MOF-derived Co–Mo bimetallic heterostructures for the selective trapping and conversion of polysulfides in lithium–sulfur batteries

Rongmei Zhu *a, Yuxuan Jiang a, Bingxin Sun a, Wang Zhang ab and Huan Pang *ac
aSchool of Chemistry and Chemical Engineering, Yangzhou University, Yangzhou, Jiangsu 225009, P. R. China. E-mail: rmzhu@yzu.edu.cn; panghuan@yzu.edu.cn; huanpangchem@hotmail.com
bResearch Institute for Convergence Science, Seoul National University, Seoul, 08826, Republic of Korea
cState Key Laboratory of Coordination Chemistry, Nanjing University, Nanjing, 210093, P. R. China

Received 18th May 2024 , Accepted 9th September 2024

First published on 23rd October 2024


Abstract

Lithium–sulfur batteries (LSBs) are promising energy storage systems, but their practical application is hindered by the polysulfide shuttle effect and slow redox kinetics. To address these challenges, we constructed ZIF-67@CoSx/MoO3 with a core–shell structure and CoSx/MoO3 with a hollow structure as separator-modified materials for LSBs by varying the degree of sulfidation of ZIF-67. The high intrinsic conductivity of CoSx facilitated ion transfer between the cathode and separator. Additionally, the introduction of MoO3 formed a heterogeneous structure with CoSx that enhanced the adsorption of LiPSs. Via in situ UV-vis and electrochemical impedance spectroscopy testing, we demonstrated the preferred selective trapping and conversion of LiPSs by CoSx/MoO3. As a result of the synergistic effect of the bimetallic heterogeneous structure, the modified LSB exhibited excellent cycling stability, with a capacity decay rate of only 0.041% after 500 cycles at 1C. Moreover, it achieved a high discharge capacity of 632 mA h g−1 at 2C. This work provides a novel concept for MOF-derived heterogeneous structures to be applied in high-performance LSBs.


Introduction

The growing demand in the emerging electric vehicle and renewable energy storage markets has prompted the search for next-generation energy storage devices.1–3 Rechargeable LSBs are considered a promising option due to their high theoretical energy density and low cost. However, the commercialization of these batteries faces challenges such as low sulfur utilization, short cycle life, and low charging and discharging efficiency. These challenges are primarily caused by the poor electrical conductivity of sulfur and its discharge products, as well as the dissolution of intermediate lithium polysulfides (LiPSs) and the shuttle effect.4–6

To address these challenges, researchers have investigated various separator-modified materials to physically and chemically adsorb LiPSs. Transition metal compounds, particularly oxides, exhibit strong adsorption ability,7,8 effectively preventing polysulfide shuttling. However, transition metal compounds typically present poor electronic conductivity and limited catalytically active sites, hindering the cycling process.

Recently, the development of metal–organic frameworks (MOFs) and their derivatives has garnered attention in the modification of lithium–sulfur battery separators9–11 due to their unique porous structure and strong chemisorption properties for LiPSs. Additionally, MOFs have been utilized as templates to construct transition metal compounds, providing various reactive sites for polysulfide adsorption and facilitating their catalytic conversion.12 Although MOF-derived sulfides exhibit high conductivity and catalytic ability, their polysulfide adsorption capability is relatively weak.13 Therefore, the construction of bimetallic transition metal sulfide/oxide heterostructures has been proposed as an effective approach for enhancing the cycling performance of LSBs.14 CoSx can chemically bond with low-order LiPSs, reducing the energy barrier in the reaction. Co9S8 exhibits good catalytic activity and, in addition to its powerful adsorption ability, can accelerate the redox reactions of LiPSs.15 Based on these findings, Su et al. employed a simple electrospinning method to prepare a core–shell heterostructure anchored on nanocarbon fibers. This heterostructure combines the synergistic effects of chemical adsorption and electrochemical catalysis, resulting in improved sulfur utilization of high sulfur-loaded electrodes.16 At the same time, density functional theory calculations further indicate that the binding strength between MoO3 and LiPSs is strong, making it an ideal host for anchoring LiPSs and improving cycling stability.17

Therefore, in this study, we utilized ZIF-67 as a template to fabricate two separator-modified materials, ZIF-67@CoSx/MoO3 and CoSx/MoO3, and investigated their performance in LSBs. The core–shell structured ZIF-67@CoSx/MoO3 was obtained by partially sulfurizing ZIF-67 and introducing the Mo element afterwards. Nevertheless, the hollow structured CoSx/MoO3 was achieved by complete sulfidation of ZIF-67, resulting in significantly improved performance as a separator in LSBs. In situ electrochemical impedance tests and in situ UV-vis confirmed that compared with ZIF-67@CoSx/MoO3, the hollow CoSx/MoO3 delivered much higher conductivity and can selectively capture LiPSs in the electrolyte and promote their conversion to insoluble LiPSs to a greater extent.

