Open Access Article
Zhe Yang†
a,
Shenglin Liu†c,
Kai Chenb,
Guodong Zhang*b,
Feng Gong
*c,
Shuangxi Xing
*a and
Jian Wang
*de
aFaculty of Chemistry, Northeast Normal University, Changchun 130024, P. R. China. E-mail: xingsx737@nenu.edu.cn
bDepartment of Physics, Research Institute for Biomimetics and Soft Matter, Fujian Provincial Key Laboratory for Soft Functional Materials, Xiamen University, Xiamen 361005, P. R. China. E-mail: zhanggd710@nenu.edu.cn
cKey Laboratory of Energy Thermal Conversion and Control of Ministry of Education, School of Energy and Environment, Southeast University, Nanjing, 211189, Jiangsu, China. E-mail: gongfeng@seu.edu.cn
dKarlsruhe Institute of Technology (KIT), D76021 Karlsruhe, Germany. E-mail: jian.wang@kit.edu; wangjian2014@sinano.ac.cn
eHelmholtz Institute Ulm (HIU)D89081 Ulm, Germany
First published on 1st December 2025
High-energy-density lithium sulfur batteries with high mass loading are restricted by the depressive electrochemical kinetics of polysulfide conversion. Herein, to enhance catalytic efficiency, abundant electron-delocalized CeO2 nanoparticles are anchored on the surface of pollen-derived carbon (PC-CeO2) via one-step carbonization, serving as a sulfur host. In this design, pollen-derived carbon (PC) with a porous structural network enhances the electrical conductivity of the sulfur cathode while alleviating volume expansion and maintaining the stability of the cathode. The strategic incorporation of electron-delocalized CeO2 nanoparticles is beneficial for the adsorption and catalysis of polysulfides, limiting the shuttle effect of polysulfides and effectively facilitating the electrochemical conversion kinetics. As a result, the fabricated sulfur cathode (PC-CeO2/S) exhibits excellent electrochemical stability with a decay rate per cycle of 0.054% after 1000 cycles at 1C and outstanding rate performance (703.3 mAh g−1 at 3C). Furthermore, it achieves an impressive areal capacity of 5.64 mAh cm−2 at 0.2C even with a high sulfur loading of 5.5 mg cm−2, demonstrating its potential for practical, high-energy-density applications in lithium–sulfur batteries.
Researchers have made many attempts to design and synthesize advanced sulfur host materials to solve the above problems in the sulfur cathode.12–15 Carbon materials, including carbon nanotubes, graphene and conductive carbon black, are widely used as sulfur host materials16,17 because of their high electronic conductivity, good compatibility with sulfur and excellent chemical stability.18–20 Nevertheless, non-polar carbon materials find it difficult to inhibit the shuttle effect effectively because of their weak interactions with LiPSs, ultimately leading to severe capacity decay.21–23 The strategic doping modification of carbon-based materials is a useful approach for improving the adsorption capacity. For example, the incorporation of non-metallic heteroatoms, such as N and O, into the carbon matrix not only enhances its intrinsic electronic conductivity but also significantly increases its polarity, greatly improving its adsorption capacity and mitigating the shuttle effect.24–27 Furthermore, researchers have frequently integrated polar metal compounds with carbon-based frameworks to obtain advanced composite materials, aiming to synergistically enhance the chemical adsorption and catalytic ability of LiPSs.28–30 This strategic modification improves the conductivity of carbon while introducing strong polar interactions that effectively adsorb and inhibit LiPSs. Simultaneously, the catalytic activity of these metal compounds accelerates the redox kinetics of LiPS conversion, which mitigates the sluggish reaction rates and suppresses the shuttle effect, thereby optimizing the overall electrochemical performance of lithium–sulfur batteries.31–34
As a carbon precursor, pollen-derived carbon has the following advantages: (1) intrinsic elemental composition of pollen: owing to its richness in various non-metallic elements, PC undergoes direct in situ modification during carbonization, which produces a conductive and polar carbon matrix without additional doping steps. (2) Hierarchical porosity and structural stability: the naturally self-organized porous network of pollen-derived carbon facilitates efficient sulfur loading while simultaneously alleviating the volumetric expansion of sulfur during lithiation and maintaining structural stability. (3) Sustainable and green chemistry approach: as an abundant and renewable biomass material, PC avoids special and complex synthesis routes, reflecting the advantages of green chemistry.35,36 In order to further improve the adsorption and catalytic performance of carbon materials, the incorporation of metallic compounds is often necessary to form an active interface. Among the available materials, cerium dioxide (CeO2) has received significant attention due to its extraordinary properties:14,37,38 (1) CeO2 possesses a unique Ce3+/Ce4+ redox couple, which actively participates in accelerating the reversible catalytic conversion of LiPSs. (2) Abundant oxygen vacancies in CeO2 serve as active sites to enhance the chemisorption and catalytic conversion of LiPSs.
