MOF derived phosphorus doped cerium dioxide nanorods modified separator as efficient polysulfide barrier for advanced lithium-sulfur batteries

Xinyun Liu Long Yuan Xiaoli Peng Shilan Li Shengdong Jing Shengjun Lu Hua Lei Yufei Zhang Haosen Fan

Citation:  Xinyun Liu, Long Yuan, Xiaoli Peng, Shilan Li, Shengdong Jing, Shengjun Lu, Hua Lei, Yufei Zhang, Haosen Fan. MOF derived phosphorus doped cerium dioxide nanorods modified separator as efficient polysulfide barrier for advanced lithium-sulfur batteries[J]. Chinese Chemical Letters, 2025, 36(10): 110369. doi: 10.1016/j.cclet.2024.110369 shu

MOF derived phosphorus doped cerium dioxide nanorods modified separator as efficient polysulfide barrier for advanced lithium-sulfur batteries

English

  • Over the past few years, the excessive consumption of fossil fuels has triggered the problems of the century such as air pollution, global warming, and the oil crisis, so it is particularly important to develop sustainable new energy sources to solve environmental problems [15]. Meanwhile, the continuously increasing requirement for electronic and portable devices have rapidly accelerated the progress of lithium-ion batteries (LIBs) and sodium/potassium ion batteries with high-energy-density [69]. However, the traditional LIBs have approached their theoretical energy density limits, typically falling below 400 Wh/kg, this limitation has hindered their ability to meet the increasing demands of practical applications. Accordingly, lithium-sulfur batteries (LSBs) have been considered as one of the promising high-energy storage systems for the next generation due to their high theoretical capacity (1675 mAh/g), specific energy density (2600 Wh/kg), plentiful natural reserves and cost-effectiveness of the sulfur element [10]. Nevertheless, despite the potential advantages of LSBs, there are still several obstacles that hinder their industrial applications, including inadequate energy density and subpar capacity retention, attributed to the following challenges: (1) The inadequate ionic/electronic conductivity with sulfur and Li2S2/Li2S, causing sluggish reaction kinetics; (2) The formation of soluble lithium polysulfides (LiPSs) as intermediate reaction products in the electrolyte, causing a serious “shuttle effect” because of the slow redox kinetics of Li2S in the multi-step sulfur redox process, which leads to reduced Coulombic efficiency, anode degradation, and rapid capacity decline; (3) The remarkable expansion in volume experienced by the sulfur cathode during charge and discharge cycles, causing structural instability. The aforementioned issues collectively present formidable challenges in achieving high-performance LSBs with excellent cyclic stability and a prolonged lifespan.

    Up to date, researchers have carried out significant efforts to alleviate the shuttling of polysulfides. Key strategies include employing host materials to encapsulate sulfur within the cathodes [11,12], preparing modified separators [13,14] and protecting lithium anode [15], with the ultimate goal of enhancing the cyclic stability of LSBs. Modified commercial separators have been more widely investigated due to their ease of operation and practicality than the search for novel separators [16,17]. Generally, modified separators bind to carbon materials because of their excellent electrical conductivity, while nanostructured carbon materials can act as physical barriers to hinder the movement of LiPSs and facilitate their utilization in the battery system [1820]. However, the weak affinity between non-polar carbon materials and polar LiPSs is not sufficient to completely prevent the shuttling of LiPSs during the cycling process [21]. Therefore, introducing some polar oxides into the separator can effectively immobilize the LiPSs by physicochemical action [22]. However, most of the poorly conductive oxide LiPSs have slow reaction kinetics, a multitude of materials have been utilized to modify the separator with the aim of promoting the catalytic conversion and chemisorption of LiPSs, and inhibiting its shuttle effect in LSBs, such as metal oxides [23,24], metal sulfides [25], metal nitrides [26], COFs (covalent organic frameworks) [13], and MOFs (metal-organic frameworks) [27,28]. Among these, metal oxides are particularly favored due to their cost-effectiveness and straightforward preparation methods. In recent reports, it has been highlighted that rare-earth-based materials, including CeO2, Sc2O3, CeF3, Eu2O3, and Sm2O3, can effectively promote the catalytic conversion of LiPSs in LSBs [2933]. CeO2, a naturally abundant material, has been commonly utilized in the fields of energy storage and catalysis [34]. Rare-earth materials typically possess distinctive physical and chemical properties. Specifically, CeO2 with a cubic fluorite structure, characterized by high ionic and electronic conductivity, has the capability to transition between Ce4+ and Ce3+ oxidation states. In addition to this, the oxygen defects of CeO2 provide plentiful active sites that can enhance their interactions [35]. However, when used as a polysulfide catalyst in LSBs, metal oxides-based separators also have serious drawbacks, such as weak electron conductivity, limited ability to bind with LSBs, and sluggish catalytic kinetics [36,37]. Studies have shown that the incorporation of C, S, P, N and other heteroatoms in metal oxides will promote their polarity and conductivity [3840], thereby improving the catalytic kinetics and binding capacity of LSBs. Particularly, in the 3p orbitals and unoccupied 3d orbitals of phosphorus (P) where possessing a lone pair of electrons, exhibits robust interactions with lithium ions, rendering it an exceptionally promising doping element for enhancing the redox kinetics in metal oxides, thereby contributing to the advancement of LSBs. However, there are few reports on the design of P-doped metal oxides for LSBs [41], the major challenge is that metal oxides usually undergo a transformation into metal phosphides during the high-temperature phosphating process, this conversion makes it hard to control and analyze the impact of surface phosphorus atoms on the properties of metal oxides.