Results and discussion

Fig. 1a presents the preparation of ZIF-67@CoSx/MoO3 and CoSx/MoO3. ZIF-67 with a particle size of about 500 nm was synthesized in one step using a solvent method, and the scanning electron microscopy (SEM) image is shown in Fig. S1. Then, by controlling the temperature and time of the hydrothermal reaction, precursor 1 (PR1) and precursor 2 (PR2) with different degrees of sulfidation were obtained upon mixing ZIF-67 with an ethanol solution of thiourea acetamide (TAA) (their SEM images are shown in Fig. S2a and b). The X-ray diffraction (XRD) spectra of PR1 and PR2 are shown in Fig. S3, with the characteristic peaks of ZIF-67 in PR1 weakened. No obvious diffraction peaks belonging to sulfides were observed, indicating that the obtained PR1 and PR2 may be amorphous.18 In the transmission electron microscopy (TEM) image (Fig. S4a), the edges of PR1 appeared rough, and Fig. S4b shows the hollow structure of PR2. Raman spectroscopy analysis further revealed that the peaks at 670 cm−1 corresponded to the vibration of Co–S in Co9S8 (Fig. S5a).19 Mixing PR1 and PR2 with NaMoO4·2H2O aqueous solution separately resulted in the formation of ZIF-67@CoSx/MoO3 and CoSx/MoO3, respectively; they all show a spherical shape and are coated with nanosheets on the outer layer (Fig. S2c and d). Similar to PR1 and PR2, the vibration of Co–S was observed in Fig. S5b, and the peaks at 817 cm−1 and 941 cm−1 were attributed to the asymmetric vibration of O–Mo–O originating from MoO3.20 Fourier transform infrared (FTIR) spectroscopy further confirmed the formation of heterostructures. The presence of Co–S (1084 cm−1) and Mo–O bonds (564 cm−1) can be attributed to CoSx and MoO3 (Fig. S5c).21,22 A high-resolution transmission electron microscopy (HRTEM) image (Fig. 1b) revealed the core–shell structure of ZIF-67@CoSx/MoO3. The energy-dispersive X-ray spectroscopy (EDX) elemental mapping (Fig. 1c and Fig. S6) confirmed the successful introduction of Co, Mo, S, and O, while the lattice fringes in Fig. 1d indicated the (440) and (400) planes of Co3S4 corresponding to 0.162 nm and 0.241 nm, respectively, and 0.223 nm was attributed to the (150) plane of MoO3.23Fig. 1e and j show the selected area electron diffraction (SAED) patterns of ZIF-67@CoSx/MoO3 and CoSx/MoO3, respectively. The Brunauer–Emmett–Teller (BET) surface area data (Fig. 1f) indicated the uniform microporous structure inside ZIF-67@CoSx/MoO3. In comparison, CoSx/MoO3 exhibited a hollow structure (Fig. 1g), and the element distribution in Fig. 1h showed the abundant distribution of Co, Mo, S, and O at the hexagonal edges. The lattice fringe analysis in Fig. 1i revealed the (422) plane of Co9S8 (0.208 nm) and the (400) plane of Co3S4 (0.236 nm). Furthermore, the XRD patterns of ZIF-67@CoSx/MoO3 and CoSx/MoO3 further confirmed the formation of the Co–Mo bimetallic heterostructure. As shown in Fig. S7a, ZIF-67@CoSx/MoO3 retains the major ZIF-67 diffraction peaks, indicating a low degree of sulfidation. Nevertheless, diffraction peaks at around 26° and 37° were attributed to MoO3 (PDF#75-0912), and diffraction peaks associated with Co9S8 (PDF#02-1459) and Co3S4 (PDF#47-1738) occurred mainly at around 30°, proving the presence of small amounts of CoSx and MoO3. Compared with ZIF-67@CoSx/MoO3, CoSx/MoO3 shows more pronounced diffraction peaks related to sulfide in Fig. S7b, which is related to its deeper sulfidation. The BET data in Fig. 1k showed that CoSx/MoO3 possessed an uneven mesoporous structure, which is attributed to the hollow structure. The abundant mesopores in CoSx/MoO3 can not only facilitate the diffusion of electrolytes and the migration of lithium ions but also suppress the loss of soluble LiPSs through confinement.24
image file: d4qi01249f-f1.tif
Fig. 1 (a) Schematic of the preparation of ZIF-67@CoSx/MoO3 and CoSx/MoO3. Characterization of ZIF-67@CoSx/MoO3 and CoSx/MoO3: (b and g) TEM images, (c and h) elemental mapping images, (d and i) HRTEM images, (e and j) SAED patterns and (f and k) N2 adsorption/desorption isotherms and pore diameter distributions.

In order to examine the uniformity and firmness of the sample loading on the separators, we conducted microscopic characterization and folding experiments. Fig. 2a–c and their insets show the SEM images of bare polypropylene (PP), ZIF-67@CoSx/MoO3/PP, and CoSx/MoO3/PP, respectively. The PP separator had a pore size of about 100 nm (Fig. S8), which easily allowed the shuttle of LiPSs to the negative electrode, causing irreversible capacity loss. Coating ZIF-67@CoSx/MoO3 or CoSx/MoO3 on PP can facilitate the physical blocking of LiPSs. In addition, the highly conductive CoSx can decrease the polarization inside the cell due to the increased thickness of the separator.25 The cross-sectional SEM image (Fig. 2d) showed a uniform functional coating of CoSx/MoO3 on PP. Additionally, Fig. 2e and f show optical photographs of CoSx/MoO3/PP and folded–unfolded different separators. The unfolded CoSx/MoO3/PP did not exhibit obvious detachment, demonstrating good adhesion and flexibility of the coating. Therefore, thermal stability is significant for the safety of batteries. We conducted heating tests on different separators, and it is clear that the modified separators have better thermal stability and resistance to thermal shrinkage than PP (Fig. S9). In the electrolyte wetting test, the modified separators also achieved better results, which facilitated the full wetting of the electrolyte (Fig. S10).


image file: d4qi01249f-f2.tif
Fig. 2 SEM images of (a) bare PP, (b) ZIF-67@CoSx/MoO3/PP, and (c) CoSx/MoO3/PP. (d) Cross-sectional SEM image of CoSx/MoO3/PP. (e) Photographs of CoSx/MoO3/PP. (f) Photographs of bare PP, ZIF-67@CoSx/MoO3/PP and CoSx/MoO3/PP under the same mechanical stress.