In this study, to enhance catalytic efficiency, abundant electron-delocalized CeO2 nanoparticles anchored on the surface of pollen-derived carbon materials (PC-CeO2) are prepared through a carbonization process. The obtained PC-CeO2 reduced the volume expansion, which maintained the stability and integrity of the structure due to its special and robust structure. Meanwhile, it provided sufficient active sites of electron delocalization for the chemical adsorption and catalytic kinetic conversion of LiPSs. DFT theoretical calculations further prove the strong adsorption and catalytic capacity of PC-CeO2 towards sulfur species. As a result, the PC-CeO2/S cathode demonstrated outstanding electrochemical performance with a low fading rate of 0.054% per cycle at 1C after 1000 cycles and an outstanding rate capacity of 703.3 mAh g−1 at 3C. Even with a high sulfur loading of 5.5 mg cm−2, the PC-CeO2/S electrode exhibited a high capacity of 5.64 mAh cm−2 at 0.2C after 60 cycles, which reached commercial standards of 4.0 mAh cm−2.
To further confirm the composition and structural characteristics of the material, the relevant characterization results are displayed in Fig. 1f–j. In the XRD patterns of PC-CeO2 and CeO2 (Fig. 1f), the diffraction peaks at 28.5°, 33.1°, 47.5° and 56.3° are attributed to the (111), (200), (220) and (311) planes of CeO2 (PDF#34-0394), respectively, while the XRD pattern of PC shows two distinct carbon peaks.40 To investigate the configuration of carbon in PC-CeO2 and PC, the Raman spectrum was carried out, as shown in Fig. 1g. Two prominent peaks emerge at 1350 and 1580 cm−1, corresponding to the D and G bands of the carbon layer, respectively. Besides, the intensity ratios (ID/IG) for PC-CeO2 and PC are estimated to be 0.98 and 0.96, respectively. This marked difference underscores more defects exhibited by the PC-CeO2 composite, which plays a pivotal role in optimizing the rapid transport of electron ions during electrochemical process, thereby improving the electrochemical performance. In addition, XPS measurements were carried out to determine the surface chemistry phase and composition of the PC-CeO2 composite. The XPS full spectrum of PC-CeO2 verifies the presence of Ce, O, N and C elements (Fig. S6a). As depicted in Fig. S6b, the C 1s spectrum displays four distinct peaks at 284.8 eV (C–C), 286.1 eV (C–N), 287.2 eV (C–O) and 289.2 eV (C
O). Fig. S6c shows three peaks, including pyridinic-N (398.4 eV), pyrrolic-N (400.7 eV) and graphitic-N (401.7 eV). As shown in Fig. 1h, the Ce 3d spectrum exhibits eight distinct splitting peaks. The peaks at 917.1, 908.2, 901.2, 898.6, 888.6, and 882.7 eV can be ascribed to Ce4+, while others located at 904.1 and 885.7 eV belong to Ce3+. In addition, the oxygen vacancy of PC-CeO2 can be verified by O1s XPS (Fig. 1i). The peaks at 529.5 eV are connected to metal oxygen, and the peaks at 531.5 eV are attributed to a low-coordinated oxygen vacancy.41 EPR is used to further estimate the oxygen vacancy of PC-CeO2 (Fig. 1j). The representative EPR signal at g = 2.002 reveals electron capture at the oxygen vacancy.42,43 The above results explain that the addition of CeO2 NPs plays a pivotal role in promoting chemical adsorption and accelerating the catalytic conversion of LiPSs.42,44 The sulfur contents in PC-CeO2/S, CeO2/S and PC/S are 75, 71.2 and 74.2 wt%, respectively (Fig. S7). Besides, the EDS elemental mapping of the PC-CeO2/S in Fig. S8 confirms the uniform distribution of sulfur. These findings demonstrate the remarkable sulfur-holding ability of the PC-CeO2 composite.