    Metal-organic frameworks (MOFs), which consist of organic ligands and inorganic metal ions, have recently been demonstrated as promising template/precursor to synthesize the metal compounds and widely used in sensors, separators, electrocatalysis and electrochemistry due to their homogeneous dispersed catalytic sites, large surface area, flexible pore size of MOFs, and excellent chemical properties [42,43]. In addition, it has previously been reported that the open metal sites of MOFs can serve as catalytic active sites for many reactions [44]. It has been reported that cerium oxide nanoparticles can adsorb polysulfides and promote their catalytic conversion [45]. In this manuscript, we have proposed a facile approach for fabrication of the phosphorus doped porous CeO2 (P-CeO2) separator materials in order to analyze the controllable influence of phosphorus atoms on the surface of metal oxides in LSBs, which is inspired by the remarkable chemical inertness [46], polyvalent states [47] and distinctive redox properties of CeO2 [48]. When combined with oxygen to form nanoparticles, CeO2 exhibits its fluorite crystal structure. Cerium atoms can easily transition quickly between Ce4+ and Ce3+ oxidation states, adjusting their electronic configuration to suit the chemical environment. In addition, due to its excellent chemical inertia, the problem of converting cerium oxide into cerium phosphide during phosphating can be solved. Firstly, the experimental analysis reveals that the phosphorus doped porous CeO2 modified PP separator (P-CeO2//PP) exhibits enhanced chemical interactions and superior catalytic activity compared with CeO2//PP separator. In addition, the substantial specific surface area and plentiful oxygen defects of P-CeO2 contribute to the manifestation of a highly exposed active interface, accelerate the rapid transformation of LiPSs, and regulate the nucleation of Li2S deposition. Surface morphology and structural characterization analyses reveal that P-CeO2 is primarily composed of numerous P-doped CeO2 nanocrystals. Benefitting from these advantages, the P-CeO2//PP separator exhibits exceptional electrochemical performance (a higher initial capacity of 1180 mAh/g at 0.5 C) and remarkable cycling stability with a low-capacity fading rate of only 0.048% per cycle over 1000 cycles at 2 C than that of bare CeO2//PP. In addition, the experimental studies have confirmed that the batteries consist of sulfur anode, lithium cathode and P-CeO2//PP separator exhibit superior adsorption capabilities for Li2S6, higher redox peak current, and facilitate the earlier precipitation of Li2S compared to the CeO2//PP batteries. To sum up, our research illustrates that the P-CeO2 porous separator coating material has the dual function of catalyzing polysulfide transformation and obstructing the polysulfide transport pathway, which can successfully mitigate the shuttle effect in LSBs. This study presents a practical method for improving the cycle stability and electrochemical performance of LSBs.

    Fig. S1 (Supporting information) schematically depicts the fabrication process of P-CeO2 modified separators and their subsequent integration into LSBs. Comprising three fundamental stages, the procedure includes: (1) Preparation of Ce-MOFs; (2) preparation of CeO2 by calcination of Ce-MOFs in air; and (3) preparation of P-CeO2 by phosphating CeO2 in argon atmosphere. Firstly, Ce-MOFs were synthesized through a straightforward hydrothermal method, wherein Ce3+ ions coordinate with trimeric acid molecules [46], and then Ce-MOFs were calcined in air to obtain CeO2. Afterwards, P was introduced by PH3 gas in argon atmosphere at high temperature, and P-CeO2 with a large amount of P doped was obtained after phosphorization process. After that, the modified separators were then assembled into button batteries by catalyzing the conversion of polysulfide intermediates and mitigating the shuttling effect of polysulfides to improve the electrochemical performance of LSBs. The morphologies of Ce-MOFs, CeO2 and P-CeO2 were observed by scanning electron microscopy (SEM) at different sizes (Figs. S2 and S3 in Supporting information, Figs. 1a and b). As depicted in Figs. S2a and b, the Ce-MOFs precursors exhibit a well-defined nanorod morphology with an average size of approximately 300 nm, which are densely arranged in a well-organized manner, and have a very smooth surface. After heat treatment at 500 ℃ for 4 h in air, CeO2 was obtained (Figs. S3a and b). During this process, the nanorod-like morphology and crystallinity of Ce-MOFs were intact even after heating at high temperatures for 4 h. After the phosphorization of CeO2 at high temperatures, P-CeO2 retains its nanorod-like structure (Figs. 1a and b), which is attributed to the complete coverage of the inert gas argon in a sealed tube furnace. It can be suggested that Ce-MOFs and CeO2 have excellent thermal stability during the heat treatment process.

