Xueyu
Wang‡
,
Peirong
Lin‡
,
Chuanhuang
Wu
,
Yuchuan
Zhu
,
Cong
Wang
,
Daying
Guo
*,
Xi'an
Chen
* and
Shun
Wang
*
Key Laboratory of Carbon Materials of Zhejiang Province, College of Chemistry and Materials Engineering, Wenzhou University, Wenzhou, China 325035. E-mail: guody@wzu.edu.cn; xianchen@wzu.edu.cn; shunwang@wzu.edu.cn
First published on 27th August 2024
Herein, we demonstrate a face-centered cubic-structure cobalt–nickel–copper–manganese–molybdenum high-entropy alloy (CoNiCuMnMo-HEA) anchored on a reduced graphene oxide (rGO) substrate as a separator intercalation layer (CoNiCuMnMo-HEA@rGO) for lithium–sulfur batteries. The CoNiCuMnMo-HEA catalyst possesses abundant exposed polymetallic sites and exhibits a strong interaction with polysulfides on the (111) densest stacking surface. Experiments and theoretical calculations prove that Mn and Mo elements play a key role in the reduction process of polysulfides in CoNiCuMnMo-HEA, while Ni and Co are very important in the oxidation process of sulfides. Meanwhile, the presence of Cu element effectively regulates the redox process, which results in HEA exhibiting excellent bidirectional catalysis. As a result, the initial specific discharge capacity of a Li–S battery with the CoNiCuMnMo-HEA@rGO modified membrane can reach 1512.4 mA h g−1 at a rate of 0.2 C. The attenuation rate of each cycle is only 0.055% after 1100 cycles at 3 C, and the coulombic efficiency remains above 91.94%.
In order to solve the above problems, many researchers mainly build multi-level structures and create physical or chemical adsorption sites based on carbon-based materials to improve the utilization rate of sulfur and inhibit the shuttle effect of lithium polysulfides (LiPSs).10 Although these strategies can suppress the problems caused by LiPSs, the reaction kinetics of polysulfides has not been improved; therefore, the stability of the Li–S battery cannot meet the practical requirements.11 Moreover, carbon-based materials loaded with too many metal compounds will significantly reduce the specific surface area and pore volume, thus reducing the sulfur loading.12,13 Fortunately, carbon-based catalytic intercalation on the battery separator not only accelerates the SRR, but also does not affect the sulfur storage at the cathode, which is an ideal strategy.12,14–16 Particularly, intercalation with mono/multi-metallic nanoalloys as catalysts has been extensively studied for Li–S batteries.17,18 For example, Gao et al.18 designed a concave nano-cubic Ni–Pt alloy with high exponential surface constraint, which was introduced as a catalyst for Li–S batteries, significantly promoting the transformation of intermediate LiPSs into solid discharge products. However, due to the limited number of metal active centers, the conversion efficiency of solid sulfur into intermediate LiPSs is low, and it is difficult to achieve a long cycle life.19 Recently, high-entropy alloys formed from various metals have attracted considerable attention due to their high entropy.20 They have many active sites such as lattice distortion effect, cocktail effect and slow diffusion effect and may have good catalytic activity.21–23 For example, Wang et al.24 designed and constructed Fe0.24Co0.26Ni0.10Cu0.15Mn0.25 HEA as the core catalyst which has high electrocatalytic activity for the transformation of solid sulfur into solid discharge products. However, the reverse reaction to polysulfides was not significantly improved. To this end, the development of HEA catalysts with high activity for both bi-directional reactions of polysulfides/sulfides is particularly necessary for the commercial application of Li–S batteries.