To investigate the adsorption capacity of the different samples for LiPSs, we performed static adsorption experiments of PR1, PR2, ZIF-67@CoSx/MoO3, and CoSx/MoO3 in a Li2S4 solution with equal mass (Fig. 3a). Benefiting from the synergistic effect of CoSx and MoO3, compared to the color of the liquid in the other bottles, the solution in bottle 5 became significantly lighter after 6 h of standing, and completed the adsorption of LiPSs in 24 h. The decrease in the absorption peak at 420 nm in the UV-Vis absorption spectra (Fig. 3b) further confirmed the efficient adsorption of LiPSs by CoSx/MoO3. However, the core of ZIF-67@CoSx/MoO3 mainly consisted of ZIF-67, which could limit the selective capture of LiPSs. Therefore, after 24 h, the Li2S4 peak in the corresponding solution was still evident. To demonstrate the inhibition of the shuttle effect more intuitively by the modified separator, we placed ZIF-67@CoSx/MoO3/PP and CoSx/MoO3/PP into H-type cells. As shown in Fig. 3c, LiPSs shuttle extensively into the transparent solution in the right compartment of the H-type cell with ZIF-67@CoSx/MoO3/PP in only 9 h. In contrast, after 18 h of standing, only a small amount of LiPSs entered the right compartment of the control group, further illustrating the effective suppression of the shuttle effect introduced by CoSx and MoO3. X-ray photoelectron spectroscopy (XPS) analysis was further conducted to study the chemical states of ZIF-67@CoSx/MoO3 and CoSx/MoO3 before and after LiPS adsorption and their full spectra are displayed in Fig. S11 and S12. As shown in Fig. 3d and g, the SOx species that appeared at high binding energy positions can effectively alleviate the shuttle effect, as reported in other articles.26,27 The Co 2p3/2 spectrum in Fig. 3e showed fitting peaks at 781.31 eV and 784.52 eV, and the spectrum in Fig. 3h showed fitting peaks at 781.72 eV and 784.01 eV, which can be attributed to Co2+ and Co3+, respectively.28,29 The Mo 3d orbitals exhibited two strong peaks at 232.73 eV and 235.81 eV in Fig. 3f, as well as 232.69 eV and 235.75 eV in Fig. 3h, which can be assigned to Mo6+.17 As can be seen visibly, the binding energies of the metal ions in ZIF-67@CoSx/MoO3 and CoSx/MoO3 shifted after the adsorption of Li2S4. The binding energy of the Co3+ fitting peak in CoSx/MoO3 decreased by 2.36 eV that is much larger than that in ZIF-67@CoSx/MoO3, indicating its stronger chemical affinity with Li2S4 and higher adsorption capacity for LiPSs.30 Additionally, the binding energy of the Mo6+ peak also shifted by 0.96 eV, indicating the involvement of Mo in the anchoring process of LiPSs. Therefore, the XPS analysis after adsorption confirmed that the hollow-structured CoSx/MoO3, with the assistance of bimetallic synergy and sulfur-rich sites, can selectively capture more soluble LiPSs and suppress the shuttle effect more efficiently. To investigate the advantages of the hollow bimetallic heterostructure in LSBs, we performed cyclic voltammetry (CV) tests on PR1/PP, PR2/PP, ZIF-67@CoSx/MoO3/PP, and CoSx/MoO3/PP (Fig. 4a and Fig. S13). All samples exhibited typical redox peaks, but the peak currents of PR1 and PR2, which contained only single metal Co, were smaller than those of samples with the Co–Mo bimetallic heterostructure. Additionally, PR2 with a higher sulfidation degree showed better cycling stability than PR1. Meanwhile, CoSx/MoO3/PP demonstrated the best redox reversibility, showing almost no shift in the oxidation–reduction peaks after 3 cycles at a scan rate of 0.1 mV s−1. Furthermore, compared to the reduction peak at around 2 V and the oxidation peak at around 2.4 V for ZIF-67@CoSx/MoO3/PP, the reduction peak of CoSx/MoO3/PP shifted positively, and the oxidation peak shifted negatively, indicating less internal polarization in the battery and faster reaction kinetics.14


image file: d4qi01249f-f3.tif
Fig. 3 (a) Digital pictures of Li2S4 solutions before and after adding PR1, PR2, ZIF-67@CoSx/MoO3, and CoSx/MoO3 and (b) corresponding UV-Vis absorption spectra. (c) Li2S4 penetration in H-type cells based on different separators. High-resolution XPS spectra of (d) S (e) Co, and (f) Mo in ZIF-67@CoSx/MoO3 and ZIF-67@CoSx/MoO3/Li2S4. (g–i) High-resolution XPS spectra of (g) S (h) Co, and (i) Mo in CoSx/MoO3 and CoSx/MoO3/Li2S4.

image file: d4qi01249f-f4.tif
Fig. 4 (a) 3-turn CV curves of LSBs with CoSx/MoO3/PP. (b) CV curves of LSBs with CoSx/MoO3/PP at different scan rates. (c) The linear relationship between Ip and ν1/2 of LSBs with CoSx/MoO3/PP. (d and e) Extracted parts of the CV curves in (a) between 1.7 and 2.8 V and Tafel plots of Peak B and Peak C (inset). (f) Rate performance of LSBs with different separators. (g) Voltage gap between the charge/discharge platforms of the LSBs with different separators at 0.2C. (h) Cycling performance of different separators at 1C.