    Figure 1

    Figure 1.  (a, b) SEM images of P-CeO2. (c) HAADF-STEM image of P-CeO2. (d) SAED pattern. (e, f) High-resolution TEM images and the corresponding EDX elemental mapping of (g) Ce, (h) O and (i) P.

    Analysis of transmission electron microscopy (TEM) images provided a detailed exploration of the morphology of P-CeO2. Fig. S5 (Supporting information) illustrates the uniform dispersion of P-CeO2 nanocrystals with the average size of approximately 10 nm, within the nanowires in TEM images. The images of high-resolution TEM (HRTEM) in Figs. 1e and f reveal distinct interplanar spacings of 0.191, 0.270, and 0.312 nm, corresponding to the (220), (200), and (111) lattice fringes of the CeO2 crystallites. Notably, these measurements are consistent with values derived from X-ray diffraction (XRD) data [47,48]. Fig. 1d showcases patterns of selected area electron diffraction (SAED), which clearly shows the presence of multiple rings, corresponding to the poly-crystalline property of CeO2. For additional verification of the elemental distribution of phosphorus within the composite, high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) images (Fig. 1c), along with the corresponding element mapping of Ce, O, and P, are presented in Fig. S4 (Supporting information) and Figs. 1gi. The mapping images further illustrate the uniform dispersion of phosphorus atoms in P-CeO2, implying that the crystal structure of CeO2 does not change after the doping of phosphorus. Obviously, the above properties demonstrated that the nanorod-like structure of P-CeO2, which provides fast electron transport channels and abundant catalytic reaction sites.

    X-ray diffraction (XRD) was employed to analyze the crystal structures and compositions of all samples. In Fig. 2a, P-CeO2 and CeO2 present closely matching XRD patterns, in which the prominent peaks at 28.5°, 33.1°, 47.4°, and 56.3° correspond precisely to the (111), (200), (220), and (311) faces of the CeO2 cubic phase (JCPDS No. 34-0394) with space group Fm-3m (225), respectively. The comparison of XRD peaks between CeO2 and P-CeO2 shows that P-CeO2 has no obvious peak-to-peak shift of XRD characteristics, which indicates that P doping does not change the crystal structure of CeO2. Afterwards, the decomposition temperature in nitrogen and air of Ce-MOFs was examined by the thermos gravimetric analysis (TGA). As shown in Fig. S6 (Supporting information), at 150 ℃, the mass loss of Ce-MOFs in air and nitrogen atmosphere are both 21.93%, when the temperature rises to 400 ℃, the mass loss of Ce-MOFs in air and nitrogen atmosphere are 64.23% and 25.41%, while when the temperature rises to 670 ℃, the mass loss are 64.99% and 52.36%, respectively. Exploration of the electronic states and detailed chemical composition of CeO2 and P-CeO2 were achieved through X-ray photoelectron spectra (XPS) analyses. Firstly, we calculated the contents of each element by XPS region scanning. The result shows that the atomic contents of O, Ce and P in CeO2 are ~75.7%, ~19.45% and ~0%, while that of O, Ce and P in P-CeO2 are ~76.9%, ~16.9% and ~10.2%, which means that we successfully doped P into CeO2 (Fig. 2b). The XPS spectra of P-CeO2 are showcased in Fig. S7 (Supporting information), distinctly illustrating the presence of C, O, P, and Ce elements in the samples. Additionally, the XPS spectra of P 2p show that P 2p was carefully deconvolved into P 2p1/2 and P 2p3/2, elucidating the presence of POx species (Fig. 2e). It is revealed that existing two characteristic spin-orbit states of Ce 3d5/2 and Ce 3d3/2 from the XPS of Ce 3d, as depicted in Figs. 2c and f. Notably, the Ce 3d3/2 peaks exhibit a subdivision into five components, of which three peaks attributed to Ce4+ (899.5, 905.8, and 915.6 eV) and two peaks attributed to Ce3+ (897.5 and 904.3 eV). High-resolution O 1s spectra provide insights into different oxygen species in P-CeO2. The peak at 531.0 eV belongs to adsorbed oxygen species and the peak at 529.2 eV is associated with metal-oxygen (Fig. 2d). Notably, the reversible conversion from Ce3+ to Ce4+ enhances the redox kinetics. The surface of P-CeO2 is rich in defective Ce3+/Ce, which indicates that the high concentration of oxygen vacancy is conducive to enhancing the chemisorption and catalytic conversion of LiPSs to accelerate the redox kinetics. These aspects will be investigated in subsequent electrochemical experiments.