Herein, we grow five-membered organometallic framework compounds (CoNiCuMnMo-MOF) on graphene oxide (GO) by a hydrothermal method. Subsequently, the CoNiCuMnMo-HEA nanoparticle-anchored reduced graphene oxide material (rGO) was successfully obtained after pyrolytic reduction (CoNiCuMnMo-HEA@rGO). As a catalytic intercalation material for the Li–S battery separator, it shows excellent performance. This is mainly due to the fact that the diameter of CoNiCuMnMo-HEA is about 12 nm, and it is dispersed in rGO with a high density to avoid agglomeration and form more catalytically active sites. In addition, CoNiCuMnMo-HEA possesses a face-centered cubic structure, and the exposed (111) close-packed surface shows good structural stability. The electrochemical test results show that CoNiCuMnMo-HEA@rGO exhibits a strong adsorption capacity for LiPSs, effectively accelerates the movement and diffusion of Li+, and at the same time has the catalytic advantage of effectively accelerating the conversion of Li2Sn (4 ≤ n ≤ 8) into Li2S2/Li2S. Furthermore, CoNiCuMnMo-HEA@rGO could accelerate the reverse catalytic conversion of Li2S into liquid-phase LiPSs, realizing the bidirectional catalytic conversion of LiPSs. Furthermore, density functional theory (DFT) calculations reveal the unique electron co-regulation properties of CoNiCuMnMo-HEA, where HEA exhibits a lower Fermi energy level than any individual metal, which is more favorable for electron movement and transfer. With these advantages, the specific capacity of CoNiCuMnMo-HEA@rGO can reach 1512.4 mA h g−1 at initial discharge and 570.2 mA h g−1 at 5 C, and the specific capacity decay rate is only 0.055% after that per cycle for 1100 cycles at 3 C. The specific capacity efficiency remains above 91.94%. The efficiency is still above 91.94%.
Fig. 1i shows that the concentration of each metal in the CoNiCuMnMo-HEA@rGO was analyzed by inductively coupled plasma mass spectrometry (ICP), where the mass ratios of Co, Cu, Mn, Mo, and Ni were 33:24:13:10:41. X-ray photoelectron spectroscopy (XPS) was utilized to analyze the surface elemental composition of CoNiCuMnMo-HEA@rGO. The XPS survey spectrum verifies the presence of the elements Co, Cu, Mn, Mo, and Ni (Fig. S5a†). In the high-resolution Mn 2p spectrum of CoNiCuMnMo-HEA@rGO (Fig. 1j), the peaks at 637.54, 641.65, 645.64 and 653.67 eV correspond to Mn0 2p3/2, Mn2+ 2p3/2, Mn3+ 2p3/2 and Mn2+ 2p1/2,26 respectively. The XPS spectra of high-resolution Cu 2p, Co 2p and Ni 2p show the existence of a metal phase and oxide phase (see Fig. S5† for detailed analysis). Compared with other elements, the highest content of Mn element shows a strong oxidation peak, which indicates that Mn metal exhibits strong activity in CoNiCuMnMo-HEA@rGO.26
In the context of electrochemical reactions in Li–S batteries, the rate of the sulfur reduction reaction (SRR) dictates the duration of the existence of soluble LiPSs, which plays a critical role in the overall performance of the battery.11 In evaluating the catalytic activity of the as-prepared materials, lithium metal was used as the anode, reduced graphene oxide with 70% sulfur content (rGO@S) was used as the cathode (Fig. S6†), and CoNiCuMnMo-HEA@rGO and rGO were used as catalytic inserts coated on the separator (named CoNiCuMnMo-HEA@rGO/PP and rGO/PP) to assemble a button cell, respectively. The effect of CoNiCuMnMo-HEA@rGO on SRR kinetics was studied by comparing with rGO. Notably, CoNiCuMnMo-HEA@rGO/PP and rGO/PP did not reveal any material separation after several manual folding tests, indicating that the macroscopic flexibility of the pristine separator was not altered by loading the coating and that it exhibited good adhesion and mechanical strength (Fig. S7†).
