Jiayi Ren and
Qihao Zhao*
School of Chemical, Marine and Life Sciences, Dalian University of Technology, Dalian 116023, China. E-mail: 18641157899@163.com
First published on 15th November 2024
Lithium–sulfur batteries (LSBs) with metal lithium as the anode and elemental sulfur as the cathode active materials have attracted extensive attention due to their high theoretical specific capacity (1675 mA h g−1), high theoretical energy density (2600 W h kg−1), low cost, and environmental friendliness. However, the discharge intermediate lithium polysulfide undergoes a shuttle side reaction between the two electrodes, resulting in low utilization of the active substances. This limits the capacity and cycle life of LSBs and further delays their commercial development. However, the number of active sites and electron transport capacity of such catalysts still do not meet the practical development needs of lithium–sulfur batteries. In view of these issues, this paper focuses on a zinc–cobalt compound catalyst, modifying it through heteroatom doping, bimetallic synergistic effect and heterogeneous structure design to enhance the performance of LSBs as a separator modification material. A carbon shell-supported boron-doped ZnS/CoS2 heterojunction catalytic material (B–ZnS/CoS2@CS) was prepared, and its performance in lithium–sulfur batteries was evaluated. A carbon substrate (CS) was prepared by pyrolysis of sodium citrate, and the boron-doped ZnS/CoS2 heterojunction catalyst was formed on the CS using a one-step solvothermal method. The unique heterogeneous interface provides numerous active sites for the adsorption and catalysis of polysulfides. The uniformly doped, electron-deficient boron further enhances the Lewis acidity of the ZnS/CoS2 heterojunction, while also regulating electron transport. The B–ZnS/CoS2@CS catalyst effectively inhibits the diffusion of LiPS anions by utilizing additional lone-pair electrons. The lithium–sulfur battery using the catalyst-modified separator achieves a high specific capacity of 1241 mA h g−1 at a current density of 0.2C and retains a specific capacity of 384.2 mA h g−1 at 6.0C. In summary, B–ZnS/CoS2@CS heterojunction catalysts were prepared through boron doping modification. They can promote the conversion of polysulfides and effectively inhibit the shuttle effect. The findings provide valuable insights for the future modification and preparation of lithium–sulfur battery catalysts.
Conventional LSBs are composed of a sulfur-based cathode, a porous diaphragm, a lithium anode, and an organic electrolyte.12–18 Among them, the sulfur-based cathode serves as the reaction center, and a complex of sulfur powder and carbon material is generally used as the main ingredient. The porous diaphragm serves as the medium for blocking the cathode and anode, and is usually an electrically insulating polyolefin-type diaphragm. The lithium anode is a commercially available lithium metal wafer. Finally, the electrolyte is usually an ether electrolyte with added lithium salts. Unlike the energy storage mechanism of lithium-ion batteries, the charging and discharging process of LSBs is more complex and involves a solid–liquid–solid phase transition process.9,10,13,14,19,20
The total equation for the sulfur reduction reaction (SRR) of LSBs during discharge is:
S8 + 16Li → 8Li2S |
The sulfur is in an unlithiated state when the battery is in its initial state, i.e., the initial state of the reaction is the discharge process. The specific reaction is mainly divided into four stages as follows.
Stage I: S8 (s) → Li2S8 (l) |
Stage II: Li2S8 (l) → Li2S6 (l)/Li2S4 (l) |
Stage III: Li2S4 (l) → Li2S2 (s)/Li2S (s) |
Stage IV: Li2S2 (s) → Li2S (s) |
Lithium–sulfur batteries are expected to be the next-generation alternative for energy storage systems due to its many advantages, including higher theoretical energy density, and abundant and low-cost sulfur materials. However, there are still many challenges to solve in practical application,21–24 including the following aspects:
(1) The electronic insulating properties of sulfur and its final discharge products.25–27 The electrical conductivity of sulfur at room temperature is 5 × 10−30 S cm−1 and that of Li2S2/Li2S is 3.6 × 10−7 S cm−1. The lower electron transport capacity leads to an increase in the polarization of the electrochemical reaction. Consequently, the deposition of the passivation layer causes irreversible loss of the active material.