    Figure 2

    Figure 2.  (a) XRD patterns of two samples. (b) The atomic element contents of CeO2 and P-CeO2, (c) XPS survey spectrum of (c) Ce 3d, (d) O 1s, (e) P 2p of P-CeO2 and (f) Ce 3d of CeO2. (g) SEM image of P-CeO2//PP. (h) Cross-sectional SEM image of P-CeO2//PP. (i) Contact angle of PP and P-CeO2//PP.

    The difference between commercial separators (PP) and P-CeO2 modified separators (P-CeO2//PP) in terms of physical properties is shown in Fig. 2. It can be observed that the original PP has a uniformly interwoven fishnet-like structure (Fig. S8 in Supporting information), which has a significant effect on the Li+ shuttling and inhibits the electron transport. Then, the uniform distribution of P-CeO2 nanorods on PP surface was observed through SEM image (Fig. 2g), and the diameter of P-CeO2//PP was about 19 mm (Fig. S9a in Supporting information). In addition, the thickness of the P-CeO2 modified layer was about 25 µm, which was confirmed by the SEM image of the membrane cross section (Fig. 2h). In addition, the mechanical toughness of the P-CeO2//PP separator is excellent and shows no delamination of P-CeO2 after various bending tests (Fig. S9b in Supporting information). Notably, we also performed electrolyte wettability tests by dropping electrolyte on the surfaces of both the P-CeO2//PP separator and the pristine PP separator and comparing the diffusion area of both (Fig. 2i). The contact angle of the P-CeO2//PP separator (7.01°) was found to be lower than that of the PP separator (33.34°), indicating that P-CeO2 has superior electrolyte wettability compared to PP. The presented findings unequivocally establish the superior qualities and promising potential of P-CeO2 as a modified diaphragm material for LSBs.

    In Fig. 3a, cyclic voltammetry (CV) curves of P-CeO2//PP, CeO2//PP, and PP batteries are presented at a uniform scan rate of 0.1 mV/s. The battery utilizing a P-CeO2//PP displays two prominent cathodic peaks at approximately 2.3 V compared to 2.0 V and a noticeable anodic peak at approximately 2.4 V. During the process of cathodic scanning, both high and low potential regions display two reduction peaks that correspond to the conversion of S8 to LiPSs and its therewith transformation into Li2S/Li2S2, respectively, and the two reduction peaks correspond to the conversion of Li2S2 to LiPSs and ultimately to S8 during the anodic scanning. Notably, the oxidation and reduction peaks of cells utilizing a P-CeO2//PP exhibit higher negative and positive shifts, respectively. This suggests faster redox kinetics and efficient utilization of sulfur. Figs. 3b and d and Fig. S11a (Supporting information) show that the CV curves of P-CeO2//PP, CeO2//PP and PP batteries at different scan rates from 0.1 mV/s to 0.5 mV/s, CV curves have no obvious deformation at different scanning rates, which proves its excellent stability. To research the ion transport velocity about different separator, the peak current curves are shown in Figs. 3c and e and Fig. S11b (Supporting information), the comparison results of linear relationships between three redox current peaks for P-CeO2, CeO2 and PP batteries are shown in Fig. S12 (Supporting information). Significantly, P-CeO2 demonstrated enhanced velocity in comparison to CeO2, emphasizing the exceptional electrical conductivity inherent in the P-CeO2 bimetallic composition as opposed to that of CeO2. The CV curves of symmetric batteries were depicted in Fig. 3f. Concurrently, in comparison to CeO2//PP and PP, P-CeO2//PP demonstrated a notable peak current density. The EIS plot for symmetric batteries, utilizing CeO2 and P-CeO2 as identical electrodes, is presented in Fig. 3g. It is an evident that the charge transfer resistance of the P-CeO2//PP symmetric batterie is lower than that of the CeO2 sample, suggesting enhanced charge-transfer capability and heightened electrocatalytic characteristics. Simultaneously, experiments involving cells with Li2S8 electrolytes were conducted to achieve nucleation experiments, thus corroborating the catalytic effectiveness of P-CeO2 towards LiPSs. The nucleation experiments for both P-CeO2 and CeO2 are graphically depicted in Figs. 3h and i. From the experimental arrangement, the P-CeO2 material played the role of the cathode, lithium served as the anode and Li2S8 was chosen as electrolyte, and a polypropylene separator was integrated into the battery assembly. In the initial phase of the experiment, in order to achieve the full conversion of Li2S8 to Li2S4, all batteries were continuously discharged to 2.05 V at a current of 0.112 mA. Following this, to foster the nucleation of Li2S, an electrostatic discharge was performed until the current dropped below 10–5 at 2.05 V. This procedure aligns with the methodologies outlined in the previous works by Yang et al., along with the equation correlating time and peak current [49]:

