |
| † These authors contributed equally to this paper. |
| This journal is © The Royal Society of Chemistry 2025 |