(2) Shuttle effect.28 In the sulfur reduction process, the long-chain polysulfides in the liquid phase of the electrolyte pass through the diaphragm into the negative electrode under the effect of a concentration gradient to react with lithium and further generate short-chain polysulfides. Under the continuous reaction, the short-chain polysulfide diffuses back to the positive electrode to carry out charging and the oxidation reaction under the impetus of the concentration gradient, and generates a long-chain polysulfide. The phenomenon is called the “shuttle effect” (Fig. 2b). This phenomenon leads to corrosion of the lithium anode and poor cycle reversibility, affecting the battery life.
(3) Volume expansion29 The large density difference between S (2.07 g cm−1) and Li2S (1.66 g cm−1) results in a volume change of about 80% inside the battery. With the continuing battery charging and discharging process, the large volume change will cause the anode carbon and sulfur composite structure to form a gap, which further affects the sulfur utilization rate.
(4) Safety hazards caused by lithium dendrites.15 Generally speaking, the uneven surface of the lithium anode, along with the non-uniform distribution of lithium ions and current, leads to uneven deposition. This results in the continuous growth of lithium dendrites, which can penetrate through the diaphragm, causing an internal short circuit in the battery and posing potential safety hazards.
In response to the above problems, researchers have developed many solutions, including the design of porous carbon structures with the appropriate pore size,30 anode composite catalyst materials,31 modified binders,32 electrolyte additives,33–36 and the assembly of solid-state electrolyte batteries.37 Compared to the above measures, the utilization of catalyst-modified diaphragms has attracted a lot of attention from researchers due to its advantages, such as simple experimental manipulation, as well as significant performance improvement.
Metal sulfides have good potential for applications in the field of LSBs because of their higher room temperature conductivity, but a single catalytic active site is still insufficient for them to realize the catalytic conversion of a large number of LiPSs. Bimetallic sulfide heterostructures have a unique role in catalyzing LiPSs because of their ability to expose more active sites due to interfacial coupling. For example, Lee and colleagues38 reported on a CoS2/Fe7S8/NG heterostructure, in which a strong built-in electric field was spontaneously generated at the contact interface between the two, owing to the equilibrium tendency of the Fermi energy levels, which greatly accelerated the electron transfer ability and LiPSs adsorption ability. However, the catalytic effect of heterostructures is not limited to this. Furthermore, the design of defective heterostructures to expose a richer variety of active sites based on interfacial engineering to achieve a higher level of battery performance has been demonstrated by researchers.
A one-step hydrothermal method was used to synthesize carbon-shell-bearing boron-doped ZnS/CoS2 heterojunction catalysts (B–ZnS/CoS2@CS) using carbon shells (CS) as substrates. The fabricated catalysts better retained the carbon shell structure, which was favorable for the transport of electrons and ions inside the LSBs, and this advantage was further amplified by the uniform introduction of electron-deficient B atoms and the construction of the internal electric field. Meanwhile, the lattice of the ZnS/CoS2 heterostructure modified by boron atoms was distorted, exposing more active sites. Combining the strong adsorption ability of CoS2 and the high catalytic activity of ZnS, the LSBs assembled with the B–ZnS/CoS2@CS-modified septa exhibited superior cycling performance. With 1500 cycles at 2.0C and a capacity decay rate of 0.043% per cycle, the cells maintained a high area-specific capacity of 6.22 mA h cm−2 after 120 cycles at 0.2C even under the harsh conditions of a liquid–sulfur ratio (E/S) of 8 μl mg−1 and a sulfur loading of 6.14 mg cm−2.