The higher polarization within ZIF-67@CoSx/MoO3/PP may be due to the low conductivity of ZIF-67 itself and the lower degree of sulfidation, which lead to limited catalytic sites and weak catalytic activity during the reaction. On the other hand, when comparing PR2 with CoSx/MoO3/PP, it can also be observed that the introduction of Mo enhances the conductivity and catalytic activity of CoSx/MoO3, resulting in the highest peak current and the smallest polarization for CoSx/MoO3/PP. CV tests were performed on the samples at different scan rates to further verify the improvement in battery stability achieved by the introduction of Mo (Fig. 4b and Fig. S14). As the scanning rate increases, the CV curve of the LSB with CoSx/MoO3/PP still shows clear redox peaks and minimal polarization, indicating that CoSx/MoO3 can effectively promote the redox reaction in LSBs. Additionally, based on the CV results, we calculated the relationship between the peak current of different oxidation–reduction peaks (Peak A–C) and the corresponding scan rates for ZIF-67@CoSx/MoO3/PP and CoSx/MoO3/PP (Fig. 4c and Fig. S15). The slope values of the fitted curves are related to the lithium-ion diffusion rate in the battery, and CoSx/MoO3/PP exhibited a steeper slope, indicating a faster lithium-ion diffusion rate.31 The flatter slope of ZIF-67@CoSx/MoO3/PP may be due to the lower degree of sulfidation and the insufficient ability to selectively adsorb and convert LiPSs, resulting in the dissolution of soluble LiPSs into the electrolyte during discharge, which increases the electrolyte concentration and increases resistance to lithium-ion transfer.32 The introduction of Mo and the deeper sulfidation can enhance the electron transfer rate of CoSx/MoO3 and greatly enhance the selective adsorption and conversion ability of CoSx/MoO3 for LiPSs.

To further examine the catalytic behavior of the CoSx/MoO3 modified separator on LiPSs, we extracted parts of the CV curves from Fig. 4a and Fig. S13 between 1.7 and 2.8 V, as demonstrated in Fig. 4d and e. The Tafel slope of the battery equipped with a separator modified by ZIF-67@CoSx/MoO3 is 50 mV dec−1 for the reduction reaction, which is higher than that observed with a separator modified by CoSx/MoO3 (31 mV dec−1). Meanwhile, the Tafel slope of the separator modified by ZIF-67@CoSx/MoO3 is observed to be 123 mV dec−1 in the oxidation reaction, which is also higher than that of the CoSx/MoO3 modified separator (105 mV dec−1). A smaller Tafel slope indicates a stronger charge transfer ability of the catalyst in the catalytic process. The above results suggest that the CoSx/MoO3 catalyst shows excellent charge transfer ability, which can accelerate the kinetics of polysulfide conversion. Therefore, it improves the discharge capacity and long cycle life of LSBs.

Therefore, according to the Randles–Sevcik equation,

image file: d4qi01249f-t1.tif

In which Ip, DLi+, and ν represent the peak current, lithium-ion diffusion coefficient, and scan rate, respectively, while n, A, and CLi+ represent the number of electrons involved in the electrochemical reaction, the effective electrode area, and the lithium-ion concentration in the electrolyte. The lithium-ion diffusion rate can be observed through the linear correlation of the peak current and the square root of the sweep rate. The lithium-ion diffusion coefficients are obtained from the Randles–Sevcik equation (Fig. S16), and DLi+ on CoSx/MoO3 is always much higher than that on ZIF-67@CoSx/MoO3 and PP.

In addition, we performed galvanostatic charge–discharge (GCD) tests at different currents (0.1–2C) to examine the discharge performance of batteries with PR1/PP, PR2/PP, ZIF-67@CoSx/MoO3/PP, and CoSx/MoO3/PP (Fig. 4f). Due to the lack of selective capture of LiPSs and catalytic conversion ability, the capacity of PP decreases significantly at 1C current. After partial sulfidation of ZIF-67 and the introduction of Mo, the batteries with modified separators showed improved discharge performance at high currents. Fortunately, CoSx/MoO3/PP exhibited the best rate performance. LSBs with CoSx/MoO3/PP show a high discharge capacity of 632 mA h g−1 at 2C rate. GCD testing is also a method used to measure the cycling stability of batteries. We tested the polarization ΔE of different samples during 100 cycles at 0.2C (Fig. 4g). At the beginning of the cycling, PP and CoSx/MoO3/PP had the highest and lowest polarization, respectively. With an increasing number of cycles, the polarization of PR1 and ZIF-67@CoSx/MoO3/PP increased significantly, which was related to the low conductivity of ZIF-67 itself and the lower degree of sulfidation, resulting in the coverage of catalytic sites. Thanks to the deeper sulfidation, CoSx/MoO3/PP maintained the smallest polarization after 100 cycles. Meanwhile, we disassembled the batteries after cycling to observe PP, ZIF-67@CoSx/MoO3/PP, and CoSx/MoO3/PP. Many pores of PP were blocked due to the lack of physical and chemical adsorption of LiPSs (Fig. S17a). Partial detachment of the coating was observed in ZIF-67@CoSx/MoO3/PP after long cycles in Fig. S17b, possibly due to the large volume expansion of the positive electrode during charge–discharge. In contrast, CoSx/MoO3/PP maintained its coating integrity after long cycles (Fig. S17c), demonstrating the best cycling stability. After disassembling the battery after cycling, we found that the negative electrode surface of the battery equipped with the CoSx/MoO3 modified separator was smoother (Fig. S18), which may be due to the ability of CoSx/MoO3 to quickly and selectively adsorb LiPSs, which greatly improves the “shuttle effect” of the battery.

