Coordinating lithium polysulfides to inhibit intrinsic clustering behavior and facilitate sulfur redox conversion in lithium-sulfur batteries

Qihou Li Jiamin Liu Fulu Chu Jinwei Zhou Jieshuangyang Chen Zengqiang Guan Xiyun Yang Jie Lei Feixiang Wu

Citation:  Qihou Li, Jiamin Liu, Fulu Chu, Jinwei Zhou, Jieshuangyang Chen, Zengqiang Guan, Xiyun Yang, Jie Lei, Feixiang Wu. Coordinating lithium polysulfides to inhibit intrinsic clustering behavior and facilitate sulfur redox conversion in lithium-sulfur batteries[J]. Chinese Chemical Letters, 2025, 36(5): 110306. doi: 10.1016/j.cclet.2024.110306 shu

Coordinating lithium polysulfides to inhibit intrinsic clustering behavior and facilitate sulfur redox conversion in lithium-sulfur batteries

English

  • Lithium-sulfur (Li-S) batteries have been hailed as a promising alternative to conventional lithium-ion batteries owing to their high theoretical specific energy density (2600 Wh/kg), low cost, and environmental friendliness [1-5]. The sulfur redox reactions occurring in the Li-S battery positive electrode commonly involve the complicated multiphase transformation processes between elemental sulfur and the Li2S discharge product accompanied by the generation of dissoluble lithium polysulfide intermediates (Li2Sn, 4 ≤ n ≤ 8) [6]. The dissolution and redeposition of polysulfide species are believed to enable the reversible conversion of the insulating sulfur/Li2S particles via a solution-mediated reaction pathway [7-9]. However, some detrimental impacts are also caused. The dissolved polysulfide species would diffuse to the lithium metal anode and participate in the parasitic reactions during the cycling process, thus leading to the rapid capacity fading and poor cycling performance of Li-S batteries [10,11]. Most of the efforts to overcome this degradation mechanism have focused on trapping and/or confining the polysulfide species to reduce the active material loss by constructing functional sulfur hosts, modified separators or interlayers [12-18]. Furthermore, other issues of concern should include the intrinsic clustering behavior of lithium polysulfides and subsequently inadequate solution-mediated kinetics [9], which have often been overlooked by researchers.

    Recent theoretical computation studies and molecular dynamics simulations demonstrate that the Li+-Sn2− bond network would be formed in the conventional ether-based electrolyte due to the electrostatic interactions between lithium cations and polysulfide dianions [19,20]. Lithium polysulfide monomers subsequently tend to aggregate to generate polysulfide clusters ([Li2Sn]m, m > 1), which are found to be the more stable configuration compared to discrete, isolated polysulfide chains, especially under low temperature [21,22]. Moreover, Anderson and colleagues presented that the major mechanism of Li+ exchange is likely the dynamic processes involving the aggregation and dissolution of polysulfide clusters that shuttle Li+ from one cluster to another [23]. However, the dissociation into individual polysulfide chains and further lithiation would be greatly suppressed by the severe steric barriers and sturdy electrostatic connections in the highly tangled and disordered aggregates [24]. Moreover, the mass transport of polysulfide species would also be limited due to the large size of formed clusters, which could restrain the electrochemical reaction kinetics and following electrodeposition of Li2S [8]. It is thus of utmost importance to develop efficient approaches to subdue the unfavorable polysulfide clustering behavior and facilitate the reaction kinetics of sulfur conversion.

    Recently, the coordination configuration of lithium polysulfide clusters could be altered and disrupted through the introduction of additional ions and solvents in the electrolyte. For example, the coordination state of lithium salt and lithium polysulfide can influence and depend on each other due to the competitive electrostatic attraction between the salt anion and the polysulfide-adjoined lithium cation [25,26]. The addition of some strongly bonded lithium salts, such as lithium trifluoroacetate (LiTFA) and lithium nitrate (LiNO3), can weaken the strong Li+-Sn2− bond network and mitigate the polysulfide clustering phenomena [27,28]. Similarly, the forcefully binding cationic groups, such as NH4+, also show the potential to deter the formation of undesirable clustered polysulfide aggregates because of the favorable NH4+-Sn2− ionic associations [29]. In addition, solvents with high donor numbers, such as dimethylsulfoxide (DMSO) and dimethylacetamide (DMA), exhibit vigorous solvation with Li+, which could contribute to the enhanced Li+-solvent interactions and attenuated polysulfide clustering, thereby promoting the solution-mediated reactions of polysulfides [30,31]. Furthermore, the enhanced mediated effect also has been proposed in recent years by introducing a redox comediator to react with lithium polysulfides to afford the products with improved redox mediation capacity [32-36]. For example, diphenyl diselenide (DPDSe) is proven to spontaneously react with lithium polysulfides to generate lithium phenylseleno polysulfides (LiPhSePSs), which could reduce the energy barrier for multiphase sulfur conversion and increase the deposition dimension of Li2S [34]. Moreover, the solubility and redox activity of polysulfide intermediates was effectively regulated by the functional couple molecules of 2,4-bis(p-tolythio)−1,3-dithia-2,4-diphosphetane-2,4-disulfide (BPHS) [37]. Sulfur conversion kinetics were consequently improved with the elimination of the shuttle effect, exhibiting excellent cycling performance of Li-S batteries.