Fig. 2a shows the cyclic voltammetry (CV) curves tested at a scan rate of 0.1 mV s−1 within the voltage range of 1.7–2.7 V. The CV curves all exhibit typical sulfur redox peaks, with two cathodic peaks appearing at approximately 2.3 and 2.0 V, indicating two reduction processes. These can be attributed to sulfur first being reduced to soluble long-chain LiPSs, and then further reduced to insoluble Li2S/Li2S2.30 An anodic peak appears at approximately 2.4 V, which can be attributed to the reversible oxidation of Li2S2/Li2S to LiPSs, and eventually to sulfur.12 Compared to the rGO/PP (peak voltage difference of 0.40 V), the redox peak-to-peak voltage difference of the CoNiCuMnMo-HEA@rGO/PP cell (0.32 V) is significantly reduced, indicating that the electrochemical polarization of the battery is smaller with the assistance of CoNiCuMnMo-HEA.31 In addition, the reduction peak of the CoNiCuMnMo-HEA@rGO/PP CV curve shows a significant shift towards higher voltages, suggesting that soluble LiPSs are rapidly transformed.16 To gain a more detailed understanding of the catalytic effect of CoNiCuMnMo-HEA@rGO on redox kinetics, the Tafel slope was calculated from the CV curves to evaluate the catalytic effect of CoNiCuMnMo-HEA@rGO.32 As shown in Fig. 2b, the slope for the reduction of S8 to LiPSs for CoNiCuMnMo-HEA@rGO/PP is 58.9 mV dec−1, which is significantly smaller than that of the rGO/PP cell (81.7 mV dec−1). Additionally, the Tafel slope for the further reduction of LiPSs to Li2S also reveals that the slope for CoNiCuMnMo-HEA@rGO/PP is the smallest (Fig. 2c). This indicates that HEA can promote the solid–liquid and liquid–solid conversion processes of LiPSs. Meanwhile, the slope for the oxidation of Li2S to LiPSs for CoNiCuMnMo-HEA@rGO/PP is 57.7 mV dec−1, which is smaller than that for rGO/PP (Fig. 2d). Analysis of the above shows that the Tafel slope for CoNiCuMnMo-HEA@rGO/PP is significantly smaller than that for rGO/PP, both during lithiation and delithiation processes, suggesting that the presence of HEA accelerates the redox reactions of LiPSs, promotes SRR behavior, and accelerates the oxidation of Li2S.8Fig. 2e presents the electrochemical impedance spectroscopy curves tested under open-circuit voltage for CoNiCuMnMo-HEA@rGO/PP and rGO/PP. The CoNiCuMnMo-HEA@rGO/PP cell exhibits a smaller resistance, which is primarily attributed to the introduction of CoNiCuMnMo-HEA facilitating the conversion of LiPSs and further accelerating the reaction kinetics of Li–S batteries. Additionally, symmetric cells were assembled to test the CV curves for further investigation into the electrocatalytic activity and reversibility of CoNiCuMnMo-HEA@rGO towards LiPS conversion.33 In symmetric cells with 0.1 M Li2S6 as the electrolyte, different catalytic electrodes exhibit distinct peak currents in the CV curves. The symmetric cell with CoNiCuMnMo-HEA@rGO as the electrode shows two pairs of oxidation–reduction peaks with larger current and sharper peak shapes, which are attributed to the two-step reduction of Li2S6 on the working electrode during the cathodic scan and the oxidation of Li2S on the counter electrode during the anodic scan. Moreover, the peak currents of the redox reactions are significantly higher than those on the rGO electrode, indicating that the CoNiCuMnMo-HEA@rGO surface exhibits better electrocatalytic activity for the redox reactions of LiPSs (Fig. 2f.34
To compare the battery response current, indicative of the internal lithium-ion diffusion rate, after the catalyst improvement, we assembled symmetric cells and conducted electrochemical testing. The CV curves of CoNiCuMnMo-HEA@rGO/PP and rGO/PP range from 0.1 mV s−1 to 0.6 mV s−1 (Fig. S8†). Notably, plotting the anodic and cathodic peak currents of these cells at varying scan rates against the square root of the scan rate revealed a steeper slope in the Randles–Sevcik plots for cells incorporating CoNiCuMnMo-HEA.35 Evidently, the absolute slope of CoNiCuMnMo-HEA@rGO/PP was significantly higher than that of rGO/PP (Fig. 2g), strongly indicating that HEA nanoparticles significantly enhance the diffusion behavior of Li+.