As shown in Fig. 4h, all modified separators showed higher initial capacity than bare PP in the long cycling test at 1C. Furthermore, compared to other samples, CoSx/MoO3 showed the highest capacity retention (81.3%) and a capacity decay rate of 0.041% after 500 cycles. During cycling, the capacity of the battery with the CoSx/MoO3 modified separator did not monotonically decrease, further demonstrating the abundant active sites and improved electron transfer efficiency of CoSx/MoO3, which can accelerate the conversion of LiPSs and promote the reuse of dead sulfur, greatly improving battery life and cycling stability.33 Compared with other transition metal compounds used as separators in LSBs (Table S1), CoSx/MoO3 exhibited competitive cycling stability and rate capability.

To further investigate the internal resistance of batteries with ZIF-67@CoSx/MoO3/PP and CoSx/MoO3/PP during charge and discharge, galvanostatic intermittent titration (GITT) was performed at 0.1C. The polarization during cycling can be quantified by the relative magnitude of ΔiR in GITT, which is calculated as follows:

ΔiR(Ω) = |ΔVQOCV–CCV|/I
where ΔVQOCV–CCV is the voltage difference between the quiescent open circuit voltage (QOCV) and closed-circuit voltage (CCV) and I is the applied current. As shown in Fig. 5a and Fig. S19, CoSx/MoO3/PP exhibited a smaller ΔiR than ZIF-67@CoSx/MoO3/PP at the Li2S nucleation and activation points, indicating that CoSx/MoO3 can better reduce the hindrance of LiPS conversion.34


image file: d4qi01249f-f5.tif
Fig. 5 (a) GITT voltage profiles of CoSx/MoO3/PP. (b and c) In situ UV/Vis spectra of ZIF-67@CoSx/MoO3/Li2S4 and CoSx/MoO3/Li2S4. (d) In situ EIS spectra of the battery with a CoSx/MoO3 modified separator. (e) DRT calculated from EIS measurements of the battery with CoSx/MoO3/PP at different voltages and (f) corresponding contour plots. (g) Schematic diagram of the CoSx/MoO3 modified separator to improve the electrochemical performance of LSBs.

To investigate the dynamic process of Li2S4 adsorption by the samples, we performed in situ UV-Vis testing on the diluted Li2S4 solution containing ZIF-67@CoSx/MoO3 and CoSx/MoO3 samples. As can be seen in Fig. 5b and Fig. S20a, the intensity of Li2S4 at 420 nm increased slightly at the beginning and decreased slowly afterwards. The optical image after Li2S4 adsorption showed a dark brown solution. This may be due to competition between the adsorption of Li2S4 by CoSx/MoO3 in the ZIF-67@CoSx/MoO3 sample and the adsorption of the solvent by parent ZIF-67. In contrast, the initial yellow solution of CoSx/MoO3 turned into a clear solution with almost no Li2S4 peak at 420 nm after 8 h of standing (Fig. 5c and Fig. S20b). This is because the hollow bimetallic heterostructure obtained under deeper sulfidation enhances the affinity for sulfur, enables the selective capture of Li2S4 and achieves rapid adsorption.35 To further characterize the impact of the samples on LiPS conversion, electrochemical impedance spectroscopy (EIS) testing was performed. Thanks to the introduction of Mo and the highly conductive sulfide, CoSx/MoO3 exhibited the lowest solution diffusion impedance and charge transfer resistance (Fig. S21). Meanwhile, we performed in situ monitoring of the impedance of ZIF-67@CoSx/MoO3/PP and CoSx/MoO3/PP during discharge. The impedance of ZIF-67@CoSx/MoO3/PP suddenly increased during the LiPS generation stage as shown in Fig. S22, indicating that the dissolution of LiPSs in the electrolyte hindered ion transfer. However, the impedance of CoSx/MoO3/PP showed no significant change during discharge as shown in Fig. 5d, indicating its effective adsorption and promotion of the conversion of LiPSs.36 Additionally, semi-quantitative analysis of the EIS spectra was performed using the distribution of relaxation time (DRT) to further verify the stability and reversibility of the batteries with CoSx/MoO3/PP, as presented in Fig. 5e. The P1 peak corresponds to the contact resistance at the current collector and electrode interface, and the P2 and P3 peaks are related to ion transport at the anode and cathode, respectively. During discharge, the peaks of P2 and P3 both gradually shifted to lower frequencies and lower intensities, indicating a rapid mass transfer process at the cathode and anode. The P4 peak is associated with the charge transfer process at the positive electrode. As the voltage dropped below 2.3 V, the P4 peak increased slightly but decreased gradually as the discharge continued, which showed the rapid selective adsorption of LiPSs by the synergistic action of CoSx and MoO3 and prevented the mass transfer process from slowing down due to the dissolution of a large amount of LiPSs in the electrolyte.37,38 The intensity color map of the DRT curves in Fig. 5f displayed more obvious changes in the P3 and P4 peaks. Overall, the improved sulfur cathode kinetics are illustrated in Fig. 5g. The hollow structure of the bimetallic heterostructure effectively suppresses the shuttle effect of LiPSs and accelerates their conversion into insoluble LiPSs.