    Here, based on the principles of disrupting the Li+-Sn2− bond network of polysulfides and the improved mediation effect, we introduce molybdenum pentachloride (MoCl5) into the electrolyte of Li-S batteries, which also has been reported as a dual-function redox mediator in Li-O2 batteries [38]. Through a series of experimental and theoretical characterizations, MoCl5 could voluntarily coordinate with lithium polysulfides via the formation of robust Mo-S bond. The as-produced complexes can not only inhibit the intrinsic clustering behavior of lithium polysulfides but also serve as an improved mediator with the bi-functional catalytic effect for Li2S deposition and activation. Moreover, due to the coordination bonding and accelerated conversion reaction, the dissolution and shuttling of polysulfides also can be greatly restrained. When tested in a Li-S coin cell, this so-called polysulfide complexant allows for good long-term cycling stability with a capacity decay of 0.078% per cycle after 400 cycles at 2 C. Additionally, it demonstrates excellent rate performance with a discharge capacity of 589 mAh/g at 4 C. An area capacity of 3.94 mAh/cm2 is achieved with a high sulfur loading of 4.5 mg/cm2 at 0.2 C. The low-temperature performance is also evidenced at −20 ℃, showing a high discharge capacity of 741 mAh/g after 80 cycles.

    The commercially available MoCl5 (Fig. S1 in Supporting information) was chosen and introduced into the electrolyte of the Li-S battery to explore its potential application for inhibiting intrinsic polysulfide clustering. As shown in Fig. 1a, the color of the Li2S6 solution would be changed from yellow to dark red after merging with the MoCl5 solution. Moreover, the UV–vis spectrum of Li2S6 solution exhibits three significant absorption peaks at 630, 420, and 300 nm, corresponding to S3•−, S42−, and S62− species, respectively [39]. After mixing with MoCl5, the characteristic peaks of S3•− and S42− would disappear, which should be attributed to the chemical interaction between polysulfides and MoCl5. This also can be confirmed by Raman spectroscopy. As presented in Fig. 1b, the Li2S6 solution could demonstrate a characteristic Raman peak located at 450 cm−1, assigning to the S32−, S42−, and S4 species [40], which are obviously observed to vanish after blending with MoCl5. Furthermore, X-ray photoelectron spectroscopy (XPS) was conducted to clarify the specific interaction between MoCl5 and Li2S6. As revealed in Fig. 1c, the XPS spectrum of S 2p for Li2S6 exhibits two typical terminal sulfur (ST) and bridge sulfur (SB) signals at 161.7/162.9 eV and 163.1/164.3 eV, respectively [41]. After reacting with MoCl5, the additional peaks at the binding energy of 163.3/164.5 eV could be observed, belonging to the emergence of the Mo-S bonds accompanied by the weakened peak intensity of ST. Moreover, in the XPS spectra of Mo 3d (Fig. S2 in Supporting information), the extra peaks at 232.5/235.7 eV could be monitored, validating the formation of Mo-S bonds. 7Li NMR analysis was further conducted for the 0.1 mol/L Li2S4-containing electrolytes with or without MoCl5 to investigate the effect of MoCl5 on the polysulfides clustering. As shown in Fig. 1d, the 7Li chemical shift moves upfield with the addition of MoCl5, indicating the enhanced electron cloud density around Li+, which should be induced by the construction of a stronger electron shielding environment due to the formed Mo-S bond [34,42]. All the above results evidentially demonstrate that MoCl5 could spontaneously react with lithium polysulfides via the coordination between Mo and ST, thus disturbing the intrinsic Li+-Sn2− bond network of polysulfides and alleviating the polysulfides clustering.

    Figure 1

    Figure 1.  Chemical interactions between MoCl5 and lithium polysulfide. (a) The UV–vis spectra and corresponding optical photos of the supernatant (Li2S6, MoCl5, and Li2S6+MoCl5 solutions). (b) The Raman spectra of DME/DOL, Li2S6, and Li2S6+MoCl5 solutions. (c) XPS spectra of S 2p for Li2S6 and Li2S6+MoCl5. (d) 7Li NMR spectra obtained from the DME/DOL-solvated Li2S4 with or without MoCl5. (e) CV curves at a scan rate of 10 mV/s and (f) EIS spectra of Li2S6 and Li2S6+MoCl5 symmetrical cells (the inset is the equivalent circuit diagram). (g) The Gibbs free energy for the reaction between Li2S6 and MoCl5 from step Ⅰ to step Ⅶ.