    $ t_m=\left(\frac{2}{\pi A k^2}\right)^{\frac{1}{3}} $

    (1)

    Figure 3

    Figure 3.  (a) CV curves of P-CeO2//PP, CeO2//PP, and PP batteries at a scan rate of 0.1 mV/s. (b, d) CV curves of P-CeO2//PP and CeO2//PP batteries at different scan rates. (c, e) Comparison of linear relationships between current peak (Ip) and the square root of the scan rates for P-CeO2//PP and CeO2//PP batteries. (f) CV curves of symmetric batteries at 10 mV/s under the Li2S6 electrolyte. (g) EIS spectra of symmetric batteries. (h, i) Li2S nucleation tests for evaluating the nucleation kinetics based on different electrodes.

    In the above formula, the peak current time is expressed in seconds, where A denotes the nucleation rate constant (cm–2 s–1), and k signifies the growth rate (cm/s). Manifestly, the results show that after data fitting, P-CeO2 manifests higher peak current, larger deposition peak area, and shorter nucleation time than CeO2, which further proves the effectiveness of phosphorus-doped metal oxides in promoting the catalytic conversion of polysulfides.

    To comprehensively investigate the performance of P-CeO2//PP LSBs, a lot of electrochemical tests were executed. Fig. 4a portrays that the P-CeO2//PP battery (with a sulfur load of 1.28 mg) obtains excellent performance of long cycle at 0.5 C. Significantly, it is obvious that P-CeO2//PP demonstrates a substantial initial discharge-specific capacity of 1180 mAh/g at 0.5 C, accompanied by an outstanding decay rate of 0.08% per cycle over 500 cycles. Conversely, cells assembled with other separators displayed lower decay rates of 0.147% and 0.154%, which highlights the exceptional cycle stability of the P-CeO2//PP separator. Additionally, charge/discharge curves of the first cycle at 0.5 C were depicted in Fig. 4b and Fig. S8 (Supporting information), revealing that P-CeO2 maintains a small polarization voltage when compared to CeO2 and the PP separator. This serves as robust evidence that the effectiveness of phosphorus-doped metal oxides in promoting catalytic conversion and reaction kinetics. Furthermore, P-CeO2//PP batteries sustained a good long cycle performance with an attenuation rate of 0.08% even at 1 C after 800 cycles, highlighting the advantages associated with ultrafast reaction kinetics (Fig. 4c). Next, in order to explore the ionic conductivity of the three batteries directly, EIS tests were performed in Fig. 4d, which indicate that the P-CeO2//PP cell displays the lowest resistance, signifying its superior conductivity and reinforcing the earlier point. Moreover, in order to explore the cyclic stability of P-CeO2//PP batteries under high sulfur loading, P-CeO2//PP batteries under sulfur loads of 2.0 and 3.0 mg at 0.5 C were depicted in Fig. 4e. It is worth noting that as the sulfur load increases, the battery consistently maintains a high specific capacity, and after 300 cycles, the coulomb efficiency is still approach 100%. Fig. 4f clearly shown the superior rate capability of P-CeO2//PP batteries, arising from the electrochemical catalysis. When the current density varies from 0.2, 0.5, 1, 2 C to 3 C, the discharge capacity shows a steady decay of 1210, 1080, 890, 788 mAh/g to 648 mAh/g, which illustrates that the P-CeO2 modified separator expedites the reaction kinetics and impedes the shuttle effect of polysulfide efficiently. Furthermore, in comparison to previous works, the P-CeO2//PP configuration also demonstrates exceptional performance, offering a valuable strategy for enhancing the performance of LSBs.

    Figure 4

    Figure 4.  (a) Cycling performances of LSBs with various separators at 0.5 C. (b) Charge-discharge profiles at 0.5 C. (c) Cycling performances of LSBs with various separators at 1 C. (d) The EIS spectra of LSBs with various separators. (e) Cycle performance with the high sulfur loading of 2.0 and 3.0 mg/cm2 of P-CeO2//PP batteries at 0.5 C. (f) Rate capacities of LSBs with various separators.