The CS morphology obtained using annealing is shown in Fig. 2a, where the material shows a stacked lamellar structure. Thus, the longer ultrasonic dispersion was carried out while stretcher-loading the catalyst. The corresponding XRD results are shown in Fig. 2b, where two distinct diffraction peaks, representing the (0 0 2) and (1 0 1) crystalline surfaces of the graphitic carbon, are shown at approximately 25.3° and 43.8° diffraction angles, respectively.39–42
The morphology of the ZnS/CoS2@CS catalyst obtained by solvothermal method is shown in Fig. 3a. It can be observed that many nanospheres are distributed between the lamellar CS layers, indicating the successful loading of the catalyst. However, the accumulation of small spheres is serious, which reduces the specific surface area of CS to some extent and is not favorable for the adsorption of polysulfides. Notably, after boron doping of the catalyst, as shown in Fig. 3c, the morphology of B–ZnS/CoS2@CS still shows a sheet-like structure without obvious stacking and changes. Meanwhile, from Fig. 3d, it is found that the catalyst spheres are more uniformly distributed, in which the size of the spheres is reduced to approximately 125 nm. This suggests that boron doping facilitates the increase of the specific surface area of the material, further exposing more polysulfide sulfide adsorption and catalytic sites.
The TEM of the catalyst is shown in Fig. 4. It can be observed that the morphology of the catalysts (Fig. 4a and b) is consistent with the SEM photographs, showing a thin sheet structure and uniformly distributed catalysts. In addition, the SAED plots in Fig. 4c show diffraction rings corresponding to the ZnS (1 0 0), ZnS (1 0 2), ZnS (1 1 2), CoS2 (2 0 0) and CoS2 (4 2 0) planes, respectively, confirming the polycrystalline nature of B–ZnS/CoS2@CS. To further investigate the microstructure of the B–ZnS/CoS2@CS catalysts, HRTEM observation and analysis were carried out. Distinguishable lattice stripes belonging to ZnS and CoS2, as well as disordered regions belonging to CS, were observed from Fig. 4d. Meanwhile, by applying the discrete Fourier transform (FFT) to the different lattice regions, the polycrystalline nature of the catalysts was confirmed in the diffracted rings, corresponding to ZnS (0.313 nm) and CoS2 (0.319 nm) with different degrees of local lattice distortion at the lattice stripe spacing. This suggests that boron doping can interfere with the lattice structure of the catalysts, favoring the creation of defects and exposing more catalytically active sites.43–46 Fig. 4g shows that Zn, Co, S, C and B elements are uniformly distributed on B–ZnS/CoS2@CS, further indicating the uniform introduction of boron.
![]() | ||
Fig. 4 (a and b) TEM image, (c) SAED pattern, (d–f) HRTEM image of B–ZnS/CoS2@CS, and (g) EDX mapping results. |
The XRD test results of the catalyst are shown in Fig. 5a. The samples before and after doping corresponded to the ZnS crystalline phase (PDF # 36-1450) and CoS2 crystalline phase (PDF # 41-1471). Together with the TEM images, these results indicated the successful preparation of the heterojunction catalysts. It was also observed that after the catalyst was loaded onto the carbon substrate, the characteristic peak of the (0 0 2) crystalline plane belonging to graphitic carbon was shifted left to 21.68°. The layer spacing of the CS was also calculated from Bragg's equation to increase from 0.35 to 0.41 nm, which was attributed to the loading of the catalyst on the CS. In addition, B–ZnS/CoS2@CS shows weaker intensity of carbon characteristic peaks, indicating that the carbon becomes less crystalline and more defective after boron doping. As shown in the Raman spectra of Fig. 5b, the different catalysts show obvious characteristic peaks (D and G peaks) at 1342 cm−1 and 1578 cm−1. The CS:
ZnS/CoS2@CS
:
B–ZnS/CoS2@CS ID/IG ratios are 1.27, 1.49 and 1.53, and were calculated based on the fitting results. These results further proved that the boron doping exposed more defective sites in the catalysts, which would be favorable for the sulfur reduction–oxidation reaction. Notably, the Raman spectrum of the B–ZnS/CoS2@CS catalyst was found to display a characteristic peak at 1020 cm−1, which may be the characteristic peak of the BO33− species.