    Symmetric cells were assembled based on Li2S6 electrolyte with or without MoCl5 to evaluate the reaction kinetics of polysulfides conversion. As demonstrated in Fig. 1e, the polysulfides symmetric cell with MoCl5 additive exhibits the higher current response, indicating that the coordination interaction of MoCl5 can facilitate more adequate polysulfides conversion. Moreover, the corresponding electrochemical impedance spectra (EIS) display that the polysulfides symmetric cell with MoCl5 additive shows a distinctly lower charge transfer resistance (Rct, 902 Ω) compared with that in the absence of MoCl5 (Rct, 1451 Ω), further confirming the promoted polysulfides redox reactions (Fig. 1f). The spontaneous reaction trend has further been verified through theoretical calculations as the Gibbs free energy for the reaction of MoCl5 and Li2S6 is −12.83 kcal/mol (Fig. 1g). Hence, their coordination reaction can proceed thermodynamically spontaneously to decouple the intrinsic clustering behavior of the intermediate polysulfides.

    To further investigate the catalytic effect for Li2S nucleation and activation, Li||Li2S8 cells were assembled to conduct the potentiostatic discharging and galvanostatic intermittent titration technique (GITT) tests, respectively. As displayed in Figs. 2a and b, and Fig. S3 (Supporting information), the Li||Li2S8 cell with MoCl5 additive shows a Li2S deposition capacity of 310.83 mAh/g, which is almost twice that of the blank cell (161.52 mAh/g). Besides, the cells were disassembled to characterize the deposited morphology of Li2S via the SEM image. The insets demonstrate the typical lateral growth of Li2S in the absence of MoCl5, which would terminate the Li2S nucleation after completely covering the conductive carbon substrate. By contrast, the smooth deposition and further two-dimensional growth of insulating Li2S could emerge after the addition of MoCl5, suggesting that the liquid-solid transformation from polysulfides towards Li2S precipitation could be facilitated by coordinating with MoCl5, then promoting the radial growth of Li2S [42]. As presented in Figs. 2c and d, the internal resistance of each electrochemical reaction is proportional to the voltage difference between quasi-open-circuit-voltage (QOCV) and closed-circuit-voltage (CCV) at the same current [43]. The Li||Li2S8 cell with MoCl5 additive shows visibly smaller internal resistance for Li2S nucleation and activation than those without MoCl5, which indicates that the introduction of MoCl5 could alleviate polarization and kinetic limitations of Li2S redox reactions [44].

    Figure 2

    Figure 2.  Catalytic effect for Li2S nucleation and activation. Potentiostatic discharge curves of Li2S deposition (a) without and (b) with MoCl5 and its corresponding SEM images. (c, d) GITT measurements. (e) The UV–vis spectra and corresponding optical photos of the supernatant (MoCl5, Li2S and Li2S+MoCl5 solutions). (f) Initial charging voltage profiles of Li||Li2S cells with or without MoCl5.

    Additionally, the interaction between Li2S and MoCl5 was studied. As shown in Fig. 2e, the colorless Li2S solution turns aurantia after mixing with the MoCl5 solution. The UV–vis absorption spectrum of the mixed solution is distinct from those of individual components. Two additional peaks could be observed at 480 nm and 400 nm, which should correspond to S82−/S62− and S42− species, respectively [39]. This demonstrates that MoCl5 can chemically oxidize Li2S into polysulfides, thus showing the potential as the redox mediator to promote the activation of insulating Li2S. Furthermore, Li2S cathodes were prepared and utilized to assemble the Li||Li2S cells. As illustrated in Fig. 2f, the activation energy barrier of Li2S can be effectively reduced by adding MoCl5 into the electrolyte, further confirming the mediation effect of MoCl5 for Li2S oxidation.

    Optically transparent Li-S cells were fabricated with lithium metal anodes and sulfur cathodes with the sulfur content of 58.28 wt% (Fig. S4 in Supporting information). As shown in Fig. 3a, plenty of yellow species would be produced from the sulfur cathode and diffuse into the blank electrolyte during the cell discharging process at 0.1 C, which should be ascribed to the generation of soluble polysulfides. In comparison, the dissolution and shuttling of polysulfides can be visibly suppressed with MoCl5 added into the electrolyte. In the discharge process of sulfur cathodes, the generated polysulfides would immediately coordinate with MoCl5 to produce slightly soluble products (Fig. S5 in Supporting information), which achieves the same effect as reducing the solubility of polysulfides. Coupling with the inhibited polysulfides clustering and faster reaction kinetics, the modified electrolyte can effectively suppress the polysulfides shuttle behaviors. Moreover, a yellow film-shaped layer is visually formed on the surface of the sulfur cathode, but this phenomenon is unobserved in the case of a blank electrolyte. The surface morphologies of the cathodes in the completely discharged state demonstrate that the sulfur electrode in the blank electrolyte would be rough and loose due to the dissolution loss of polysulfides, while a compact and smooth layer can be constructed to efficaciously inhibit the dissolution and diffusion of polysulfides. Furthermore, the chemical composition of the dense layer was characterized by XPS analysis. As displayed in the XPS spectra of S 2p in Fig. 3b, the peaks at 168.6/169.8 eV, 167.2/169.4 eV, and 160.6/161.8 eV correspond to typical polythionate, thiosulfate, and lithium sulfide (Li2S) signals [41]. Compared to the sulfur cathode in the blank electrolyte, the peak intensity of thiosulfate is significantly reduced, and a new Mo-S peak appears after adding MoCl5. Besides, the XPS spectra of Mo 3d also verify the formation of the Mo-S bond (Fig. 3c) [45]. It can be concluded that MoCl5 reacts with polysulfides to form complexes via Mo-S bonds during the discharging process of sulfur cathodes, thereby effectively inhibiting the dissolution and diffusion of polysulfides.