    Adsorption experiments were conducted to investigate the adsorption mechanism (Figs. 5a and b). Based on the observed color change, successful adsorption of Li2S6 was achieved. Furthermore, as shown in Fig. 5c, XPS analysis was performed on P-CeO2 powder after 24 h adsorption. Two adsorption peaks for Ce3+ and three adsorption peaks for Ce4+ shift from the initial 898.3, 903.5, 900.6, 906.9 and 916.4 eV to 896.5, 901.4, 898.6, 904.7 and 915.1 eV. Thus, the material exhibits both chemical and physical interactions with Li2S6, providing compelling evidence that P-CeO2 can promote the chemisorption and catalytic conversion of polysulfide. Simultaneously, the visual representation in Fig. 5d illustrates the impact of different separators on polysulfide through shuttle experiments. The PP and CeO2//PP separators exhibit a noticeable shuttle effect of polysulfide, while the P-CeO2//PP separator shows superior performance to mitigate the effect. From the experiment, we can see that the shuttle effect of polysulfide is most significant in PP separator, followed by CeO2//PP separator, and P-CeO2//PP separator has the best effect to outcome the shuttle of polysulfide. This serves as direct evidence confirming the effective inhibition of polysulfide shuttling by the modified separator. Besides, the calculation results presented the adsorption energy of the (100) crystal plane for P-CeO2 on Li2S6 can reach –10.44 eV (Fig. S13 in Supporting information), which is higher than that of CeO2, presenting an outstanding adsorption effect on Li2S6 and suggesting that the doping of P can promote the electron transfer and can effectively slow down the shuttle effect during the charge/discharge processes. Additionally, Raman shifts were utilized to identify surface defects in P-CeO2 and CeO2. Notably, the peak observed at 463.7 cm–1 represents the symmetric stretching vibration of the CeO2 lattice, while the faint peak approximately 600 cm–1 is due to the presence of Ce3+ (Fig. S14 in Supporting information).

    Figure 5

    Figure 5.  (a, b) Adsorption test of P-CeO2 in Li2S6 electrolyte. (c) XPS spectrum of P-CeO2 after adsorption of Li2S6. (d) Shuttling effect measurements of three separators at the Li2S6 electrolyte.

    In summary, the phosphorus doped porous CeO2 (P-CeO2) were successfully synthesized by solvothermal method, calcination and phosphorization, then the commercial PP separator is modified by coating process. The P-CeO2 modified separator surpasses the commercial PP separator due to the synergistic effects in P-CeO2, which combines the catalytic conversion and chemical absorption for polysulfides. Turn out, when the sulfur load is 1.28 mg/cm2, the battery capacity is 1180 mA/h, and the decay rate per cycle is 0.10% at 0.5 C. When the sulfur load is increased to 2.0 mg/cm2, the decay rate per cycle is 0.048% at 0.5 C. These performances are superior to previously reported separators modified by cerium-based compounds [50]. Furthermore, the P-CeO2//PP battery maintained a high capacity of 658 mAh/g under a high sulfur loading of 3.0 mg/cm2 after 300 cycles. In this study, a strategy of phosphorus-doped porous CeO2 is introduced, which shows that P-CeO2 separator can promote the chemical absorption and catalytic conversion of polysulfides in advanced LSBs, thus achieving high performance and long cycle life.

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

    Xinyun Liu: Writing – original draft, Investigation, Conceptualization. Long Yuan: Data curation. Xiaoli Peng: Formal analysis. Shilan Li: Methodology. Shengdong Jing: Software. Shengjun Lu: Conceptualization. Hua Lei: Writing – original draft, Supervision. Yufei Zhang: Writing – review & editing, Supervision. Haosen Fan: Writing – review & editing, Supervision.

    This work was supported by National Natural Science Foundation of China (Nos. 52472194, 52101243), Natural Science Foundation of Guangdong Province, China (No. 2023A1515012619) and the Science and Technology Planning Project of Guangzhou (No. 202201010565). We would like to thank Analysis and Test Center of Guangzhou University for their technical support.

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2024.110369.


    1. [1]

      G. Zhou, H. Chen, Y. Cui, Nat. Energy 7 (2022) 312-319. doi: 10.1038/s41560-022-01001-0

    2. [2]

      B. Dunn, H. Kamath, J.M. Tarascon, Science 334 (2011) 928-935. doi: 10.1126/science.1212741

    3. [3]

      S. J. Lu, J. Y. Lin, C. H. Wang, et al., Rare Met. 43 (2024) 3713-3723. doi: 10.1007/s12598-024-02650-8

    4. [4]

      L. Chen, Z. Liu, W. Yang, et al., J. Colloid Interface Sci. 666 (2024) 416-423. doi: 10.1016/j.jcis.2024.04.044

    5. [5]

      Y. Tsao, H. Gong, S. Chen, et al., Adv. Energy Mater. 11 (2021) 2101449. doi: 10.1002/aenm.202101449

    6. [6]

      B. Qin, M. Wang, S. Wu, et al., Chin. Chem. Lett. 35 (2024) 108921. doi: 10.1016/j.cclet.2023.108921

    7. [7]

      S. M. Wu, F. Xu, Y. N. Li, et al., J. Colloid Interface Sci. 649 (2023) 741-749. doi: 10.1016/j.jcis.2023.06.159

    8. [8]