Conclusions

In this work, MOF-derived heterostructure modified separators were designed to improve the electrochemical performance of LSBs. By varying the degree of sulfidation of ZIF-67, we successfully prepared core–shell structured ZIF-67@CoSx/MoO3 and hollow CoSx/MoO3, respectively. LSBs with CoSx/MoO3-modified separators show excellent electrochemical performance due to the complete sacrifice of ZIF-67 and the bimetallic synergistic system. The MoO3 has strong chemical interactions with LiPSs; in addition, CoSx has high conductivity that can speed up the transport of ions between the positive and negative sides and the conversion of LiPSs. The LSBs equipped with CoSx/MoO3/PP exhibit a high discharge capacity of 632 mA h g−1 at 2C rate and a capacity decay rate of 0.041% after 500 cycles at 1C, demonstrating excellent cycling stability and rate performance. The excellent electrochemical performance of CoSx/MoO3 highlights the potential of MOF-derived heterostructure materials in the field of LSBs and provides new ideas for the development of high-performance LSBs.

Author contributions

Rongmei Zhu and Yuxuan Jiang conceived the ideas and wrote the paper; Yuxuan Jiang synthesized the materials and performed the experiments; Bingxin Sun performed the in situ experiments; Wang Zhang helped analyze the data; Rongmei Zhu and Huan Pang supervised the work and provided funding.

Data availability

The data supporting this article have been included as part of the ESI.

Conflicts of interest

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

Acknowledgements

This work was supported by the Natural Science Foundation of Jiangsu Province (Grant No. BK20230069, BX2023026 and BK20200044), the National Natural Science Foundation of China (U1904215), the Changjiang Scholars’ Program of the Ministry of Education (Q2018270), the Six Talent Peaks Project in Jiangsu Province, the Top Talent Project of Yangzhou University, and the Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (NRF-2022R1A6A1A03063039). We also acknowledge the Priority Academic Program Development of Jiangsu Higher Education Institutions and the technical support received from the Testing Center of Yangzhou University.