The N2 adsorption and desorption curves of the catalysts are shown in Fig. 6a. As seen, the specific surface areas of the catalysts before and after doping were 63.4 m2 g−1 and 74.9 m2 g−1, respectively. The specific surface area of the B–ZnS/CoS2@CS catalyst was slightly higher than that of the ZnS/CoS2@CS, suggesting that more polysulfide adsorption and catalytically active sites were exposed by the B–ZnS/CoS2@CS catalyst, which agreed with the analytical results of SEM. In addition, the pore sizes of the two catalysts were mainly concentrated at 2–40 nm with a mesoporous structure (Fig. 6b). The pore structure of the B–ZnS/CoS2@CS catalyst was more distributed in this interval, which was conducive to the selective transmission of ions and improvement of the physical properties of the modified diaphragm to reduce the cell polarization.47–49
![]() | ||
Fig. 6 Pore structure of ZnS/CoS2@CS and B–ZnS/CoS2@CS: (a) N2 adsorption–desorption curve and (b) pore size distribution. |
Further analysis of the chemical states of the boron-doped heterojunction catalysts was carried out by XPS. The Zn 2p XPS of the ZnS/CoS2@CS and B–ZnS/CoS2@CS catalysts are shown in Fig. 7a. After boron doping, the binding energies of the 2p3/2 and 2p1/2 characteristic peaks of Zn were changed from 1022.73 eV and 1045.75 eV, respectively, to 1022.50 eV and 1045.52 eV, shifting by 0.23 eV in the direction of low binding energy. Fig. 7b shows the 2p XPS plots of Co before and after boron doping, and it can be found that the characteristic peak binding energies of the 2p3/2 and 2p1/2 peaks of Co2+ are changed from 781.41 eV and 797.68 eV to 781.81 eV and 798.08 eV, respectively, and are shifted by 0.4 eV towards the direction of high binding energy. The XPS fitting results indicate that the introduction of boron causes the flow of electrons from CoS2 to ZnS inside the heterojunction. This enhances the reduction of the negatively charged polysulfide anions due to the gain of electrons by ZnS, and enhances the adsorption of polysulfides due to the loss of electrons by CoS2, which is positively charged. The results indicate that the uniform doping of boron can effectively modulate the electronic properties of the ZnS/CoS2 heterostructures, thus exhibiting enhanced catalytic activity. Fig. 7c shows the S 2p XPS spectra of the B–ZnS/CoS2@CS catalyst, in which the four sets of spin-split peaks, 161.94/163 eV; 163.02/164.04 eV; 164.08/165.29 eV and 168.95/170.02 eV, represent the S–S, M–S, C–S–C, and SO42−, respectively.50–52 The presence of the C–S–C bonds indicates that the catalyst is stretcher-loaded onto the carbon shell in the form of bonding, and this structural advantage lies in the ability of the carbon substrate to provide the catalyst with a strong electron transporting capability. The XPS spectrum of B 1s is shown in Fig. 7d, which displays a distinct characteristic peak at 187.56 eV. This is due to the generation of metal–boron bonding as a result of the doping of boron into the catalyst, suggesting a successful doping of boron.53
![]() | ||
Fig. 7 (a) Zn 2p, (b) Co 2p, (c) S 2p and (d) B 1s XPS spectra of the ZnS/CoS2@CS and B–ZnS/CoS2@CS catalysts. |
The adsorption capacity of the catalytic materials for LiPSs determines the utilization of the active substance, for which visual adsorption experiments were performed on the catalysts. As shown in Fig. 8a, 100 mg of CS, ZnS/CoS2@CS and B–ZnS/CoS2@CS catalytic materials were placed in transparent glass vials, respectively, and 10 ml of 0.05 M Li2S6 solution was added to the interior and left to stand for 12 h. It was observed that the LiPSs solution impregnated with the B–ZnS/CoS2@CS material became almost colorless. Conversely, the LiPSs solution impregnated with the CS- and the ZnS/CoS2@CS materials still exhibited varying degrees of yellow color, although the LiPS solution with ZnS/CoS2@CS was lighter in color compared to CS. The observed results indicated that the boron-doped catalyst could effectively improve the adsorption capacity of ZnS/CoS2@CS on LiPSs. Meanwhile, it was found that the degree of adsorption capacity enhancement was less desirable after the ZnS/CoS2 heterojunction catalysts were stretched on CS. This was due to the fact that the agglomerated morphology of the ZnS/CoS2@CS catalysts affected the chemical adsorption strength of LiPSs to a certain extent. After the adsorption experiments, the supernatant was aspirated and diluted 100-fold with DOL/DME solvent mixture for ultraviolet-visible spectroscopy (UV-vis) testing. As shown in Fig. 8b, the UV-vis results agreed with the adsorption experimental observations. The solution impregnated with CS showed the highest Li2S6 peak intensity, the solution impregnated with ZnS/CoS2@CS exhibited a slightly lower peak intensity, and the solution impregnated with B–ZnS/CoS2@CS solution showed the lowest peak intensity, indicating that the boron-doped heterojunction catalyst has the best LiPSs adsorption capacity.