      R. Sun, F. Xu, C.H. Wang, et al., Rare Met. 43 (2024) 1906-1931. doi: 10.1007/s12598-023-02563-y

    9. [9]

      F. Xu, S.L. Li, S.D. Jing, et al., J. Colloid Interface Sci. 660 (2024) 907-915. doi: 10.1016/j.jcis.2024.01.138

    10. [10]

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

    11. [11]

      J. Shen, X. Xu, J. Liu, et al., Adv. Energy Mater. 11 (2021) 2100673. doi: 10.1002/aenm.202100673

    12. [12]

      C. Qi, L. Xu, J. Wang, et al., ACS Sustain. Chem. Eng. 8 (2020) 12968-12975. doi: 10.1021/acssuschemeng.0c03536

    13. [13]

      J. Xu, S. An, X. Song, et al., Adv. Mater. 33 (2021) 2105178. doi: 10.1002/adma.202105178

    14. [14]

      W. Yang, W. Yang, B. Sun, et al., ACS Appl. Mater. Interfaces 10 (2018) 39695-39704. doi: 10.1021/acsami.8b14045

    15. [15]

      L. Tan, X. Li, Z. Wang, H. Guo, J. Wang, ACS Appl. Mater. Interfaces 10 (2018) 3707-3713. doi: 10.1021/acsami.7b18645

    16. [16]

      N. Deng, W. Kang, Y. Liu, et al., J. Power Source. 331 (2016) 132-155. doi: 10.1016/j.jpowsour.2016.09.044

    17. [17]

      Z. Li, L. Tang, X. Liu, et al., Electrochim. Acta 310 (2019) 1-12. doi: 10.1016/j.electacta.2019.04.057

    18. [18]

      J. Song, C. Zhang, X. Guo, et al., J. Mater. Chem. A 6 (2018) 16610-16616. doi: 10.1039/c8ta04800b

    19. [19]

      W. Xue, D. Yu, L. Suo, et al., Matter 1 (2019) 1047-1060. doi: 10.1016/j.matt.2019.07.002

    20. [20]

      X.Y. Jin, S.Y. Dong, X.C. Guo, et al., Appl. Surf. Sci. 686 (2025) 162181. doi: 10.1016/j.apsusc.2024.162181

    21. [21]

      F. Wu, S. Zhao, L. Chen, et al., Energy Storage Mater. 14 (2018) 383-391. doi: 10.1016/j.ensm.2018.06.009

    22. [22]

      S. Deng, Y. Yan, L. Wei, et al., ACS Appl. Energy Mater. 2 (2019) 1266-1273. doi: 10.1021/acsaem.8b01815

    23. [23]

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

    24. [24]

      J. K. Nørskov, T. Bligaard, J. Rossmeisl, C. H. Christensen, Nat. Chem. 1 (2009) 37-46. doi: 10.1038/nchem.121

    25. [25]

      P. Guo, K. Sun, X. Shang, et al., Small 15 (2019) 1902363. doi: 10.1002/smll.201902363

    26. [26]

      X.C. Guo, P.F. Wan, P. Xia, et al., J. Colloid Interface Sci. 678 (2025) 393–406.

    27. [27]

      X. Zhang, G. Li, Y. Zhang, et al., Nano Energy 86 (2021) 106094. doi: 10.1016/j.nanoen.2021.106094

    28. [28]

      C. Ma, Y. Zhang, Y. Feng, et al., Adv. Mater. 33 (2021) 2100171. doi: 10.1002/adma.202100171

    29. [29]

      L. Ma, R. Chen, G. Zhu, et al., ACS Nano 11 (2017) 7274-7283. doi: 10.1021/acsnano.7b03227

    30. [30]

      J. Xu, Q. Zhang, X. Liang, et al., Nanoscale 12 (2020) 6832-6843. doi: 10.1039/d0nr00160k

    31. [31]

      N. Deng, J. Ju, J. Yan, et al., ACS Appl. Mater. Interfaces 10 (2018) 12626-12638. doi: 10.1021/acsami.7b19746

    32. [32]

      L. Peng, Z. Yu, M. Zhang, et al., Nanoscale 13 (2021) 16696-16704. doi: 10.1039/d1nr04855d

    33. [33]

      B. Liu, S. Gu, H. Li, et al., Mater. Today Energy 17 (2020) 100484. doi: 10.1016/j.mtener.2020.100484

    34. [34]

      Y. Huang, X. Zhai, T. Ma, et al., Coord. Chem. Rev. 450 (2022) 214236. doi: 10.1016/j.ccr.2021.214236

    35. [35]

      Y. Xiao, H. Li, K. Xie, Angew. Chem. 133 (2021) 5300-5304. doi: 10.1002/ange.202013633

    36. [36]

      J. Wang, W.Q. Han, Adv. Funct. Mater. 32 (2022) 2107166. doi: 10.1002/adfm.202107166

    37. [37]