References

  1. W. Zhang, Z. Zheng, L. Lin, X. Zhang, M. Bae, J. Lee, J. Xie, G. Diao, H. Im, Y. Piao and H. Pang, Ultrafast Synthesis of Graphene-Embedded Cyclodextrin-Metal-Organic Framework for Supramolecular Selective Absorbency and Supercapacitor Performance, Adv. Sci., 2023, 10, 2304062 CrossRef CAS PubMed.
  2. S. Zheng, Y. Sun, H. Xue, P. Braunstein, W. Huang and H. Pang, Dual-ligand and hard-soft-acid-base strategies to optimize metal-organic framework nanocrystals for stable electrochemical cycling performance, Natl. Sci. Rev., 2022, 9, nwab197 CrossRef CAS PubMed.
  3. C. Liu, Y. Bai, W. Li, F. Yang, G. Zhang and H. Pang, In Situ Growth of Three-Dimensional MXene/Metal–Organic Framework Composites for High-Performance Supercapacitors, Angew. Chem., 2022, 134, e202116282 CrossRef.
  4. D. Luo, et al., Constructing Abnormal Step-Scheme Nano-Heterointerfaces as Sulfur Electrocatalysts with Desirable Electron Confinement for Practical Li-S Battery, Renewables, 2024, 2, 61–72 CrossRef.
  5. X. C. Liu, et al., Design of Defect-Rich MgCo-Layered Double Hydroxide Microspheres as Highly Effective Sulfur Reduction Reaction Electrocatalyst for High-Performance Li-S Batteries, Renewables, 2024, 1–9 Search PubMed.
  6. Y. Jiang, M. Du, P. Geng, B. Sun, R. Zhu and H. Pang, CoO/MoO3@Nitrogen-Doped carbon hollow heterostructures for efficient polysulfide immobilization and enhanced ion transport in Lithium-Sulfur batteries, J. Colloid Interface Sci., 2024, 664, 617–625 CrossRef CAS PubMed.
  7. L. Ni, G. Zhao, Y. Wang, Z. Wu, W. Wang, Y. Liao, G. Yang and G. Diao, Coaxial Carbon/MnO2 Hollow Nanofibers as Sulfur Hosts for High-Performance Lithium-Sulfur Batteries, Chem. – Asian J., 2017, 12, 3128–3134 CrossRef CAS PubMed.
  8. Z. Wu, W. Wang, Y. Wang, C. Chen, K. Li, G. Zhao, C. Sun, W. Chen, L. Ni and G. Diao, Three-dimensional graphene hollow spheres with high sulfur loading for high-performance lithium-sulfur batteries, Electrochim. Acta, 2017, 224, 527–533 CrossRef CAS.
  9. X. Wang, X. Zhang, Y. Zhao, D. Luo, L. Shui, Y. Li and Z. Chen, Accelerated Multi-step Sulfur Redox Reactions in Lithium-Sulfur Batteries Enabled by Dual Defects in Metal-Organic Framework-based Catalysts, Angew. Chem., 2023, 135(42), e202306901 CrossRef.
  10. L. Wu, J. Hu, X. Yang, Z. Liang, S. Chen, L. Liu, H. Hou and J. Yang, Synergistic effect of adsorption and electrocatalysis of CoO/NiO heterostructure nanosheet assembled nanocages for high-performance lithium–sulfur batteries, J. Mater. Chem. A, 2022, 10, 23811–23822 RSC.
  11. S. Qiu, J. Zhang, X. Liang, Y. Li, J. Cui and M. Chen, Tunable MOFs derivatives for stable and fast sulfur electrodes in Li-S batteries, Chem. Eng. J., 2022, 450, 138287 CrossRef CAS.
  12. J. Wang and J. Li, Cobalt-based zeolitic imidazolate frameworks modified separator as efficient polysulfide adsorbent for high performance lithium-sulfur batteries, J. Colloid Interface Sci., 2021, 584, 354–359 CrossRef CAS PubMed.
  13. Y. Wen, Z. Shen, J. Hui, H. Zhang and Q. Zhu, Co/CoSe Junctions Enable Efficient and Durable Electrocatalytic Conversion of Polysulfides for High-Performance Li–S Batteries, Adv. Energy Mater., 2023, 13, 2204345 Search PubMed.
  14. Y. Luo, M. Wu, D. Zhang, J. Liu, Y. He, W. Zhang, S. Liu, Y. Dong, C. Xiang, L. Yang, H. Liu, H. Shu, X. Wang and M. Chen, Boosted Polysulfide Conversion by Co, Mn Bimetallic-Modulated Nitrogen–Carbon Material for Advanced Lithium–Sulfur Batteries, ACS Sustainable Chem. Eng., 2023, 11, 1087–1099 CrossRef CAS.
  15. X. Guan, M. Huang, L. Yang, G. Wang and X. Guan, Facial design and synthesis of CoSx/Ni-Co LDH nanocages with rhombic dodecahedral structure for high-performance asymmetric supercapacitors, Chem. Eng. J., 2019, 372, 151–162 CrossRef CAS.
  16. B. Li, Q. Su, L. Yu, J. Zhang, G. Du, D. Wang, D. Han, M. Zhang, S. Ding and B. Xu, Tuning the Band Structure of MoS2 via Co9S8@MoS2 Core-Shell Structure to Boost Catalytic Activity for Lithium-Sulfur Batteries, ACS Nano, 2020, 14, 17285–17294 CrossRef CAS PubMed.
  17. K. Wen, L. Huang, L. Qu, T. Deng, X. Men, L. Chen and J. Wang, g-C3N4/MoO3 composite with optimized crystal face: A synergistic adsorption-catalysis for boosting cathode performance of lithium-sulfur batteries, J. Colloid Interface Sci., 2023, 649, 890–899 CrossRef CAS PubMed.
  18. F. Wang, H. Fu, F.-X. Wang, X.-W. Zhang, P. Wang, C. Zhao and C.-C. Wang, Enhanced catalytic sulfamethoxazole degradation via peroxymonosulfate activation over amorphous CoSx@SiO2 nanocages derived from ZIF-67, J. Hazard. Mater., 2022, 423, 126998 CrossRef CAS PubMed.
  19. J. Bai, T. Meng, D. Guo, S. Wang, B. Mao and M. Cao, Co9S8@MoS2 Core Shell Heterostructures as Trifunctional Electrocatalysts for Overall Water Splitting and Zn Air Batteries, ACS Appl. Mater. Interfaces, 2018, 10, 1678–1689 CrossRef CAS PubMed.
  20. J. Yang, Y. Qu, X. Lin, L. Wang, Z. Zheng, J. Zhuang and L. Duan, MoO3/MoS2 flexible paper as sulfur cathode with synergistic suppress shuttle effect for lithium-sulfur batteries, Electrochim. Acta, 2022, 418, 140378 CrossRef CAS.