![]() | ||
Fig. 8 (a) Digital photographs and (b) UV-vis spectra of different catalysts after soaking in Li2S6. |
The particulate samples after the adsorption experiments were dried and analyzed for their chemical states using XPS. As shown in Fig. 9a, the Zn 2p XPS spectrum showed a set of characteristic peaks at 1022.30 eV and 1045.32 eV. These peaks were shifted by 0.2 eV towards low binding energy compared to that before adsorption, which was typical of Lewis acid–base interactions. The Co 2p XPS spectrum is shown in Fig. 9b, and the intensity of the characteristic peaks belonging to Co3+ was significantly reduced. From calculations, it was found that the percentage of Co3+ chemical state orbitals decreased from 28.1% to 11.1% before and after adsorption. Meanwhile, the characteristic peak of SO42− shown in Fig. 9c belongs to the polysulfate complex. The increase of its peak intensity indicates the enhancement of oxidative strength from the side, which corresponds to the decrease of Co3+.54 Notably, the Li 1s XPS spectra (Fig. 9d) showed two characteristic peaks at 55.4 eV and 56.4 eV, corresponding to Li–S and Li–O bonds, respectively.55,56 This suggests that the Li in the polysulfides is capable of strong chemical bonding with the B–ZnS/CoS2@CS.
![]() | ||
Fig. 9 XPS spectra of B–ZnS/CoS2@CS after Li2S6 adsorption: (a) Zn 2p, (b) Co 2p, (c) S 2p and (d) Li 1s. |
![]() | ||
Fig. 10 Top SEM images of (a–c) CS, ZnS/CoS2@CS and B–ZnS/CoS2@CS-modified separators. Cross-sectional SEM images of (d–f) CS, ZnS/CoS2@CS and B–ZnS/CoS2@CS-modified separators. |
Electrolyte contact angle tests of different catalyst-modified diaphragms (Fig. 11a–c) revealed that the contact angle of the B–ZnS/CoS2@CS modified membrane (7.93°) was smaller than that of ZnS/CoS2@CS (9.28°) and CS (17.26°). The results indicate that the B–ZnS/CoS2@CS catalyst has the strongest ability to infiltrate the electrolyte, and the boron doping can increase the surface polarity of the catalyst and promote off-permeation.
![]() | ||
Fig. 11 The electrolyte contact angles of different separators: (a) CS, (b) ZnS/CoS2@CS, and (c) B–ZnS/CoS2@CS. |
![]() | ||
Fig. 12 (a) CV curves and (b) AC impedance spectra of LSBs assembled by the CS, ZnS/CoS2@CS and B–ZnS/CoS2@CS catalyst-modified separators. |
The cyclic voltammetry curves of LSBs with catalyst-modified diaphragms at 0.1–0.4 mV s−1 sweep rates are shown in Fig. 13a–c. As can be seen from the figure, the cell assembled with the B–ZnS/CoS2@CS modified membrane has the largest peak response current compared to the ZnS/CoS2@CS and CS cells, indicating that the B–ZnS/CoS2@CS catalyst has the strongest redox catalytic effect on LiPSs. Moreover, the R1, R2 and O1 peaks were accompanied by an increase in the sweep rate, an increase in the polarization voltage, and a more pronounced shift in the linear increase of the peak current, suggesting that the redox kinetics of LiPSs during charging and discharging of LSBs were mainly affected by ion diffusion. Therefore, it is usually believed that the diffusion ability of Li+ affects the conversion process of LiPSs, which in turn has a certain degree of influence on the performance of LSBs.