To further demonstrate the adsorption capacity of different materials for LiPSs, visual adsorption experiments were carried out. As illustrated in Fig. 2a, the color in Li2S6 with PC-CeO2 nearly vanishes after 2 h, and the UV-vis data further confirm its exceptional adsorption ability. Furthermore, XPS analysis of LiPS adsorption was performed to analyze the possible mechanism of the strong anchoring effect for LiPSs on PC-CeO2, as shown in Fig. 2b and c. After exposure to Li2S6, the Ce 3d peaks exhibit an obvious shift toward lower binding energies, indicating that Ce cations actively participate in the redox interactions. This shift confirms the involvement of the Ce4+/Ce3+ redox couple, highlighting their crucial role as electron transfer mediators during the Li2S6 adsorption process. As depicted in Fig. 2c, the peaks at 165.2 and 164.0 eV correspond to S22−, and the signals at 168.4 and 169.4 eV correspond to SO32− and SO42− (thiosulfate), respectively.44 All the above data indicate a strong interaction between PC-CeO2 and Li2S6. Li2S plays a crucial role in the reaction process of lithium–sulfur batteries, affecting the charge and discharge capacity of the batteries. Thus, nucleation (Fig. 2d–f) and dissolution experiments of Li2S (Fig. 2g–i) were carried out. The results indicate that the PC-CeO2 electrode exhibits the earliest current peak at 2130 s and the largest deposition capacity at 189.98 mAh g−1. Besides, as shown in Fig. 2g–i, the dissolution profiles of Li2S display that the PC-CeO2 electrode exhibits the earliest current response and largest current peak (829.9 s/0.498 mA). Therefore, PC-CeO2 significantly enhances liquid–solid reactions and accelerates the conversion of LiPSs in Li–S batteries.
The cyclic voltammetry (CV) profiles for Li–S batteries equipped with PC-CeO2, CeO2 and PC electrodes are depicted in Fig. 3a. Two notable reduction peaks, labeled peaks A and B, appear around 2.30 V and 2.02 V, respectively. These correspond to the stepwise reduction process of S8 to lithium polysulfides (Li2Sn, 4 ≤ n ≤ 8) and their subsequent transformation into shorter, insoluble species, like Li2S2 and Li2S. The oxidation peak, denoted as peak C at 2.31 V, marks the reverse reaction, where Li2S2 and Li2S are oxidized back to S8. Among these, the PC-CeO2 electrode stands out with the highest current response and the smallest reaction polarization (0.295 V for PC-CeO2 vs. 0.314 V for CeO2 and 0.358 V for PC), signifying its superior electrochemical performance in Li–S batteries. Fig. 3b and c illustrate the enlarged CV curve segments around peaks A and C in Fig. 3a. Tafel plots were derived based on these peaks, revealing that the PC-CeO2 electrode exhibits lower Tafel slopes (27.3 mV dec−1 for peak A and 98.91 mV dec−1 for peak C) than the CeO2 electrode (28.2 mV dec−1 and 124.53 mV dec−1) and PC electrode (54.9 mV dec−1 and 111.38 mV dec−1). This suggests the fastest electron transfer rate of PC-CeO2 during the redox process of LiPSs. To verify the catalytic capabilities of the different electrodes in facilitating LiPS conversion, symmetric cells were assembled. Fig. 3d shows that the PC-CeO2 electrode demonstrates the highest current density and the largest curve area, underscoring its excellent catalytic activity in the conversion of LiPSs. In addition, galvanostatic intermittent titration technique (GITT) profiles were measured at 0.1C to explore the kinetic performance of the host materials (Fig. 3e and S9a–f). Two important transition points about Li2S nucleation and activation are shown in Fig. 3e, and the voltage difference at Li2S nucleation and Li2S activation points can be calculated (Fig. 3f). The cell with the PC-CeO2 electrode shows the lowest Li2S nucleation voltage difference (22.8 mV) and Li2S activation voltage difference (79.1 mV), indicating its lower polarization and faster kinetics. An EIS test was conducted to analyze the redox kinetics (Fig. 3g). The resulting Nyquist plots are presented in Fig. 3g. Notably, the PC-CeO2 cathode exhibits a relatively low interface resistance (Rs) of 2.72 Ω, compared to 4.74 Ω for the CeO2 cathode and 2.78 Ω for the PC cathode. Additionally, the PC-CeO2 