      S. Li, B. Chen, Y. Wang, et al., Nat. Mater. 20 (2021) 1240-1247. doi: 10.1038/s41563-021-01006-2

    38. [38]

      C. Yan, G. Chen, X. Zhou, J. Sun, C. Lv, Adv. Funct. Mater. 26 (2016) 1428-1436. doi: 10.1002/adfm.201504695

    39. [39]

      H. Jin, X. Liu, S. Chen, et al., ACS Energy Lett. 4 (2019) 805-810. doi: 10.1021/acsenergylett.9b00348

    40. [40]

      D. Voiry, H. Yamaguchi, J. Li, et al., Nat. Mater. 12 (2013) 850-855. doi: 10.1038/nmat3700

    41. [41]

      W. Liu, H. Zhi, X. Yu, Energy Storage Mater. 16 (2019) 290-322. doi: 10.1016/j.ensm.2018.05.020

    42. [42]

      X. Li, H. Liang, B. Qin, et al., J. Colloid Interface Sci. 625 (2022) 41-49. doi: 10.1016/j.jcis.2022.05.155

    43. [43]

      L. Wang, Z. Wang, L. Xie, L. Zhu, X. Cao, ACS Appl. Mater. Interfaces 11 (2019) 16619-16628. doi: 10.1021/acsami.9b03365

    44. [44]

      S. Chen, K. Li, F. Zhao, et al., Nat. Commun. 7 (2016) 13169. doi: 10.1038/ncomms13169

    45. [45]

      L. Ma, R. Chen, G. Zhu, et al., ACS Nano 11 (2017) 7274-7283. doi: 10.1021/acsnano.7b03227

    46. [46]

      C. Du, X. Gao, C. Cheng, et al., Electrochim. Acta 266 (2018) 348-356. doi: 10.1016/j.electacta.2018.02.035

    47. [47]

      W. Dong, Y. Huang, Microchim. Acta 187 (2020) 1-10. doi: 10.1155/2020/3691421

    48. [48]

      W. Qi, W. Jiang, F. Xu, et al., Chem. Eng. J. 382 (2020) 122852. doi: 10.1016/j.cej.2019.122852

    49. [49]

      P.F. Wan, X.C. Guo, S.D. Jing, et al., Chem. Eng. J. 504 (2025) 159014. doi: 10.1016/j.cej.2024.159014

    50. [50]

      Y. Liu, X. Zhao, S. Li, et al., Chem Res. Chin. Univ. 38 (2022) 147-154. doi: 10.1007/s40242-021-1386-x

  • Figure 1  (a, b) SEM images of P-CeO2. (c) HAADF-STEM image of P-CeO2. (d) SAED pattern. (e, f) High-resolution TEM images and the corresponding EDX elemental mapping of (g) Ce, (h) O and (i) P.

    Figure 2  (a) XRD patterns of two samples. (b) The atomic element contents of CeO2 and P-CeO2, (c) XPS survey spectrum of (c) Ce 3d, (d) O 1s, (e) P 2p of P-CeO2 and (f) Ce 3d of CeO2. (g) SEM image of P-CeO2//PP. (h) Cross-sectional SEM image of P-CeO2//PP. (i) Contact angle of PP and P-CeO2//PP.

    Figure 3  (a) CV curves of P-CeO2//PP, CeO2//PP, and PP batteries at a scan rate of 0.1 mV/s. (b, d) CV curves of P-CeO2//PP and CeO2//PP batteries at different scan rates. (c, e) Comparison of linear relationships between current peak (Ip) and the square root of the scan rates for P-CeO2//PP and CeO2//PP batteries. (f) CV curves of symmetric batteries at 10 mV/s under the Li2S6 electrolyte. (g) EIS spectra of symmetric batteries. (h, i) Li2S nucleation tests for evaluating the nucleation kinetics based on different electrodes.

    Figure 4  (a) Cycling performances of LSBs with various separators at 0.5 C. (b) Charge-discharge profiles at 0.5 C. (c) Cycling performances of LSBs with various separators at 1 C. (d) The EIS spectra of LSBs with various separators. (e) Cycle performance with the high sulfur loading of 2.0 and 3.0 mg/cm2 of P-CeO2//PP batteries at 0.5 C. (f) Rate capacities of LSBs with various separators.

    Figure 5  (a, b) Adsorption test of P-CeO2 in Li2S6 electrolyte. (c) XPS spectrum of P-CeO2 after adsorption of Li2S6. (d) Shuttling effect measurements of three separators at the Li2S6 electrolyte.

  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  95
  • HTML全文浏览量:  0
文章相关
  • 发布日期:  2025-10-15
  • 收稿日期:  2024-06-25
  • 接受日期:  2024-08-24
  • 修回日期:  2024-08-07
  • 网络出版日期:  2024-08-28
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

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

/

返回文章