The diffusion behavior of Li+ was investigated by CV curves with different sweep speeds. The B–ZnS/CoS2@CS catalyst has the strongest ability to promote the diffusion of Li+. This is mainly due to the uniformly doped electron-deficient boron enhancing the Lewis acid nature of the ZnS/CoS2@CS heterojunction, which in turn enhances the ability to inhibit the diffusion of the anions of LiPSs with extra lone-pair electrons. Based on the law of charge conservation, this will effectively accelerate the diffusion of lithium ions into the cathode of lithium ions to the anode.
The Li+ diffusion coefficient (DLi+) can be calculated based on the slope of Ip × v−0.5. The values for different electrochemical reaction processes are shown in Table 1, and the DLi+ values of the cells assembled with B–ZnS/CoS2@CS modified membranes were 3.00 × 10−10, 0.29 × 10−10, and 0.26 × 10−10 cm2 s−1 for the R1, R1, and O1 processes, respectively. These values were higher than those of the CS and ZnS/CoS2@CS catalysts, indicating that the B–ZnS/CoS2@CS catalyst has the strongest ability to promote Li+ diffusion. The reason is the uniformly doped electron-deficient boron enhancing the Lewis acid nature of the ZnS/CoS2@CS heterojunction, which in turn enhances the ability to inhibit the diffusion of LiPSs anions with extra lone-pair electrons, which, based on the law of charge conservation, will effectively accelerate the diffusion of lithium ions to the cathode.
Peak | n | Slope of Ip × v−0.5 | DLi+ × 1010 cm2 s−1 | |
---|---|---|---|---|
CS | R1 | 0.5 | −0.57 | 0.28 |
R2 | 1.5 | −1.25 | 0.05 | |
O1 | 2 | −1.82 | 0.04 | |
ZnS/CoS2@CS | R1 | 0.5 | −1.21 | 1.27 |
R2 | 1.5 | −2.18 | 0.16 | |
O1 | 2 | −2.90 | 0.11 | |
B–ZnS/CoS2@CS | R1 | 0.5 | −1.88 | 3.00 |
R2 | 1.5 | −3.00 | 0.29 | |
O1 | 2 | −4.41 | 0.26 |
Symmetric cells were assembled using different catalyst electrodes for CV and EIS tests. The CV curves at a sweep rate of 2 mV s−1 are shown in Fig. 14, where the CS symmetric cell shows the lowest response peak current and the largest polarization voltage. Meanwhile, the B–ZnS/CoS2@CS symmetric cell exhibits a larger peak current and smaller polarization voltage than the ZnS/CoS2@CS cell.
The conversion kinetics of LiPSs on the anodes of different catalysts was further investigated by linear scanning voltammetry (LSV). As shown in Fig. 15a, the peak response currents (absolute values) of the LSV curves for B–ZnS/CoS2@CS, ZnS/CoS2@CS, and CS were 9.723, 8.212, and 5.411 mA, respectively. This indicated that the catalytic ability of the B–ZnS/CoS2@CS catalysts for the transition process of the soluble, long-chained LiPSs to the solid-phase Li2S2/Li2S phase was stronger. The fitted Tafel curves (Fig. 15b) showed that the Tafel slopes of the B–ZnS/CoS2@CS, ZnS/CoS2@CS and CS are 0.538, 0.617 and 0.769 V dec−1, respectively. The B–ZnS/CoS2@CS electrodes had the smallest Tafel slopes, which indicated that they had the strongest reaction kinetics. The above results indicate that the boron-doped heterostructure contributes to the accelerated conversion kinetics of LiPSs.