    Figure 3

    Figure 3.  Inhibition effect for polysulfides shuttling. (a) In situ visual observations of polysulfide dissolution and diffusion in glass beakers filled with the blank or MoCl5-based electrolytes and the corresponding SEM images of sulfur cathodes in the fully discharged state; XPS spectra of (b) S 2p and (c) Mo 3d for sulfur cathodes collected from the beaker cells.

    To verify the catalytic behaviors of the sulfur redox reactions, cyclic voltammetry (CV) measurements were carried out. As shown in Fig. 4a, the first-cycle CV profiles of the Li-S coin cells exhibit two typical cathodic peaks at the potential of around 2.3 V and 2.0 V, attributing to the reduction of sulfur to soluble polysulfides, then to the insoluble Li2S2 and Li2S. During the anodic segment, two corresponding oxidation peaks can be observed in the potential range of 2.3–2.5 V, which is ascribed to the oxidation of Li2S2/Li2S to polysulfides and then back to elemental sulfur [46]. By introducing MoCl5 into the blank electrolyte, the deposition and oxidation of Li2S would exhibit higher redox peak current and lower polarization, which is consistent with the results in Fig. 2. The good repeatability of CV profiles in the subsequent cycles suggests the favorable reversibility of sulfur conversion (Fig. S6 in Supporting information). Accordingly, Li-S cells with different electrolytes were measured by CV at various scanning rates to evaluate the Li ions diffusion rates and reaction kinetics (Fig. S7 in Supporting information). Based on the Randles-Sevcik equation, a linear relationship between the peak current density and the square root of the scanning rate at each redox peak position suggests the typical characteristic of ion diffusion-controlled reactions [47]. The higher slopes for the Li-S cell using MoCl5-based electrolyte reveal that the Li-ion transport between the formed MoCl5-polysulfide complexes is more achievable [48]. EIS spectra of Li-S cells were conducted before and after 200 cycles at 1 C (Fig. 4b and Fig. S8 in Supporting information). When adding MoCl5 to the electrolyte, the cell exhibits a lower charge transfer resistance (Rct, 15.2 Ω) than the cell with blank electrolyte (31.6 Ω) before cycling. After 200 cycles, the Rct (2.86 Ω) of the cell without MoCl5 additive is still higher than the cell with MoCl5 additive (1.84 Ω), indicating the accelerated charge transfer at the electrode surface in the presence of MoCl5 [49]. Besides, as shown in Fig. S9 (Supporting information), the insulating discharging product of Li2S exhibits a high potential barrier at the initial delithiation process in the cell with blank electrolyte at 0.2 C, while the addition of MoCl5 significantly decreases the activation barrier [50].

    Figure 4

    Figure 4.  Electrochemical performance of Li-S cells. (a) CV curves at the scan rate of 0.1 mV/s. (b) Nyquist plots of fresh Li-S cells. (c) Rate performance (sulfur loading: 1.2 mg/cm2, 1 C corresponding to 1.68 A/gs). (d) Cycling performance at 0.5 C (sulfur loading: 1.7 mg/cm2). (e) Long-term cycling performance at 2 C (sulfur loading: 1.0 mg/cm2). (f) Cycling performance of cells with MoCl5 with high sulfur loading and electrolyte of 30 µL at 0.2 C. (g) Cycle performance and (h) corresponding charge/discharge curves at 0.2 C in −20 ℃.

    As displayed in Fig. 4c, the Li-S cells with different electrolytes deliver subtle differences in discharge capacity at relatively low rates (0.1 C and 0.2 C), but the addition of MoCl5 into electrolyte can enable the higher capacity contribution at high rates (≥0.5 C). Specifically, the cell with MoCl5 additive could exhibit better rate performance, which provides the discharge capacity of 1261, 1136, 1002, 898, 810, 713, and 589 mAh/g at 0.1, 0.2, 0.5, 1, 2, 3 and 4 C, respectively (Fig. S10 in Supporting information). Besides, due to the strong interaction between MoCl5 and polysulfides, the cell with MoCl5 delivers higher capacity at the first plateau, which implies the increased sulfur utilization degree and rapid polysulfides conversion (Fig. S11a in Supporting information) [51]. The obvious valley between the first and second discharge plateaus represents the beginning of Li2S nucleation. The cell with MoCl5 additive displays a distinctly lower polarization voltage gap between Li2S nucleation and activation at different rates (Fig. S11b in Supporting information), again confirming that the introduction of MoCl5 can improve the electrochemical kinetics of the Li2S redox reactions [43]. As shown in Fig. S11c (Supporting information), for the cell with MoCl5, the increase of polarization voltage gap from 0.1 C to 4 C is only 256 mV (from 130 mV to 386 mV), which is considerably smaller than that of the cell without MoCl5 (369 mV, from 124 mV to 493 mV).