The kinetics of the liquid-to-solid transformation of LiPSs/Li2S and the ion transfer capabilities were further supported by the Li2S nucleation and dissociation experiments.36 The nucleation and dissociation processes in the LiPS redox reactions were monitored using chronopotentiometry. Fig. 2h and i depict the constant potential current distribution on the surfaces of CoNiCuMnMo-HEA@rGO and rGO precipitated with Li2S, respectively. Prior to nucleation, CR2025 coin cells were assembled using CoNiCuMnMo-HEA@rGO and rGO as cathodes and 20 μL Li2S8 as the electrolyte. In detail, the nucleation time of Li2S on CoNiCuMnMo-HEA@rGO (1074 s) was earlier than on rGO (1821 s), which may be attributed to the synergistic interface of CoNiCuMnMo-HEA and rGO with high absorption capacity for LiPSs and rapid reaction kinetics, facilitating the nucleation of Li2S. Moreover, the precipitation capacity of Li2S was calculated based on Faraday's law. The precipitation capacity of the CoNiCuMnMo-HEA@rGO electrode (196.5 mA h g−1) is significantly higher than that of the rGO electrode (147.6 mA h g−1), indicating that CoNiCuMnMo-HEA@rGO can significantly promote the liquid-to-liquid and liquid-to-solid transformations of LiPSs compared to the rGO matrix.
Furthermore, to visually demonstrate the interaction between CoNiCuMnMo-HEA@rGO and LiPSs, we conducted an adsorption experiment with Li2S6. As evidenced by the optical images, over time, the initially pale yellow Li2S6 solution became transparent in the presence of CoNiCuMnMo-HEA@rGO, whereas the pale yellow color remained clearly observable in the rGO (see the inset of Fig. S9a†). This phenomenon indicates that CoNiCuMnMo-HEA@rGO possesses a stronger physical adsorption capacity for LiPSs compared to the original rGO. Additionally, ultraviolet-visible (UV-vis) spectroscopy analysis of the adsorbed solution (Fig. S9a†) shows a significant decrease in the intensity of the characteristic peak at approximately 260 nm, corresponding to S62−, indicating its strong physical adsorption capability for LiPSs, consistent with the aforementioned analyses. Subsequently, the strong interaction between Li2S6 and CoNiCuMnMo-HEA@rGO was further verified by XPS (Fig. S9b–f†). It was observed that after adsorbing Li2S6, the peaks of Co 2p and Ni 2p shifted towards higher binding energies, indicating an increase in the electron density at the Co and Ni centers. This is due to the chemical interaction of Co and Ni with Li2S6.15,37 In the XPS spectrum of Mo, the characteristic peak of Mo6+ decreased after the adsorption experiment, indicating a strong chemical reaction between Mo6+ and Li2S6. Additionally, the position of the Mo 3d peak shifted slightly towards higher binding energy, suggesting a strong electronic interaction between Mo and LiPSs.38 The binding energies corresponding to Mn, Cu, and their respective ions (Mn2+ and Cu2+) were significantly higher than those before adsorption, indicating interactions between Mn, Cu, and LiPSs. These results further confirm that the polar adsorption between CoNiCuMnMo-HEA@rGO and LiPSs leads to intense chemical anchoring and effectively captures LiPSs. In addition, Fig. S10† shows the ex situ SEM and maps of S elements on the electrode containing different modified separators with the distribution. For the electrode containing CoNiCuMnMo-HEA@rGO/PP, the S element is uniformly distributed on the rGO surface. In Fig. S10b,† for the electrode containing rGO/PP, the sulfur content almost disappears in the middle of the discharge. The results show that CoNiCuMnMo-HEA@rGO has a good adsorption capacity for polysulfide and can well prevent LiPSs from passing through the separator.