![]() | ||
Fig. 15 (a) The LSV curves and (b) the corresponding Tafel curves of CS, ZnS/CoS2@CS and B–ZnS/CoS2@CS. |
The above results indicate that the B–ZnS/CoS2@CS catalyst has the strongest kinetic promotion of LiPSs, which is further demonstrated by the Li2S nucleation pattern of the catalyst cell. As shown in Fig. 16a–c, comparing the nucleation process of Li2S on the B–ZnS/CoS2@CS, ZnS/CoS2@CS and CS cathodes, it was found that its specific capacity contribution on B–ZnS/CoS2@CS (532.8 mA h g−1) was higher than that on ZnS/CoS2@CS (447.5 mA h g−1) and CS (394.7 mA h g−1), indicating that the B–ZnS/CoS2@CS material has superior catalytic ability for Li2S nucleation. The dimensionless analysis of the i–t curves based on the Li2S nucleation process is shown in Fig. 16d–f. It can be observed that the Li2S nucleation on the CS electrode is mainly dominated by two-dimensional instantaneous (2DI) and three-dimensional progressive (3DP) diffusion. Meanwhile, the modeling of the Li2S nucleation on both ZnS/CoS2@CS and B–ZnS/CoS2@CS electrodes exhibits three-dimensional progressive (3DP) and three-dimensional instantaneous (3DI) mixed diffusion models, in which the electrodes of the boron-doped catalysts have some degree of improvement on the Li2S nucleation.
Next, the Li2S growth on the surfaces of different electrodes under the same test conditions was characterized by SEM. It was found that only a small amount of Li2S was deposited on the CS electrode (Fig. 17a). Conversely, the Li2S deposited on the ZnS/CoS2@CS electrode (Fig. 17b) was significantly increased, but still not as much as that on the B–ZnS/CoS2@CS electrode (Fig. 17c). The dense and uniform Li2S grown on the surface may be due to the doping of boron, which exposes more nucleation sites in the heterostructure. The above results indicate that the B–ZnS/CoS2@CS catalyst can effectively promote the transition kinetics of the liquid–solid phase and the uniform growth of Li2S. This ensures the uniform charge transfer for the subsequent reaction, and then activates the oxidation of Li2S and enhances the utilization of the active substance.
![]() | ||
Fig. 17 SEM images of the Li2S deposition on different catalyst electrodes: (a) CS; (b) ZnS/CoS2@CS and (c) B–ZnS/CoS2@CS. |
![]() | ||
Fig. 18 GITT curves of (a) CS, (b) ZnS/CoS2@CS, (c) B–ZnS/CoS2@CS batteries, and (d) the internal resistance of different batteries based on the normalized discharge/charge time. |
LSBs button batteries were assembled using different catalytic materials modified diaphragms, and their performance was tested and analyzed. The battery multiplicity performance is shown in Fig. 19a, from which it can be seen that the specific capacity of discharge at adopting B–ZnS/CoS2@CS are 0.2, 0.5, 1, 2, 3, 4, 5 and 6C multiplicities are 1241.5, 1058.5, 954.2, 861.9, 732.1, 610.8, 495.6 and 370.7 mA h g−1, respectively. The specific capacity also almost returned to the initial state when the cell recovered to 0.5C and 0.2C. Compared with the B–ZnS/CoS2@CS catalyst-modified diaphragm, the LSBs with ZnS/CoS2@CS and CS-modified membranes provided a lower discharge specific capacity, especially at 6C. The ZnS/CoS2@CS and CS batteries can only contribute 136.9 and 50 mA h g−1 specific capacity, respectively. The results indicate that the boron-doped heterostructured catalysts have a significant effect on the improvement of the battery performance.
![]() | ||
Fig. 19 Rate performance of batteries assembled with different separators: (a) rate performance and (b) charge/discharge curve of the 0.2C first cycle. |
The first turn constant-current charge/discharge curves of CS, ZnS/CoS2@CS and B–ZnS/CoS2@CS at 2C are shown in Fig. 19b. It can be observed that the LSBs cell with the B–ZnS/CoS2@CS diaphragm has the smallest polarization voltage, and the smallest charging/discharging overpotential during Li2S nucleation and dissolution. It is noteworthy that ZnS/CoS2@CS shows no significant improvement in polarization relative to CS, and presents a larger charging/discharging overpotential on the constant-current charging/discharging curves. This may be attributed to the agglomeration of the ZnS/CoS2 heterojunction catalysts loaded on CS, resulting in a larger electron transfer energy barrier during the reduction–oxidation reaction of Li2S.