    The cycling performance of Li-S cells with different electrolytes was also tested at different rates. The optimal concentration of MoCl5 additive is determined to be 5 mmol/L by comparative results (Fig. S12 in Supporting information), which has been used in other experiments unless otherwise stated. The cell employing a blank electrolyte presents an initial discharging capacity of 950 mAh/g, which would decrease continuously to 817 mAh/g after 150 cycles at 0.5 C (Fig. 4d). In contrast, after adding MoCl5 into the electrolyte, the cell exhibits superior cycling performance with an initial discharge capacity of 1094 mAh/g and a retained capacity of 975 mAh/g for 150 cycles, which demonstrates a very competitive performance compared to the literature results (Table S1 in Supporting information). Even at a relatively higher rate of 2 C (Fig. 4e), the cell using MoCl5 still delivers a discharge capacity of over 590 mAh/g after 400 cycles with a capacity-decay rate of 0.078% per cycle. As a comparison, the cell without MoCl5 can only be cycled 238 times with a large capacity fade rate of 0.135%, which should be caused by the severe shuttle effect of polysulfides and the unfavorable growth of lithium dendrites. Furthermore, the cycling performance of the cell with MoCl5 additive under high sulfur loading was evaluated. As shown in Fig. 4f, under high sulfur loading of 3.2, 3.5, and 4.5 mg/cm2 and low E/S ratio (the ratio of electrolyte volume to sulfur mass) of 9, 8, and 6 µL/mg, the cells could exhibit outstanding initial specific capacities of 972, 949 and 902 mAh/g, which are corresponding to areal capacities of 3.1, 3.3 and 4.1 mAh/cm2, respectively. Particularly, the cell with a sulfur loading of 3.5 mg/cm2 can maintain good cycling stability with a capacity retention rate of 98% after 80 cycles. The two discharge plateaus at different cycles could be obviously observed (Fig. S13 in Supporting information), showing the satisfactory inhibition effect of MoCl5 for polysulfides shuttling and the catalytic activity for sulfur conversion reactions.

    Furthermore, the cycling performance of Li-S cells with or without MoCl5 was estimated at low temperatures (−20 ℃), where the polysulfides clustering issue has been believed to be more serious [52]. As shown in Fig. 4g, the cell using MoCl5 can show good cycling stability at 0.2 C with a high discharge capacity of 741 mAh/g after 80 cycles. In comparison, the Li-S cell would be subjected to the obvious capacity fading with a discharge capacity of 504 mAh/g in the absence of MoCl5. Moreover, the second discharge plateau corresponding to the Li2S nucleation could still be maintained after cycling in the cell with MoCl5, while that in the cell without MoCl5 would almost vanish. The lower polarization voltage also further indicates the distinct catalytic effect and the suppressed behavior of polysulfides clustering (Fig. 4h).

    In the end, the surface morphology of the sulfur cathodes and lithium metal anodes after battery cycling at a fully discharged state were characterized via SEM images. As shown in Fig. 5a, the cell with a blank electrolyte shows serious corrosion on the surface of the cycled Li anode after 200 cycles at 1 C. The morphology of the Li anode becomes rough, appearing irregular pores and dendrites. By adding MoCl5 into the electrolyte, the relevant cycled Li anode shows high surface uniformity with large Li metal grains (Fig. 5b). This should be ascribed to the inhibiting effect of MoCl5 for polysulfide shuttling, then limiting the erosion of metallic Li anodes by polysulfides. Furthermore, the microstructures of cycled sulfur cathodes are obviously diverse (Figs. 5c and d). Significantly fewer precipitations could be observed on the cycled sulfur cathode in the blank electrolyte, confirming the emergence of serious polysulfides shuttling. In the MoCl5-based electrolyte, large particles would be formed and the interfacial passivation could be effectively prevented due to the three-dimensional pattern of nucleation and growth, which can promote sulfur utilization [53]. The interaction between MoCl5 and the lithium metal anode is further clarified, and the introduction of MoCl5 shows a great positive influence on the Li deposition morphology and cycling stability of the cells (Figs. S14 and S15 in Supporting information). Consequently, Li-S cell with MoCl5 demonstrates more standardized interface morphology and framework integrity of electrodes.

    Figure 5

    Figure 5.  Morphology and chemical characterizations of electrodes after battery cycling. The SEM images of (a, b) Li anodes and (c, d) sulfur cathodes. (e, f) XPS spectra of S 2p for sulfur cathodes from Li-S cells without or with MoCl5 (after 200 cycles at 1 C) at a fully discharged state.

    Besides, the interface composition of cycled electrodes also reveals the working mechanism of MoCl5 via XPS analysis. Figs. 5e and f present XPS spectra of S 2p for cycled sulfur cathodes collected from cells with different electrolytes. Compared to the cathode with a blank electrolyte, the introduction of MoCl5 induces new peaks of Mo-S bonds but without the polysulfides signal (162.8/164 eV) (Fig. S16 in Supporting information) [45,54]. This verifies that polysulfides could be effectively anchored via the Mo-S bond, helping to achieve the complete conversion of polysulfides to Li2S. Moreover, other sulfur-containing species (polythionate and thiosulfate) with higher intensity can be clearly observed, which also has been proven to show mediation effect for polysulfides conversion [41]. Consequently, this inner mechanism can accelerate the reaction kinetics of sulfur conversion and thus increase the reversible capacity of the sulfur electrodes.

