Design of carbon@WS2 host with graham condenser-like structure for tunable sulfur loading of lithium-sulfur batteries

Yue Wang Wenli Hu Binchao Shi He Jia Shilin Mei Chang-Jiang Yao

Citation:  Yue Wang, Wenli Hu, Binchao Shi, He Jia, Shilin Mei, Chang-Jiang Yao. Design of carbon@WS2 host with graham condenser-like structure for tunable sulfur loading of lithium-sulfur batteries[J]. Chinese Chemical Letters, 2025, 36(6): 110065. doi: 10.1016/j.cclet.2024.110065 shu

Design of carbon@WS2 host with graham condenser-like structure for tunable sulfur loading of lithium-sulfur batteries

English

  • On pursuing future energy storage systems (ESSs) with high output performances and sustainability, a series of new ESSs beyond lithium-ion batteries (LIBs) have attracted growing attention due to the bottle-neck of LIBs towards higher energy/power density [1-3]. As one of the most promising candidates for the next-generation ESSs, lithium-sulfur (Li-S) batteries outperform others due to the low cost of nature abundant sulfur and high theoretical capacity (1675 mAh/g). Up to date, significant progresses have been achieved regarding the notorious problems that hinder the practical application of Li-S batteries, such as the "shuttle effect" of lithium polysulfides in state-of-the-art ether-based electrolytes, the sluggish kinetics due to the insulating nature of sulfur species, and the destructive volume expansion of the lithiated cathode [4-7]. Numerous cathodes involving advanced catalysis and well-defined structures have been developed to realize high performance even with lean electrolyte [8-13]. With respect to prototype Li-S batteries, high capacity, high energy density, and acceptable power density are undoubtedly necessary to adopt their practical use, which requires a high mass loading of the active material [14-16]. However, in pouch cells with high-loading-cathodes, achieving fast kinetics remains a long-standing challenge since the aforementioned problems are even more sever in these thick electrodes. Specifically, the reaction kinetics is even more sluggish due to the proportional increase in the electron- and ion transport distance. Moreover, on account of the increased sulfur concentration gradient in the high-loading-cathode, the shuttle-effect tends to be intensified, leading to sever capacity decay during cycling. Besides, increasing areal loading usually leads to new challenges such as fracturing and delamination of the thick electrodes. In principle, building long-range conducting paths for efficient electron and ion transfer has been the most straightforward approach to eliminate the retarded kinetics and improve the stability of thick electrodes, thus to realize comparable electrochemical performances to that under low sulfur loading [14].

    Among various sulfur host materials, low-tortuous arrays featuring both short and continuous electron- and ionic conducting paths have been demonstrated beneficial to achieving high-rate and prolonged cycling life of lithium batteries [17-21]. Besides facilitated mass transport, strong confinement and fast conversion of polysulfides are additional preset for sulfur cathode for practical application. In this regard, one-dimensional nanostructures such as hollow carbon nanotube (CNT) stand out from many other structures due to their straight conducting route and tubular space for encapsulating and confining of sulfur species [22,23]. In addition, strong interactions between the nonpolar carbon and the polar Li2S can be introduced by surface modification of amphiphilic organic molecules on the inner carbon wall to further constrain the sulfur species. Although more complicated structures were reported afterwards such as small CNT embedded micro carbon tubes and spheres-in-tube structures [24,25], the rational design of one-dimensional sulfur host still needs systematic study to realized optimized output performances under high areal loading of sulfur.

    Motivated by the aforementioned research, a multifunctional C@WS2 host with coil-in-tube structure inspired by the Graham condenser was designed in purpose to induce rapid electron and ion transfer, tunable sulfur loading, strong confinement and fast conversion of polysulfide species. The one-dimensional carbon nanotubes with internal carbon coils allow efficient charge transfer across through the thickness of the cathode and provide strong physical confinement to polysulfide diffusion towards both the lateral and longitudinal directions. The Few-layer WS2 loaded in the carbon coils provides additional adsorption sites for polysulfides, performing a synergistic role in suppressing the shuttle-effect and boosting the cathodic kinetics.