The catalytic performance of the CoNiCuMnMo-HEA@rGO catalyst for the reverse reaction in Li–S batteries, specifically its oxidation catalytic effect on Li2S, was investigated using UV-vis and XPS techniques to analyze the interaction between Li2S and CoNiCuMnMo-HEA@rGO. Compared with Li2S and rGO (inset of Fig. 3a), the supernatant of the reaction between the insoluble Li2S and CoNiCuMnMo-HEA@rGO exhibited a significant color change (from colorless to yellowish-green), indicating the formation of LiPSs39 (the black object at the bottom is the magnetic stirrer). From Fig. 3a, the presence of LiPSs was further confirmed by UV-vis spectroscopy, which showed characteristic peaks at approximately 260 nm for S62− and at 375 nm for S42−.
It was observed that after the reaction between CoNiCuMnMo-HEA@rGO and Li2S, the characteristic peaks of metallic Co, Ni, Mo, Mn, and Cu in the XPS spectra shifted. Specifically, the Co2+ 2p1/2 peak shifted to a higher binding energy by 0.4 eV, while the Co0 2p3/2 and Co0 2p1/2 peaks shifted to a lower binding energy by 0.7 and 0.9 eV (Fig. 3b), respectively. For Cu, the metallic oxidation states Cu2+ 2p3/2 and Cu2+ 2p1/2 shifted to higher binding energies by 0.3 and 0.35 eV (Fig. 3c), respectively. From Fig. 3d, Mn exhibited three valence states, and all six characteristic peaks shifted to various degrees. Among them, Mn0 2p3/2 and Mn0 2p1/2 shifted to higher binding energies by 0.35 and 0.85 eV, while Mn2+ 2p3/2, Mn2+ 2p1/2, Mn3+ 2p3/2, and Mn3+ 2p1/2 shifted by 0.3, 0.25, 0.6, and 1.2 eV, respectively. In Fig. 3e, the Mo6+ 3d5/2 and Mo6+ 3d3/2 peaks also shifted to higher binding energies by 0.4 and 0.5 eV, respectively. The oxidation states Ni 2p3/2 and Ni 2p1/2 orbitals shifted by 0.35 and 0.25 eV (Fig. 3f), respectively. The areas of the oxidized metal states all increased, indicating a reduction in the content of these states after reacting with polysulfides. This demonstrates that CoNiCuMnMo-HEA@rGO form a strong interaction with polysulfides during the oxidation of Li2S, effectively promoting the formation of polysulfides and accelerating the redox reaction of Li2S, thereby enhancing the performance of Li–S batteries.15
To investigate the superiority of CoNiCuMnMo-HEA@rGO in the Li2S dissociation process, a similar battery configuration was used for constant current discharge to 1.7 V (to ensure that LiPSs were almost completely converted to Li2S), followed by constant potential charging at 2.35 V. The current response of the CoNiCuMnMo-HEA@rGO electrode (Fig. 3g) was higher than that of the rGO electrode (Fig. 3h), indicating that CoNiCuMnMo-HEA@rGO can effectively reduce the overpotential during the reverse transformation of Li2S, facilitating the decomposition of Li2S. Additionally, compared to rGO (265.6 mA h g−1), the CoNiCuMnMo-HEA@rGO electrode (300.4 mA h g−1) exhibited a higher dissociation capacity, demonstrating that CoNiCuMnMo-HEA@rGO promotes the dissociation of Li2S. These results confirmed that the existence of CoNiCuMnMo-HEA enhanced the reaction kinetics of the transformation of Li2S into polysulfide/sulfur (Fig. 3i).