    In summary, we introduce molybdenum pentachloride (MoCl5) as an electrolyte additive to promote the electrochemical performance of Li-S batteries. The inorganic salt MoCl5 could voluntarily coordinate with lithium polysulfides via the formation of tough Mo-S bonds to generate complexes, which can not only inhibit the intrinsic clustering behavior of lithium polysulfides, but also serve as a modified mediator with the bi-functional catalytic effect for Li2S deposition and activation. In addition, the coordination bonding and accelerated conversion reaction can greatly restrain the dissolution and shuttling of polysulfides. Consequently, the Li-S coin cell with MoCl5 electrolyte additive can exhibit good long-term cycling stability with a capacity decay of 0.078% per cycle after 400 cycles at 2 C, and excellent rate performance with a discharge capacity of 589 mAh/g at 4 C. An area capacity of 3.94 mAh/cm2 is also achieved with a high sulfur loading of 4.5 mg/cm2 at 0.2 C. Moreover, the modified cells work well at −20 ℃ with a high capacity of 741 mAh/g. In this work, we demonstrate the coordinated interaction of MoCl5 with polysulfides and the optimized mediation effect, which could promote the efficient conversion of sulfur species and alleviate the capacity fading issue of Li-S batteries. Our work provides a new perspective for the rational design of electrolyte systems with multiple functions, including the polysulfide clustering inhibition and enhanced catalytic effect for high-performance Li-S batteries.

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

    Qihou Li: Writing – original draft, Supervision, Methodology, Investigation, Formal analysis, Data curation. Jiamin Liu: Writing – original draft, Methodology, Investigation, Formal analysis, Data curation. Fulu Chu: Writing – original draft, Validation, Investigation, Formal analysis, Data curation. Jinwei Zhou: Validation, Methodology, Investigation. Jieshuangyang Chen: Validation, Methodology, Investigation. Zengqiang Guan: Validation, Methodology, Investigation. Xiyun Yang: Writing – original draft. Jie Lei: Writing – review & editing, Writing – original draft, Supervision, Methodology. Feixiang Wu: Writing – review & editing, Writing – original draft, Supervision, Resources, Project administration, Funding acquisition, Conceptualization.

    The authors gratefully acknowledge the National Natural Science Foundation of China (Nos. 51904344, 52172264), and the Natural Science Foundation of Hunan Province of China (Nos. 2021JJ10060 and 2022GK2033).

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2024.110306.


    1. [1]

      A. Manthiram, Y. Fu, Y.S. Su, Acc. Chem. Res. 46 (2012) 1125–1134. doi: 10.1021/ar300179v

    2. [2]

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

    3. [3]

      F. Wu, F. Chu, G.A. Ferrero, et al., Nano Lett. 20 (2020) 5391–5399. doi: 10.1021/acs.nanolett.0c01778

    4. [4]

      R. Deng, M. Wang, H. Yu, et al., Energy Environ. Mater. 5 (2022) 777–799. doi: 10.1002/eem2.12257

    5. [5]

      R. Deng, F. Chu, F. Kwofie, et al., Angew. Chem. Int. Ed. 61 (2022) e202215866. doi: 10.1002/anie.202215866

    6. [6]

      Z. Wang, Y. Li, H. Ji, et al., Adv. Mater. 34 (2022) 2203699. doi: 10.1002/adma.202203699

    7. [7]

      G. Li, S. Wang, Y. Zhang, et al., Adv. Mater. 30 (2018) 1705590. doi: 10.1002/adma.201705590

    8. [8]

      J. Lei, T. Liu, J. Chen, et al., Chem 6 (2020) 2533–2557. doi: 10.1016/j.chempr.2020.06.032

    9. [9]

      Z. Guan, X. Chen, F. Chu, et al., Adv. Energy Mater. 13 (2023) 2302850. doi: 10.1002/aenm.202302850

    10. [10]

      X. Fang, H. Peng, Small 11 (2015) 1488–1511. doi: 10.1002/smll.201402354

    11. [11]

      Y. Huang, L. Lin, C. Zhang, et al., Adv. Sci. 9 (2022) 2106004. doi: 10.1002/advs.202106004

    12. [12]

      X. Liu, J.Q. Huang, Q. Zhang, L. Mai, Adv. Mater. 29 (2017) 1601759. doi: 10.1002/adma.201601759

    13. [13]

      P. Wang, Z. Zhang, B. Hong, et al., J. Electroanal. Chem. 832 (2019) 475–479. doi: 10.1016/j.jelechem.2018.11.044

    14. [14]

      J. Wu, T. Ye, Y. Wang, et al., ACS Nano 16 (2022) 15734–15759. doi: 10.1021/acsnano.2c08581

    15. [15]

      D. Yang, J. Wang, C. Lou, et al., ACS Energy Lett. 9 (2024) 2083–2091. doi: 10.1021/acsenergylett.4c00771

    16. [16]