    Nanoporous AAO membrane (200 nm in diameter) with vertically aligned nanochannels and block copolymer of polystyrene-b-poly(2-vinylpyridine) (PS-b-P2VP) are used as hard and soft template, respectively. The inner coil is induced by PS-b-P2VP, which forms helix micelles inside the AAO channels driven by the confined phase separation [26]. As shown in Fig. S1 (Supporting information), PS-b-P2VP/THF+DMF solution was firstly filled into the AAO channels via drop wetting and subsequently immersed into water to induce a fast phase separation. Helix micelles were generated due to the microphase separation under the spatial confinement of AAO nanochannels driven by the fast extraction of THF+DMF from PS-b-P2VP into water. The total molecular weight which determines the micelle size and the volume ratio of PS: P2VP that presents its hydrophilic property are found to be crucial in the formation of helix coils. PS50000-b-P2VP16500 with a volume ratio of PS: P2VP = 3.03 was demonstrated adoptable for the fabrication of helix coils under the confinement of 200 nm nanochannels, while those with larger or smaller PS content would form disordered structures due to the weaker or stronger interactions between the hydrophilic P2VP and water. In order to transform the PS-b-P2VP micelles into carbon nanotubes, polydopamine (PDA) was applied as carbon source and uniformly coated on the surface of PS-b-P2VP as well as the inner wall of AAO channels (Fig. 1a, step 1). WS2 nanodomains are generated by the adsorption of WS42− ions and the subsequent decomposition under acid conditions (pH-3) (Fig. 1a, step 2). Finally, the PS-b-P2VP template was removed by a thermal annealing process and carbon tubes with helix coils inside can be obtained. The prepared C@WS2 host was filled with sulfur and used as cathode in the subsequent process (Fig. 1a, step 3).

    Figure 1

    Figure 1.  (a) Schematic illustration of the fabrication process for the C@WS2 composites. (b, c) SEM images of AAO and AAO-C@WS2 surfaces. (d-g) SEM images of PS-b-P2VP coils inside the AAO channels. (h) EDS mapping of the cross-section view of the AAO-C@WS2 membrane. (i) TEM image of the C@WS2 coil-in-tube structures. Inset: HRTEM image of the WS2 domains.

    Figs. 1b and c show the morphology change of the AAO membrane before and after filling with the C@WS2 tubes, in which the color changes from white to black. Fig. 1d shows the thickness of the AAO membrane is about 60 µm, which is filled with PS-b-P2VP helix coils (Figs. 1eg) inside the channels. EDS mapping of the C@WS2 tubes inside the AAO membrane show homogeneous distribution of S, C, N, W elements, which were derived from the carbonized PDA and incorporated WS2 (Fig. 1h). TEM image further confirms the replica of the PS-b-P2VP template by PDA, showing the well-defined coil-in-tube structure after carbonization (Fig. 1i). High-resolution TEM (HRTEM) image reveals the formation of few-layer WS2 inside the amorphous carbon where crystal domains with an adjacent lattice fringe of 0.62, 0.27, 0.16, and 0.26 nm corresponding to the (002), (100), (101) and (110) plane of WS2 can be found, suggesting the successful loading of WS2 (Fig. S2 in Supporting information).