Based on our experimental characterization, we utilized a simplified CoNiCuMnMo-HEA particle model to perform density functional theory (DFT) computational simulations, further elucidating the unique advantages of CoNiCuMnMo-HEA in catalyzing the SRR at the atomic level. To this end, we simulated the adsorption behavior of cyclic S8 and Li2Sn (1 ≤ n ≤ 8) on CoNiCuMnMo-HEA@rGO and rGO (Fig. S11†), respectively. Compared to rGO, the Li2Sn (1 ≤ n ≤ 8) adsorbed by CoNiCuMnMo-HEA@rGO exhibited slight or obvious deformation, primarily due to the formation of chemical bonds between the S atoms and the metal atoms. This is also important proof that CoNiCuMnMo-HEA can effectively adsorb LiPSs and suppress shuttle effects. For the multi-phase catalytic conversion of LiPSs, the tight adsorption of active species on catalytically active sites is a prerequisite.40 As can be seen from Fig. 4a, the binding energies of various LiPSs on the surface of CoNiCuMnMo-HEA@rGO based on DFT calculations are much stronger than those on rGO, and for soluble Li2Sn, CoNiCuMnMo-HEA@rGO exhibits a much stronger binding energy, which is consistent with the results of the adsorption experiments.
In addition, we further considered the transformation pathways of intermediates such as *S8, *Li2S8, *Li2S6, *Li2S4, *Li2S2, and *Li2S to evaluate the catalytic performance of CoNiCuMnMo-HEA@rGO (Fig. 4b). Compared to rGO, the presence of CoNiCuMnMo-HEA@rGO led to a significant reduction in the overall free energy of LiPSs and greatly decreased the energy barriers for each step of their transformation, indicating that CoNiCuMnMo-HEA@rGO accelerates the catalytic conversion of LiPSs. This demonstrates that the multi-metal synergistic effect of CoNiCuMnMo-HEA is the core to accelerate the catalytic conversion of LiPSs.41 Additionally, we calculated the density of states (DOS) of Li2S6 on different substrates and found that compared to other single metal substrates, CoNiCuMnMo-HEA has more dense energy bands, which are more conducive to electron motion and transition, indicating a stronger interaction between Li2S6 and CoNiCuMnMo-HEA. This can be attributed to the renowned cocktail effect of CoNiCuMnMo-HEA (Fig. 4c). These analyses indicate that CoNiCuMnMo-HEA, due to the multi-metal synergistic catalysis resulting from the cocktail effect, enhances the adsorption of LiPSs, accelerates redox reactions, and thereby effectively suppresses shuttle effects.
Fig. 5a presents the CV curve of the CoNiCuMnMo-HEA@rGO/PP battery obtained at a scanning rate of 0.1 mV s−1, from which it can be clearly observed that the cathodic and anodic peak curves in the first three cycles are almost identical, indicating that the catalytic effect of CoNiCuMnMo-HEA@rGO/PP exhibits good chemical reversibility and stability. Fig. 5b shows the charge–discharge curve of the battery at a rate of 0.2 C. The discharge capacity at the second discharge plateau of the CoNiCuMnMo-HEA@rGO/PP battery significantly increased, indicating that the conversion efficiency from LiPSs (Li2Sn, 4 ≤ n ≤ 8) to Li2S/Li2S2 has been significantly improved. Notably, the potential difference (ΔE1) between the second charge and discharge curves of the CoNiCuMnMo-HEA@rGO/PP is relatively smaller than that of the rGO/PP, indicating that the battery with a CoNiCuMnMo-HEA functional separator has lower polarization and better reversibility. Subsequently, at different current rates from 0.2 to 5 C, the HEA@rGO/PP battery exhibited a higher discharge specific capacity (Fig. 5c). The initial discharge specific capacity of HEA@rGO/PP was 1512.4 mA h g−1. When the current density increased to 5 C, the discharge capacity stabilized at 570.2 mA h g−1. When the current density was restored to 0.2 C, the battery capacity rebounded to 796.5 mA h g−1, demonstrating good chemical reversibility and a high utilization rate of the active material (Table S1†). The constant current charge–discharge voltage curves of CoNiCuMnMo-HEA@rGO/PP and rGO/PP batteries between 1.5 and 3.0 V showed two discharge plateau regions (Fig. S12†), corresponding to the conversion of cyclic S8 to LiPSs, which then transform into solid Li2S2/Li2S. In addition, the thickness of catalytic intercalation has a significant influence on the performance (Fig. S13†), and the best performance is obtained when the intercalation thickness is about 20 μm.