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

    17. [17]

      D. Yang, Y. Han, M. Li, et al., Adv. Funct. Mater. 34 (2024) 2401577. doi: 10.1002/adfm.202401577

    18. [18]

      C. Li, D. Yang, J. Yu, et al., Adv. Energy Mater. 14 (2024) 2303551. doi: 10.1002/aenm.202303551

    19. [19]

      R.S. Assary, L.A. Curtiss, J.S. Moore, J. Phys. Chem. C 118 (2014) 11545–11558. doi: 10.1021/jp5015466

    20. [20]

      B. Wang, S.M. Alhassan, S.T. Pantelides, Phys. Rev. Appl. 2 (2014) 034004. doi: 10.1103/PhysRevApplied.2.034004

    21. [21]

      C. Park, A. Ronneburg, S. Risse, et al., J. Phys. Chem. C 123 (2019) 10167–10177. doi: 10.1021/acs.jpcc.8b10175

    22. [22]

      A. Gupta, A. Bhargav, J.P. Jones, R.V. Bugga, A. Manthiram, Chem. Mater. 32 (2020) 2070–2077. doi: 10.1021/acs.chemmater.9b05164

    23. [23]

      A. Andersen, N.N. Rajput, K.S. Han, et al., Chem. Mater. 31 (2019) 2308–2319. doi: 10.1021/acs.chemmater.8b03944

    24. [24]

      M. Vijayakumar, N. Govind, E. Walter, et al., Phys. Chem. Chem. Phys. 16 (2014) 10923–10932. doi: 10.1039/C4CP00889H

    25. [25]

      Q. Zou, Z. Liang, G.Y. Du, et al., J. Am. Chem. Soc. 140 (2018) 10740–10748. doi: 10.1021/jacs.8b04536

    26. [26]

      H. Chu, H. Noh, Y.J. Kim, et al., Nat. Commun. 10 (2019) 188. doi: 10.1038/s41467-018-07975-4

    27. [27]

      A. Gupta, A. Bhargav, A. Manthiram, ACS Energy Lett. 6 (2021) 224–231. doi: 10.1021/acsenergylett.0c02461

    28. [28]

      S. Kim, J. Jung, I. Kim, et al., Energy Storage Mater. 59 (2023) 102763. doi: 10.1016/j.ensm.2023.04.002

    29. [29]

      A. Gupta, A. Bhargav, A. Manthiram, Chem. Mater. 33 (2021) 3457–3466. doi: 10.1021/acs.chemmater.1c00893

    30. [30]

      Q. Zou, Y.C. Lu, et al., J. Phys. Chem. Lett. 7 (2016) 1518–1525. doi: 10.1021/acs.jpclett.6b00228

    31. [31]

      A. Gupta, A. Bhargav, A. Manthiram, Adv. Energy Mater. 9 (2019) 1803096. doi: 10.1002/aenm.201803096

    32. [32]

      Z. Lin, Z. Liu, W. Fu, N.J. Dudney, C. Liang, Adv. Funct. Mater. 23 (2013) 1064–1069. doi: 10.1002/adfm.201200696

    33. [33]

      H.L. Wu, M. Shin, Y.M. Liu, K.A. See, A.A. Gewirth, Nano Energy 32 (2017) 50–58. doi: 10.1016/j.nanoen.2016.12.015

    34. [34]

      M. Zhao, X. Chen, X.Y. Li, B.Q. Li, J.Q. Huang, Adv. Mater. 33 (2021) 2007298. doi: 10.1002/adma.202007298

    35. [35]

      H. Ye, J. Sun, X.F. Lim, Y. Zhao, J.Y. Lee, Energy Storage Mater. 38 (2021) 338–343. doi: 10.1016/j.ensm.2021.03.023

    36. [36]

      W. Zhang, F. Ma, Q. Wu, et al., Energy Environ. Mater. 6 (2023) e12369. doi: 10.1002/eem2.12369

    37. [37]

      Q. Zheng, Q. Hou, Z. Shu, et al., Angew. Chem. Int. Ed. 62 (2023) e202308726. doi: 10.1002/anie.202308726

    38. [38]

      X.G. Wang, Z. Zhang, Q. Zhang, et al., J. Mater. Chem. A 7 (2019) 14239–14243. doi: 10.1039/c9ta03628h

    39. [39]

      J. Häcker, D.H. Nguyen, T. Rommel, et al., ACS Energy Lett. 7 (2022) 1–9. doi: 10.1021/acsenergylett.1c02152

    40. [40]

      J.D. McBrayer, T.E. Beechem, B.R. Perdue, C.A. Apblett, F.H. Garzon, J. Electrochem. Soc. 165 (2018) A876. doi: 10.1149/2.0441805jes

    41. [41]

      X. Liang, C. Hart, Q. Pang, et al., Nat. Commun. 6 (2015) 5682. doi: 10.1038/ncomms6682

    42. [42]

      H.J. Peng, G. Zhang, X. Chen, et al., Angew. Chem. Int. Ed. 55 (2016) 12990–12995. doi: 10.1002/anie.201605676