    XRD pattern of the C@WS2 host is shown in Fig. 2a. Sharp reflexes at 2θ of 33.6° and 58.4° corresponding to the 101 and 110 plane of WS2 are observed (PDF No. 84-1398) [27,28]. In addition, a broad peak centered at 24.8° can be indexed to the amorphous phase-dominated carbon derived from PDA. In XPS spectra, the W 4f spectrum (Fig. 2b) [29-33] is fitted by three characteristic peaks of W4+, ascribed to W 4f7/2 (32.9 eV) W 4f5/2 (35.0 eV). Besides, the peaks at 36.0 and 38.7 eV are assigned to W 4f7/2, W 4f5/2 of W6+, respectively, which can be attributed to a slight oxidation of the WS2 surface. In the S 2p [34-36] spectrum (Fig. S3 in Supporting information), the peaks at 162.5 and 163.3 eV belong to S 2p3/2, S 2p1/2 of sulfidic nature. The splitting signals positioned at approximately 164.3 and 165.6 eV are assigned to the C—S—C bond, which is derived from the interaction between PDA and WS42- during the synthetic process of WS2. In the C 1s spectrum (Fig. 2c), peaks at 284.5 and 286.1 eV correspond to the C—C and C—O bonds, respectively. The existing of C-N bond at 285.2 eV indicates the generation of N-doped carbon from PDA. The N 1s [37,38] spectrum in Fig. 2d shows three peaks at 398.9, 400.5, and 401.4 eV, which can be ascribed to three nitrogen's chemical states, namely the pyridinic N (N-6), pyrrolic N (N-5), and graphitic N (N-G), respectively. The N-doped carbon not only can provide additional polar sites to adsorb polysulfides, but also can accelerate the redox reaction via efficient electron transfer of graphitic N.

    Figure 2

    Figure 2.  (a) XRD pattern and high-resolution XPS spectrum of (b) W 4f, (c) C 1s, and (d) N 1s for the C@WS2 composites.

    The S/C@WS2 cathode was prepared by the melting-infusion method at 155 ℃ with different areal loading of sulfur and assembled into Li-S coin cells with 100 µm lithium foil as the anode. TEM image shows that the melt sulfur was successfully filled inside the tubular structure (Fig. S4 in Supporting information). The cathode kinetics is systematically analyzed to evaluate the functions of the designed C@WS2 host with coil-in-tube structure. Carbon nanotubes (CTs) generated from AAO without coils and WS2 nanocatalyst were applied for comparison. The galvanostatic intermittent titration technique (GITT) is adopted to decouple the total cathodic polarization (ηtotal) during discharge into activation polarization (ηac), concentration polarization (ηcon), and ohmic polarization (ηohm) [39]. Specifically, the sum of ηohm and ηac can be indicated by the instantaneous voltage jump at the end of the discharge step, and the ηcon can be represented by the following voltage recovery to the equilibrium voltage in the relaxation stage. As shown in Figs. 3ac, the decoupled ηtotal at the Li2S nucleation stage (depth of discharge between 27% and 70%, Fig. S5 in Supporting information) is compared between the C@WS2 and CTs host. C@WS2 host with WS2 nanocatalyst and coil-in-tube structure show smaller ηohm + ηac and ηcon, suggesting that the C@WS2 host shall significantly reduce the polarization required for interfacial charge transfer determined by the activation energy of electrode reaction (ηac) and ohmic resistance of electrolyte under a certain current (ηohm), as well as the concentration polarization caused by the concentration differences of the reaction species between the electrode surface and the bulk electrolyte (ηcon). In addition, the Li2S precipitation test with constant potential discharge at 2.05 V vs. Li/Li+ is performed to validate the catalytic performance of the C@WS2 host [40-43]. As shown in Fig. 3d and Fig. S6 (Supporting information), the C@WS2 host exhibits clear response of Li2S nucleation (at 6450 s), while the CTs reference shows no obvious Li2S nucleation peak, indicating efficient electrocatalytic activity of WS2 towards lithium polysulfides. According to Faraday's law, the Li2S precipitation capacitiy of C@WS2 based cathode is 294.4 mAh/g, which is higher than that of CTs based cathode (59.19 mAh/g), implying more conversion of lithium polysulfides to Li2S. Therefore, C@WS2 not only promotes the initial nucleation but also guides the subsequent deposition of Li2S, showing a favorable catalytic effect on the redox reaction owing to the highly active sites of few-layer WS2. Further evaluation on the Li2S growth behavior is conducted using the classical Scharifer-Hills (SH) and Bewick-Fleischman-Thirsk (BFT) models to fit the curves of current vs. time in the potentiostatic discharging process (Fig. 3e). The fitted result shows that the Li2S growth on the C@WS2 host matches better with the 2DI nucleation model, suggesting the Li2S growth rate will be controlled by lattice bonding, which can be attributed to the existence of sulfurphilic WS2.