  21. K. Ji, Q. Che, Y. Yue and P. Yang, Ni-Doped Co3S4 Hollow Nanobox for the Hydrogen Evolution Reaction, ACS Appl. Nano Mater., 2022, 5, 9901–9909 CrossRef CAS.
  22. J. Gong, W. Luo, Y. Zhao, M. Xie, J. Wang, J. Yang and Y. Dai, Co9S8/NiCo2S4 core-shell array structure cathode hybridized with PPy/MnO2 core-shell structure anode for high-performance flexible quasi-solid-state alkaline aqueous batteries, Chem. Eng. J., 2022, 434, 134640 CrossRef CAS.
  23. D. Lei, W. Shang, X. Zhang, Y. Li, S. Qiao, Y. Zhong, X. Deng, X. Shi, Q. Zhang, C. Hao, X. Song and F. Zhang, Facile Synthesis of Heterostructured MoS2-MoO3 Nanosheets with Active Electrocatalytic Sites for High-Performance Lithium–Sulfur Batteries, ACS Nano, 2021, 15, 20478–20488 CrossRef CAS PubMed.
  24. Z. Wu, W. Wang, Y. Wang, C. Chen, K. Li, G. Zhao, C. Sun, W. Chen, L. Ni and G. Diao, Three-dimensional graphene hollow spheres with high sulfur loading for high-performance lithium-sulfur batteries, Electrochim. Acta, 2017, 224, 527–533 CrossRef CAS.
  25. J. Fu, C. Bie, B. Cheng, C. Jiang and J. Yu, Hollow CoSx Polyhedrons Act as High-Efficiency Cocatalyst for Enhancing the Photocatalytic Hydrogen Generation of g-C3N4, ACS Sustainable Chem. Eng., 2018, 6, 2767–2779 CrossRef CAS.
  26. Y. Tang, F. Jing, Z. Xu, F. Zhang, Y. Mai and D. Wu, Highly Crumpled Hybrids of Nitrogen/Sulfur Dual-Doped Graphene and Co9S8 Nanoplates as Efficient Bifunctional Oxygen Electrocatalysts, ACS Appl. Mater. Interfaces, 2017, 9, 12340–12347 CrossRef CAS PubMed.
  27. J. Wang, J. Xu, X. Guo, T. Shen, C. Xuan, B. Tian, Z. Wen, Y. Zhu and D. Wang, Synergistic regulation of nickel doping/hierarchical structure in cobalt sulfide for high performance zinc-air battery, Appl. Catal., B, 2021, 298, 120539 CrossRef CAS.
  28. L. Yang, L. Zhang, G. Xu, X. Ma, W. Wang, H. Song and D. Jia, Metal–Organic-Framework-Derived Hollow CoSx@MoS2 Microcubes as Superior Bifunctional Electrocatalysts for Hydrogen Evolution and Oxygen Evolution Reactions, ACS Sustainable Chem. Eng., 2018, 6, 12961–12968 CrossRef CAS.
  29. J. Li, B. Gao, Z. Shi, J. Chen, H. Fu and Z. Liu, Graphene/Heterojunction Composite Prepared by Carbon Thermal Reduction as a Sulfur Host for Lithium-Sulfur Batteries, Materials, 2023, 16, 4956 CrossRef CAS PubMed.
  30. H. Liu, X. Yang, B. Jin, M. Cui, Y. Li, Q. Li, L. Li, Q. Sheng, X. Lang, E. Jin, S. Jeong and Q. Jiang, Coordinated Immobilization and Rapid Conversion of Polysulfide Enabled by a Hollow Metal Oxide/Sulfide/Nitrogen-Doped Carbon Heterostructure for Long-Cycle-Life Lithium-Sulfur Batteries, Small, 2023, 19, 2300950 CrossRef CAS PubMed.
  31. Z. Gu, C. Cheng, T. Yan, G. Liu, J. Jiang, J. Mao, K. Dai, J. Li, J. Wu and L. Zhang, Synergistic effect of Co3Fe7 alloy and N-doped hollow carbon spheres with high activity and stability for high-performance lithium-sulfur batteries, Nano Energy, 2021, 86, 106111 CrossRef CAS.
  32. Y. Yoshie, K. Hori, T. Mae and S. Noda, High-energy-density Li–S battery with positive electrode of lithium polysulfides held by carbon nanotube sponge, Carbon, 2021, 182, 32–41 CrossRef CAS.
  33. L. Zhang, T. Li, X. Zhang, Z. Ma, Q. Zhou, Y. Liu, X. Jiang, H. Zhang, L. Ni and G. Diao, Core–shell polyoxometalate-based zeolite imidazole framework-derived multi-interfacial MoSe2/CoSe2 @NC enabling multi-functional polysulfide anchoring and conversion in lithium–sulfur batteries, J. Mater. Chem. A, 2023, 11, 3105–3117 RSC.
  34. Y. Wu, N. Wu, X. Jiang, S. Duan, T. Li, Q. Zhou, M. Chen, G. Diao, Z. Wu and L. Ni, Bifunctional K 3PW12O40/Graphene Oxide-Modified Separator for Inhibiting Polysulfide Diffusion and Stabilizing Lithium Anode, Inorg. Chem., 2023, 62, 15440–15449 CrossRef CAS PubMed.
  35. Z. Ma, W. Liu, X. Jiang, Y. Liu, G. Yang, Z. Wu, Q. Zhou, M. Chen, J. Xie, L. Ni and G. Diao, Wide-Temperature-Range Li–S Batteries Enabled by Thiodimolybdate [Mo2S12]2− as a Dual-Function Molecular Catalyst for Polysulfide Redox and Lithium Intercalation, ACS Nano, 2022, 16, 14569–14581 CrossRef CAS PubMed.
  36. L. Ni, J. Gu, X. Jiang, H. Xu, Z. Wu, Y. Wu, Y. Liu, J. Xie, Y. Wei and G. Diao, Polyoxometalate-Cyclodextrin-Based Cluster-Organic Supramolecular Framework for Polysulfide Conversion and Guest–Host Recognition in Lithium-sulfur Batteries, Angew. Chem., Int. Ed., 2023, 62, e202306528 CrossRef CAS PubMed.
  37. Y. Zhang, C. Kang, W. Zhao, Y. Song, J. Zhu, H. Huo, Y. Ma, C. Du, P. Zuo, S. Lou and G. Yin, d-p Hybridization-Induced “Trapping–Coupling–Conversion” Enables High-Efficiency Nb Single-Atom Catalysis for Li–S Batteries, J. Am. Chem. Soc., 2023, 145, 1728–1739 CrossRef CAS PubMed.
  38. Q. Liang, S. Wang, X. Lu, X. Jia, J. Yang, F. Liang, Q. Xie, C. Yang, J. Qian, H. Song and R. Chen, High-Entropy MXene as Bifunctional Mediator toward Advanced Li–S Full Batteries, ACS Nano, 2024, 18, 2395–2408 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4qi01249f

This journal is © the Partner Organisations 2024
Click here to see how this site uses Cookies. View our privacy policy here.