cathode displays a relatively small charge transfer resistance (Rct) of 63.9 Ω, in contrast to 81.5 Ω for the CeO2 and a significantly high value of 167.5 Ω for the PC cathode (shown in Table S2). These results indicate that PC-CeO2 facilitates the rapid kinetics conversion of LiPSs in Li–S batteries. Furthermore, the CV curves of Li–S cells with different cathodes were recorded over a scan range of 0.1–0.5 mV s−1 (Fig. 3h, S10a and S10b). The CV tests were conducted 15 times from 0.1 to 0.5 mV s−1, and the CV curves at the same scanning speed are very close, as depicted in Fig. S11, demonstrating the exceptional electrochemical stability and reversibility of the PC-CeO2 electrode. The peak current is plotted against the square root of the scan rate, as shown in Fig. 4i. According to the classical Randles–Sevcik equation (IP = 2.69 × 105n1.5ADLi+0.5CLi+ν0.5), the Li+ diffusion coefficients are calculated. For the PC-CeO2 electrode, the Li+ diffusion rates are found to be DLi+ (A) = 2.35 × 10−8 cm2 s−1, DLi+ (B) = 4.9 × 10−8 cm2 s−1 and DLi+ (C) = 5.21 × 10−8 cm2 s−1, which are higher than those of the CeO2 and PC electrodes (Fig. S12). These results demonstrate that the PC-CeO2 electrode delivers the highest Li+ diffusion coefficients, leading to faster electrochemical reactions and enhanced performance in Li–S batteries.
Li–S batteries were assembled using PC-CeO2, CeO2 and PC electrodes paired with Li anodes. The electrochemical results are summarized in Fig. 4. The charge/discharge profiles show two distinct discharge plateaus and one charge plateau at 0.2C (Fig. 4a), consistent with the CV curve data. Besides, the PC-CeO2 electrode exhibits a smaller polarization potential (ΔE = 182 mV) than the CeO2 electrode (ΔE = 278 mV) and PC electrode (ΔE = 207 mV), revealing its enhanced redox kinetics and reversibility in Li–S batteries (Fig. 4b). Moreover, the capacity ratio between the two discharge plateaus (Q2 and Q1) serves as a crucial indicator of catalytic activity. Q1 corresponds to sulfur reduction to soluble LiPSs, while Q2 reflects their conversion into Li2S. The PC-CeO2 electrode shows the highest Q2/Q1 ratio of 2.17 (vs. 2.07 for CeO2 and 2.15 for PC), suggesting more efficient sulfur utilization and superior catalytic activity. The cycling performance of the electrodes at 0.2C is depicted in Fig. S13, where the PC-CeO2 cathode displays an initial discharge capacity of 1174 mAh g−1, with a coulombic efficiency (CE) consistently above 99.8% over 200 cycles, surpassing both the CeO2 and PC cathodes.
As shown in Fig. 4c, the rate capabilities were further evaluated from 0.2 to 3C. The PC-CeO2 electrode demonstrates a remarkable rate performance, maintaining initial discharge capacities of 1024.4, 891.5, 829.1, 744.8, and 703.3 mAh g−1 at 0.2, 0.5, 1, 2, and 3C, respectively. Even when the current density is reversed back to lower rates (2, 1, 0.5, and 0.2C), the discharge capacities remain at 756.8, 827.3, 883.2, and 957.8 mAh g−1, achieving capacity retention rates of 101.6% at 2C, 99.7% at 1C, 99.1% at 0.5C, and 93.5% at 0.2C, respectively. The corresponding galvanostatic discharge/charge voltage curves of PC-CeO2, CeO2 and PC at 0.2–3 C are presented in Fig. S14a–c. The PC-CeO2 electrode maintains a stable discharge/charge voltage plateau at 3C owing to its enhanced ionic/electronic conductivity, which accelerates the redox reactions of LiPSs. Besides, the PC-CeO2 cathode delivers an impressive initial specific areal capacity of 7.5 mAh cm−2 and maintains a reversible capacity of 5.6 mAh cm−2 after 60 cycles at 0.2C at a sulfur loading of 5.5 mg cm−2 (Fig. 4d). The long-term cycling stability was tested at 1C (Fig. 4g), where the PC-CeO2 cathode achieved an initial discharge capacity of 826.2 mAh g−1, gradually declining to 375.1 mAh g−1 after 1000 cycles, with an exceptionally low capacity decay rate of just 0.054% per cycle and 0.035% per cycle at 2C, as depicted in Fig. S15. This demonstrates the remarkable cycling stability of the PC-CeO2 cathode. A pouch cell was constructed to assess the practical viability of the PC-CeO2 electrode. As displayed in Fig. 4f, the battery exhibits an impressive initial