    Figure 3

    Figure 3.  (a, b) Decoupled total cathodic polarization (ηtotal) of the S/C@WS2 and S/CTs batteries based on the GITT measurements. (c) Polarization analysis at the Li2S nucleation stage. (d) Current-time plots of catholyte Li2S8 on the S/C@WS2 cathodes potentiostatically discharged at 2.06 V. (e) The dimensionless transient of Li2S deposition in comparison with theoretical 3D and 2D electroplating models (t: time, tm: time needed to reach the maximum current; i: current, im: maximum current). (f) Lithium-ion diffusion rate and (g, h) tafel plots of different cathodes.

    The electrochemical properties of the S/C@WS2 electrode were further revealed by CV measurements under various scan rates ranging from 0.02 mV/s to 0.1 mV/s (Fig. S7 in Supporting information). The C@WS2 cell exhibits two typically cathodic peaks at about 2.3 V and 1.93 V and two closed (merged) anodic peaks at around 2.4 V corresponds to the reversible sulfur redox reactions (Figs. S7a and b). Li+ diffusion coefficients (DLi+) can be calculated by the classical Randles-Sevcik equation (), where ip is the peak current measured in the CV curve, z is the number of transferred charges, A is the area of the electrode, ν is the scanning rate, CLi is the concentration of Li+ in the electrolyte. The linear fitting relations of ip/ν0.5 of the two sulfur cathodes are shown in Figs. S7c and d and the DLi+ is summarized in Fig. 3f. In comparison with the CTs host, the S/C@WS2 exhibits greatly improved Li+ transfer in the cell. Moreover, the kinetics in liquid-solid reaction and solid-liquid reaction is further verified by the Tafel slope derived from CV curve (Fig. S8 in Supporting information). As shown in Figs. 3g and h, the S/C@WS2 cathode shows smaller Tafel slope for all the three redox peaks, indicating better bidirectional catalysis of C@WS2 than that of CTs.

    Fig. 4a shows the galvanostatic charge/discharge profiles of the S/C@WS2 cathode, from which small voltage hysteresis (ΔE = 0.14 V) can be observed, corresponding to small electrochemical polarization. Electrochemical impedance spectroscopy (EIS) is used to monitor the changes of the inner resistance related to the structure and electrochemical evolution during the cycling. Compared to the initial state before cycling, the charge transfer resistance (Rct) and Warburg impedance of the S/C@WS2 battery greatly decrease after 5 cycles due to the redistribution of sulfur species after the activation process. In the subsequent cycles the impedance stable, suggesting good reversibility of the redox reactions (Fig. S9 in Supporting information). Fig. 4b illustrates the cycling performances of the S/C@WS2 battery at moderate sulfur loading of approximately 3 mg/cm2 under different C rates. Reversible specific capacities of 1180, 1014, 920, 695 and 200 mAh/g are obtained at 0.1, 0.2, 0.5, and 1 C, respectively. When the current rate returns to 0.1 C, reversible specific capacity of 986 mAh/g is sustained, which is comparable to the initial state. Long-term cycling of the S/C@WS2 battery at 0.5 C has been investigated, which shows a high discharge capacity of 900 mAh/g and a remarkable capacity retention of 79% after 500 cycles. Analysis on the capacity contributions (denoted as QH and QL, corresponding to the conversion of long-chain polysulfides (the higher plateau) and short-chain polysulfides (the lower plateau), respectively) have been conducted for the battery with different sulfur loading. As shown in Fig. 4c, the highest QL/QH of 3.09 has been achieved with the areal loading of 8.4 mg/cm2 at 0.02 C, which is quite close to the ideal contribution of the redox conversion from Li2S4 to Li2S (-75% of the total discharge capacity). Another two high QL/QH values can be observed in the case of 4.5 and 5.4 mg/cm2 at 0.1 C (2.95 and 2.79, respectively), suggesting that full redox conversion shall be realized under high mass loading with appropriate current density.