    43. [43]

      J. Park, E.T. Kim, C. Kim, et al., Adv. Energy Mater. 7 (2017) 1700074. doi: 10.1002/aenm.201700074

    44. [44]

      Z. Wu, S. Chen, L. Wang, et al., Energy Storage Mater. 38 (2021) 381–388. doi: 10.1016/j.ensm.2021.03.026

    45. [45]

      H. Shi, Z. Sun, W. Lv, et al., J. Mater. Chem. A 7 (2019) 11298–11304. doi: 10.1039/c9ta00741e

    46. [46]

      F. Liu, C. Zong, L. He, et al., Chem. Eng. J. 443 (2022) 136489. doi: 10.1016/j.cej.2022.136489

    47. [47]

      Y. Dong, D. Cai, T. Li, et al., ACS Nano 16 (2022) 6414–6425. doi: 10.1021/acsnano.2c00515

    48. [48]

      K. Wu, Y. Hu, Z. Cheng, et al., J. Membrane Sci. 592 (2019) 117349. doi: 10.1016/j.memsci.2019.117349

    49. [49]

      D.Q. Cai, J.L. Yang, T. Liu, S.X. Zhao, G. Cao, Nano Energy 89 (2021) 106452. doi: 10.1016/j.nanoen.2021.106452

    50. [50]

      H. Liu, P. Zeng, Y. Li, et al., J. Alloys Compd. 835 (2020) 155421. doi: 10.1016/j.jallcom.2020.155421

    51. [51]

      S. Chen, J. Luo, N. Li, et al., Energy Storage Mater. 30 (2020) 187–195. doi: 10.1016/j.ensm.2020.05.002

    52. [52]

      F. Chu, M. Wang, J. Liu, et al., Adv. Funct. Mater. 32 (2022) 2205393. doi: 10.1002/adfm.202205393

    53. [53]

      H. Pan, J. Chen, R. Cao, et al., Nat. Energy 2 (2017) 813–820. doi: 10.1038/s41560-017-0005-z

    54. [54]

      M. Xiang, J. Li, S. Feng, et al., J. Colloid. Interface Sci. 624 (2022) 471–481. doi: 10.1016/j.jcis.2022.05.118

  • Figure 1  Chemical interactions between MoCl5 and lithium polysulfide. (a) The UV–vis spectra and corresponding optical photos of the supernatant (Li2S6, MoCl5, and Li2S6+MoCl5 solutions). (b) The Raman spectra of DME/DOL, Li2S6, and Li2S6+MoCl5 solutions. (c) XPS spectra of S 2p for Li2S6 and Li2S6+MoCl5. (d) 7Li NMR spectra obtained from the DME/DOL-solvated Li2S4 with or without MoCl5. (e) CV curves at a scan rate of 10 mV/s and (f) EIS spectra of Li2S6 and Li2S6+MoCl5 symmetrical cells (the inset is the equivalent circuit diagram). (g) The Gibbs free energy for the reaction between Li2S6 and MoCl5 from step Ⅰ to step Ⅶ.

    Figure 2  Catalytic effect for Li2S nucleation and activation. Potentiostatic discharge curves of Li2S deposition (a) without and (b) with MoCl5 and its corresponding SEM images. (c, d) GITT measurements. (e) The UV–vis spectra and corresponding optical photos of the supernatant (MoCl5, Li2S and Li2S+MoCl5 solutions). (f) Initial charging voltage profiles of Li||Li2S cells with or without MoCl5.

    Figure 3  Inhibition effect for polysulfides shuttling. (a) In situ visual observations of polysulfide dissolution and diffusion in glass beakers filled with the blank or MoCl5-based electrolytes and the corresponding SEM images of sulfur cathodes in the fully discharged state; XPS spectra of (b) S 2p and (c) Mo 3d for sulfur cathodes collected from the beaker cells.

    Figure 4  Electrochemical performance of Li-S cells. (a) CV curves at the scan rate of 0.1 mV/s. (b) Nyquist plots of fresh Li-S cells. (c) Rate performance (sulfur loading: 1.2 mg/cm2, 1 C corresponding to 1.68 A/gs). (d) Cycling performance at 0.5 C (sulfur loading: 1.7 mg/cm2). (e) Long-term cycling performance at 2 C (sulfur loading: 1.0 mg/cm2). (f) Cycling performance of cells with MoCl5 with high sulfur loading and electrolyte of 30 µL at 0.2 C. (g) Cycle performance and (h) corresponding charge/discharge curves at 0.2 C in −20 ℃.

    Figure 5  Morphology and chemical characterizations of electrodes after battery cycling. The SEM images of (a, b) Li anodes and (c, d) sulfur cathodes. (e, f) XPS spectra of S 2p for sulfur cathodes from Li-S cells without or with MoCl5 (after 200 cycles at 1 C) at a fully discharged state.

  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  109
  • HTML全文浏览量:  8
文章相关
  • 发布日期:  2025-05-15
  • 收稿日期:  2024-06-02
  • 接受日期:  2024-07-30
  • 修回日期:  2024-07-02
  • 网络出版日期:  2024-07-31
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