As shown in Fig. 5d, CoNiCuMnMo-HEA@rGO/PP shows excellent rate performance compared with HEA-based lithium–sulfur batteries reported in the literature.24,42–50Fig. 5e shows the long-term cycle performance tested at 3 C with different interlayers when the sulfur loading is 1.0 mg cm−2. CoNiCuMnMo-HEA@rGO/PP exhibits experienced 1100 cycles, and the decay rate of each cycle is only 0.055%, and the Coulomb efficiency exceeds 91.94%, which is significantly better than that of rGO/PP. As shown in Fig. S14,† compared to the pristine lithium metal with a metallic luster, the surface topography of the lithium metal electrodes becomes rough as the current density increases. Moreover, after cycling for 100 cycles at 3C, many black spots appear on the surface of the lithium metal electrodes, indicating the formation of ‘dead lithium’, which has been believed to be one of the most important reasons for the decline of coulombic efficiency. The above excellent long-cycle performance is attributed to the high catalytic activity and structural stability of the close-packed catalyst surface with a face-centered cubic structure. Tests were conducted on batteries with high sulfur loadings (Fig. 5f). The initial area specific capacity of the HEA@rGO/PP battery at 0.1 C reached 2.63 mA h cm−2, and the capacity remained at 2 mA h cm−2 after 100 cycles. This is mainly due to the fact that CoNiCuMnMo-HEA@rGO/PP can accelerate the bidirectional reaction power of polysulfide/sulfide, thus effectively inhibiting the shuttle effect and improving the battery performance.
Based on the bidirectional catalytic conversion of polysulfides/sulfides by CoNiCuMnMo-HEA, it can effectively inhibit the shuttle effect of polysulfides and reduce the generation of lithium dendrites and dead lithium.51 The formation of Li2S on the anode side is a result of the slow shuttling of LiPSs to the lithium side, which reacts with lithium to form ‘dead sulfur’ on the one hand, leading to irreversible capacity degradation, and generating ‘dead lithium’ and electrolyte loss on the other hand, thus resulting in the polarization increasing and performance degradation of Li–S batteries. The batteries after cycling were subjected to EIS tests to monitor the trend of external impedance. As shown in Fig. S15,† according to the equivalent circuit model, Rct is the charge transfer resistance resulting from the reaction kinetics of the conversion of long-chain LiPSs to short-chain Li2S2/Li2S on the cathode corresponding to the semicircle in the mid- to high-frequency region. RSEI is the solid electrolyte interface (SEI) resistance. The post-cycling EIS shows two semicircles in the high and mid-frequency regions. The formation of RSEI indicates the resistance of Li+ in the SEI film, confirming the generation of a stable SEI at the electrolyte interface in contact with lithium metal. The presence of the SEI film prevents the continuous electron transfer between the Li anode and the LiPSs in the electrolyte, thus stabilizing the surface structure of the Li-anode. The total impedance of the Li–S battery decreased after cycling, which could be attributed to the redistribution of sulfur in the cathode. The EIS results confirmed that CoNiCuMnMo-HEA@rGO could induce the uniform generation of the SEI layer, which reduces the ‘dead lithium’, improves the utilization rate of the active lithium, and ultimately prolongs the cycle life of the battery. To verify this hypothesis, the battery was disassembled after 100 cycles, and the corrosion degree of the lithium anode was observed using scanning electron microscopy. The anode corresponding to rGO/PP suffered severe corrosion compared to CoNiCuMnMo-HEA@rGO/PP (Fig. S12†). These analyses indicate that introducing CoNiCuMnMo-HEA into the intercalated separator can effectively accelerate the catalytic conversion of LiPSs, inhibit the formation of lithium dendrites, and significantly improve the performance of Li–S batteries.
Footnotes |
† Electronic supplementary information (ESI) available: Detailed characteristics of the catalysts and additional reduction results. See DOI: https://doi.org/10.1039/d4qi01434k |
‡ These authors contributed equally to this work. |
This journal is © the Partner Organisations 2024 |