    Figure 4

    Figure 4.  (a) Charge/discharge profile of the S/C@WS2 battery at 0.1 C. (b) Rate performance and cycling performance at 0.5 C. (c) Capacity contributions of QL and QH of the batteries with different sulfur loading at different C rates. (d-f) Cycling performance and the corresponding areal capacity of the batteries different sulfur loading at different C rates.

    In practical applications, the amount of sulfur loading at the cathode is one of the key indicators that determine the actual energy density of Li-S batteries. In previous studies, a low sulfur loading usually helps to improve the specific capacity and cycling stability, but it leads to a dramatic decrease in the overall energy density of the battery. For the sake of further confirming the functionality of the C@WS2 host, tunable sulfur loading of 2.6, 3.4, 4.5, 5.4 and up to 8.4 mg/cm2 has been investigated with the E/S ratio of 10 µL/mgS (Figs. 4df). Typical discharge plateau is clearly displayed in the charge/discharge curves at different current densities (Fig. S10 in Supporting information). Notably, the kinetic voltage of the S/C@WS2 battery shows almost no decrease as the sulfur loading increased to 8.4 mg/cm2, dictating the smooth electron/ion transfer and efficient catalysis of the C@WS2 host. As shown in Figs. 4df, the initial specific capacity of the batteries reaches 1145 mAh/g for 2.6 mg/cm2 at 0.2 C, 1028 mAh/g for 3.4 mg/cm2 at 0.2 C, 1045 mAh/g for 2.6 mg/cm2 at 0.1 C, 881 mAh/g for 8.4 mg/cm2 at 0.02 C, corresponding to areal capacity of 3.0, 3.5, 5.6 and 7.4 mg/cm2, respectively. The coulombic efficiency for different sulfur loading keeps above 98.8%, suggesting good reversibility of the charge-discharge reactions under high sulfur loading.

    The excellent electrochemical performances can be attributed to the synergistic effect of the C@WS2 host with coil-in-tube structure, which not only provides short path for charge transfer, but also strong adsorption to polysulfides and efficient catalysis for the conversion reaction (Fig. 5a). Specifically, the host material serves as efficient conducting framework which enable electron transfer from both the tube wall and the inner coil. The one-dimensional tubular structure shall regulate the ion transfer within a short pathway, which is favorable for fast kinetics. Notably, sulfur species could be seriously constrained inside the Graham condenser-like tubes. During discharging, the lateral diffusion of lithium polysulfides would be prohibited by the carbon tube wall and the longitudinal diffusion shall pass through the inner coil, during which lithium polysulfides would be blocked by the strong adsorption (Fig. S11 in Supporting information) of the WS2 nanodomains [44] and the N-doped carbon. Simultaneously, the kinetics of soluble polysulfides could be facilitated at the "triple-phase" interface. Fig. 5b shows the diffusion experiment of Li2S6 performed with Celgard, AAO-CTs, and AAO-C@WS2 membranes as filters. The colorless solution on the right side of the chamber after 24 h confirms the significant intercept of polysulfides for the C@WS2 host, while the two references show very limited intercepting capability. The strong confinement of the coil-in-tube structure to polysulfides is further verified by decomposition of the cycled batteries and the electron probe X-ray microanalysis (EPMA) results of the lithium anode (Figs. 5c and d). The lighter color in the separator and lithium anode of the S/C@WS2 battery as well as the much less sulfur species on the surface of lithium anode in comparison with that of the S/CTs batteries confirm the mitigated "shuttle effect" by the Graham condenser-like C@WS2 host.

    Figure 5

    Figure 5.  (a) Schematic illustration of the synergistic effect derived from the C@WS2 host with coil-in-tube structure. (b) Li2S6 permeation experiment in H-shaped glass container with AAO-C@WS2 separator, AAO-C separator, and Celgard. (c, d) Images of disassembled coin cells using CTs and C@WS2 as sulfur host and the corresponding SEM images and EPMA mapping of the Li anode.