specific capacity of 845.2 mAh g−1 at 0.2C. Furthermore, as demonstrated in Fig. S16, a rigorous test of the pouch battery's mechanical properties was conducted by folding it at various angles. Despite these deformations, the battery effortlessly illuminates a “NENU” panel, showing its remarkable anti-folding characteristics and mechanical durability. This underscores the potential of Li–S batteries with PC-CeO2 cathode for flexible and reliable energy storage applications. In order to further verify the influence of different host materials in Li–S batteries, the cells with the PC-CeO2 and PC electrodes are disassembled after 500 cycles at 2C, and the SEM images are shown in Fig. S17. First, the SEM images of the sulfur cathodes paired with the PC-CeO2 and PC electrodes are shown in Fig. S17a and b, respectively. Compared with the cathode assembled with the PC-CeO2 electrode, the cathode with the PC electrode is severely split after 500 cycles. Besides, the SEM images of Li anodes show that the lithium electrode assembled with the PC-CeO2 electrode is well protected with a smooth surface and uniform deposition (Fig. S17c). On the contrary, the surface of the lithium electrode with the PC electrode is severely damaged after 500 cycles (Fig. S17d). As illustrated in Fig. S18, the EDS mapping on the cycled separator (a) and cathode (b) shows the uniform distribution of Ce, S, O and N elements. Apart from that, in the XRD patterns of PC-CeO2 cathode (Fig. S19), the diffraction peak at 27° is indexed to the (111) planes of Li2S (PDF#26-1188), which emphasizes the accumulation of Li2S. In conclusion, the cells with the PC-CeO2 host material can not only inhibit the shuttle of LiPSs but also adsorb LiPSs and catalyze the conversion. Meanwhile, as illustrated in Fig. 4g and Tables S3 and 4, the electrochemical performance of PC-CeO2 is more competitive than other sulfur host materials in Li–S batteries.45–50
To further explain the adsorption and catalytic capacity of LiPSs for the PC-CeO2, the density functional theoretical calculations were conducted, and the (111) plane and (110) planes of CeO2 were chosen as the calculation plane owing to the HRTEM image and the XRD results of PC-CeO2. The corresponding results are presented in Fig. 5a–c. The adsorption models of sulfur species on PC-CeO2(111), PC-CeO2(110) and PC are shown in Fig. 5a, S20 and S21, respectively. As shown in Fig. 5b, PC-CeO2(111) exhibits the highest adsorption energy towards different sulfur species compared with the PC, and the adsorption energies among Li2S, Li2S2, Li2S4, Li2S6, Li2S8, S8 and PC-CeO2 are calculated as −4.61, −3.73, −2.71, −2.48, −2.73 and −1.10 eV, respectively. The above results further illustrate the strong adsorption capacity between PC-CeO2(111) and LiPSs. Fig. 5c depicts the Gibbs free energy curves for the reduction of sulfur from S8 to Li2S on PC-CeO2(111), PC-CeO2(110) and PC. The process from S8 to Li2S8 for PC-CeO2(111), PC-CeO2(110) and PC is calculated as −4.47, −3.68 and −2.82 eV, respectively, showing that the step is a spontaneous reaction, and the results further show that the process from S8 to Li2S8 for PC-CeO2(111) is the easiest to proceed. Besides, the highest positive Gibbs free energy (ΔG) in the entire sulfur reduction process is the conversion from Li2S2 to Li2S, marking it the rate-limiting step in the LiPS reduction reaction. Notably, the ΔG for PC-CeO2(111) is only 0.05 eV, which is lower than 0.62 eV for PC-CeO2(110) and 1.11 eV for PC, indicating that the PC-CeO2(111) can accelerate the electrochemical redox reactions of sulfur. Finally, the mechanism of the PC-CeO2/S cathode to enhance reaction kinetics in Li–S batteries is illustrated in Fig. 5d. On the one hand, the 3D porous structure offers ample space for the loading of CeO2 NPs as active sites, further offering enough chemisorption sites for LiPSs. On the other hand, the CeO2 shows remarkable adsorption and catalysis ability to LiPSs, improving the cycle stability and electrochemical performance of the batteries significantly.
Footnote |
| † These authors contributed equally to this work. |
| This journal is © The Royal Society of Chemistry 2026 |