LSBs with different catalyst-modified diaphragms were subjected to a 0.5C short cycle test, and the results are shown in Fig. 20. The initial specific capacities of LSBs assembled with B–ZnS/CoS2@CS, ZnS/CoS2@CS, and CS catalyst-modified diaphragms were 1016 mA h g−1, 965.8 mA h g−1, and 893.6 mA h g−1, respectively. After 200 charging and discharging cycles, the specific capacities decreased to 806.2 mA h g−1, 698.6 mA h g−1 and 567.9 mA h g−1, respectively. Furthermore, the capacity retention of the batteries assembled with different modified membranes were calculated to be 79.4%, 72.4% and 63.6% for B–ZnS/CoS2@CS, ZnS/CoS2@CS, and CS catalyst-modified membranes, respectively. The results show that the B–ZnS/CoS2@CS catalyst-modified membranes can effectively improve the cycling stability of LSBs. The surface morphology of the post-cycling diaphragms was observed by SEM. It was found that the surface of the post-cycling CS-modified membranes (Fig. 20b) showed many inhomogeneous Li2S2/Li2S passivation layers. Conversely, the passivation layers on the surfaces of the ZnS/CoS2@CS (Fig. 20e) and B–ZnS/CoS2@CS (Fig. 20h)-modified membranes exhibited a thinner and more homogeneous deposition. Even in the case of the B–ZnS/CoS2@CS modified film, the surface can also be observed with obvious catalytic materials, indicating that the boron-doped catalyst can effectively enhance the utilization of sulfur as the active substance by the heterojunction. In addition, by observing the surface of the lithium wafers after different cell cycles, the growth of lithium dendrites was found to be significantly different. Among them, dense lithium dendrites were distributed on the surface of lithium wafers of LSBs assembled with CS-modified film (Fig. 20c). The lithium dendrite growth on the surface of the lithium wafers of the battery with ZnS/CoS2@CS-modified film was improved to a great extent (Fig. 20f). Only a small amount of dendritic crystal growth existed in localized areas. In the battery with boron-doped catalysts, there was hardly any obvious dendritic crystals observed on the surface of the lithium wafers (Fig. 20i). This indicated that the B–ZnS/CoS2@CS catalyst can effectively inhibit the growth of lithium dendrites.
Subsequently, B–ZnS/CoS2@CS catalysts with different boron doping amounts were used to modify the diaphragm and assemble the cell, and the cycling performance was tested at 1.0C. As shown in Fig. 21, the 2B–ZnS/CoS2@CS catalyst prepared with 5 mmol of boron-sourced dopant endowed the battery with the best long-cycle performance, with an initial specific capacity of 815.4 mA h g−1 for the LSBs. The specific capacity was maintained at 519.1 mA h g−1 after 900 cycles, with a retention rate of 63.66% and an average decay rate of 0.0404% per cycle, which is a good cycling performance. The average decay rate per cycle was 0.0648% for the 1B–ZnS/CoS2@CS catalyst prepared by 1 mmol boron source dopant, causing the cell to exhibit the worst cycling specific capacity retention rate (45.7%) in a shorter cycle. Average decay rate per cycle is 0.0958%.
![]() | ||
Fig. 21 Comparison of the cycle performance of the B–ZnS/CoS2@CS catalyst batteries with different boron doping amounts at 1.0C. |
In this paper, the structures of bimetallic sulfides were modified by different boron doping strategies. The catalytic effect of the catalyst-modified diaphragms on the redox of polysulfides was further investigated, which suppressed the shuttle effect and achieved higher sulfur utilization to different extents, providing a certain reference value for the enhancement of metal compounds in the field of lithium–sulfur battery diaphragm catalysts for higher performance. However, the following shortcomings of this paper are still worth exploring in the future:
(1) The nucleation and dissolution kinetics of Li2S have a crucial position in the field of lithium–sulfur battery catalysis. Investigating the role of catalysts in controlling the nucleation behavior of Li2S through theoretical calculations and advanced in situ techniques will help to design more efficient catalytic materials.
(2) Heteroatoms and transition metal compounds belong to two major classes of systems. It is worth studying the various synergistic catalytic effects from the doping of heteroatoms (other than boron) and compounds (other than oxides and sulfides).
This journal is © The Royal Society of Chemistry 2024 |