    In summary, we designed and fabricated a novel C@WS2 host material with coil-in-tube structure inspired by the Graham condenser for tunable loading of sulfur. Due to the advantages of both the novel structure and well-designed chemical components, the cathode material shows good electron-and ionic conductivity, strong adsorption and electrocatalytic ability to lithium polysulfides, which effectively promotes excellent electrochemical performances of Li-S batteries with tunable sulfur loading. Under a moderate sulfur loading of 3 mg/cm2, the S/C@WS2 cathode delivers remarkable high specific capacity of 1180 mAh/g at 0.1 C and long-cycling stability with a 79% capacity retention at 0.5 C after 500 cycles. Under high sulfur loading of 8.4 mg/cm2, impressive areal capacity of 7.4 mAh/cm2 has been realized at 0.02 C. The synergistic effect derived from the combination of structural and compositional design propose a new way to tackle the key problems of Li-S batteries on the way towards practical use.

    The authors have no conflicts to declare.

    Yue Wang: Writing – original draft, Investigation, Formal analysis, Data curation. Wenli Hu: Visualization, Formal analysis. Binchao Shi: Investigation, Data curation. He Jia: Methodology, Visualization. Shilin Mei: Writing – review & editing, Supervision, Methodology, Conceptualization. Chang-Jiang Yao: Writing – review & editing, Project administration, Funding acquisition, Formal analysis.

    We are grateful to the National Natural Science Foundation of China (Nos. 22075027, 52003030), Starting Grant from Beijing Institute of Technology and financial support from the State Key Laboratory of Explosion Science and Safety Protection (Nos. YBKT21-06, YBKT23-05), and Beijing Institute of Technology Research Fund Program for Young Scholars.

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2024.110065.


    1. [1]

      X. Ji, K.T. Lee, L.F. Nazar, Nat. Mater. 8 (2009) 500–506. doi: 10.1038/nmat2460

    2. [2]

      P.G. Bruce, S.A. Freunberger, L.J. Hardwick, J.M. Tarascon, Nat. Mater. 11 (2012) 19–29. doi: 10.1038/nmat3191

    3. [3]

      Y.X. Yin, S. Xin, Y.G. Guo, L.J. Wan, Angew. Chem. Int. Ed. 52 (2013) 13186–13200. doi: 10.1002/anie.201304762

    4. [4]

      W. Sun, Z. Song, Z. Feng, et al., Nano-Micro Lett. 14 (2022) 222. doi: 10.1007/s40820-022-00954-x

    5. [5]

      Z. Ye, Y. Jiang, L. Li, F. Wu, R. Chen, Adv. Mater. 32 (2020) 2002168. doi: 10.1002/adma.202002168

    6. [6]

      Y. Tian, G. Li, Y. Zhang, et al., Adv. Mater. 32 (2020) 1904876. doi: 10.1002/adma.201904876

    7. [7]

      Q. Shao, S. Zhu, J. Chen, Nano Res. 16 (2023) 8097–8138. doi: 10.1007/s12274-022-5227-0

    8. [8]

      S. Evers, L.F. Nazar, Acc. Chem. Res. 46 (2013) 1135–1143. doi: 10.1021/ar3001348

    9. [9]

      A. Manthiram, Y. Fu, S.H. Chung, C. Zu, Y.S. Su, Chem. Rev. 114 (2014) 11751–11787. doi: 10.1021/cr500062v

    10. [10]

      S. Chandra, P. Dutta, K. Biswas, ACS Nano 16 (2022) 7–14. doi: 10.1021/acsnano.1c10584

    11. [11]

      D. Yang, Z. Liang, C. Zhang, et al., Adv. Energy Mater. 11 (2021) 2101250. doi: 10.1002/aenm.202101250

    12. [12]

      D. Yang, Z. Liang, P. Tang, et al., Adv. Mater. 34 (2022) 2108835. doi: 10.1002/adma.202108835

    13. [13]

      D. Yang, C. Li, M. Sharma, et al., Energy Storage Mater. 66 (2024) 103240. doi: 10.1016/j.ensm.2024.103240

    14. [14]

      M. Wang, Z. Bai, T. Yang, et al., Adv. Energy Mater. 12 (2022) 2201585. doi: 10.1002/aenm.202201585

    15. [15]

      B.Q. Li, Q. Zhang, Sci. China Chem. 64 (2021) 337–338. doi: 10.1007/s11426-020-9915-y

    16. [16]

      J.S. Kim, T.H. Hwang, B.G. Kim, J. Min, J.W. Choi, Adv. Funct. Mater. 24 (2014) 5359–5367. doi: 10.1002/adfm.201400935

    17. [17]

      B. Wang, W. Guo, Y. Fu, ACS Appl. Mater. Interfaces 12 (2020) 5831–5837. doi: 10.1021/acsami.9b19196

    18. [18]

      Y. Shi, B. Li, Y. Zhang, et al., Adv. Energy Mater. 11 (2021) 2003663. doi: 10.1002/aenm.202003663

    19. [19]

      K. Huang, P. Zhai, J.S. Chen, et al., Electrochem. Commun. 144-145 (2022) 107395. doi: 10.1016/j.elecom.2022.107395

    20. [20]

      Z. Wu, Z. Cai, B. Fang, et al., ACS Appl. Mater. Interfaces 13 (2021) 25890–25897. doi: 10.1021/acsami.1c02951

    21. [21]

      S.X. Liu, J.L. Tian, W. Zhang, Nanotechnology 32 (2021) 222001. doi: 10.1088/1361-6528/abe25f

    22. [22]

      G. Zheng, Q. Zhang, J.J. Cha, et al., Nano Lett. 13 (2013) 1265–1270. doi: 10.1021/nl304795g

    23. [23]

      Z. Li, B.Y. Guan, J. Zhang, X.W. Lou, Joule 1 (2017) 576–587. doi: 10.1016/j.joule.2017.06.003

    24. [24]

      Y. Ge, Z. Chen, S. Ye, et al., J. Mater. Chem. A 6 (2018) 14885–14893. doi: 10.1039/c8ta05041d

    25. [25]

      F. Jin, S. Xiao, L. Lu, Y. Wang, Nano Lett. 16 (2016) 440–447. doi: 10.1021/acs.nanolett.5b04105

    26. [26]

      P. Hou, H. Fan, Z. Jin, Macromolecules 48 (2015) 272–278. doi: 10.1021/ma501933s

    27. [27]

      Y. Liu, J. Li, B. Liu, et al., ChemSusChem 16 (2023) e202201200. doi: 10.1002/cssc.202201200

    28. [28]

      L. Yang, X. Zhu, S. Xiong, et al., ACS Appl. Mater. Interfaces 8 (2016) 13966–13972. doi: 10.1021/acsami.6b04045

    29. [29]

      S. Xiao, J. Zhang, Y. Deng, et al., ACS Appl. Energy Mater. 3 (2020) 4923–4930. doi: 10.1021/acsaem.0c00488

    30. [30]

      M.S. Bazarjani, M. Hojamberdiev, K. Morita, et al., J. Am. Chem. Soc. 135 (2013) 4467–4475. doi: 10.1021/ja3126678

    31. [31]

      F. Xie, M. Xiong, J. Liu, et al., ChemElectroChem 9 (2022) e202200474. doi: 10.1002/celc.202200474

    32. [32]

      B. Zhang, C. Luo, Y. Deng, et al., Adv. Energy Mater. 10 (2020) 2000091. doi: 10.1002/aenm.202000091

    33. [33]

      S. Wei, M. Serra, S. Mourdikoudis, et al., ACS Appl. Mater. Interfaces 14 (2022) 46386–46400. doi: 10.1021/acsami.2c06295

    34. [34]

      S. Dey, K. Manjunath, A. Zak, G. Singh, ACS Omega 8 (2023) 10126–10138. doi: 10.1021/acsomega.2c07464

    35. [35]

      S. Huang, Y. Wang, J. Hu, et al., ACS Nano 12 (2018) 9504–9512. doi: 10.1021/acsnano.8b04857

    36. [36]

      H. Lin, L. Yang, X. Jiang, et al., Energy Environ. Sci. 10 (2017) 1476–1486. doi: 10.1039/C7EE01047H

    37. [37]

      J. Jang, S.H. Song, H. Kim, et al., ACS Appl. Mater. Interfaces 13 (2021) 14786–14795. doi: 10.1021/acsami.1c02892

    38. [38]

      T. Li, B. Ding, J. Wang, et al., ACS Appl. Mater. Interfaces 12 (2020) 14993–15001. doi: 10.1021/acsami.9b18883

    39. [39]

      X. Guan, H. Pei, X. Chen, et al., J. Colloid Interface Sci. 652 (2023) 997–1005. doi: 10.1016/j.jcis.2023.08.130

    40. [40]

      C. Ding, M. Niu, C. Cassidy, et al., Small 19 (2023) 2301755. doi: 10.1002/smll.202301755

    41. [41]

      R. Sun, M. Qu, L. Peng, et al., Small 19 (2023) 2302092. doi: 10.1002/smll.202302092

    42. [42]

      T. Huang, G. Zhang, R. Chen, et al., ACS Appl. Mater. Interfaces 15 (2023) 21075–21085. doi: 10.1021/acsami.3c01558

    43. [43]

      X. Zhang, Z. Liu, W. Liu, J. Han, W. Lv, ACS Appl. Mater. Interfaces 15 (2023) 19002–19010. doi: 10.1021/acsami.3c01512

    44. [44]

      T. Lei, W. Chen, J. Huang, et al., Adv. Energy Mater. 7 (2017) 1601843. doi: 10.1002/aenm.201601843

  • Figure 1  (a) Schematic illustration of the fabrication process for the C@WS2 composites. (b, c) SEM images of AAO and AAO-C@WS2 surfaces. (d-g) SEM images of PS-b-P2VP coils inside the AAO channels. (h) EDS mapping of the cross-section view of the AAO-C@WS2 membrane. (i) TEM image of the C@WS2 coil-in-tube structures. Inset: HRTEM image of the WS2 domains.

    Figure 2  (a) XRD pattern and high-resolution XPS spectrum of (b) W 4f, (c) C 1s, and (d) N 1s for the C@WS2 composites.

    Figure 3  (a, b) Decoupled total cathodic polarization (ηtotal) of the S/C@WS2 and S/CTs batteries based on the GITT measurements. (c) Polarization analysis at the Li2S nucleation stage. (d) Current-time plots of catholyte Li2S8 on the S/C@WS2 cathodes potentiostatically discharged at 2.06 V. (e) The dimensionless transient of Li2S deposition in comparison with theoretical 3D and 2D electroplating models (t: time, tm: time needed to reach the maximum current; i: current, im: maximum current). (f) Lithium-ion diffusion rate and (g, h) tafel plots of different cathodes.

    Figure 4  (a) Charge/discharge profile of the S/C@WS2 battery at 0.1 C. (b) Rate performance and cycling performance at 0.5 C. (c) Capacity contributions of QL and QH of the batteries with different sulfur loading at different C rates. (d-f) Cycling performance and the corresponding areal capacity of the batteries different sulfur loading at different C rates.

    Figure 5  (a) Schematic illustration of the synergistic effect derived from the C@WS2 host with coil-in-tube structure. (b) Li2S6 permeation experiment in H-shaped glass container with AAO-C@WS2 separator, AAO-C separator, and Celgard. (c, d) Images of disassembled coin cells using CTs and C@WS2 as sulfur host and the corresponding SEM images and EPMA mapping of the Li anode.

  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  155
  • HTML全文浏览量:  3
文章相关
  • 发布日期:  2025-06-15
  • 收稿日期:  2024-03-16
  • 接受日期:  2024-05-27
  • 修回日期:  2024-04-27
  • 网络出版日期:  2024-05-28
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章