Next Article in Journal
Multiscale Evaluation of an Electrically Heated Thermal Battery for High-Temperature Industrial Energy Storage
Previous Article in Journal
Environmental Performance of the Sewage Sludge Gasification Process Considering the Recovered CO2
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

High-Efficiency Polysulfide Trapping with g-C3N4/CNT Hybrids for Superior Lithium-Sulfur Batteries

1
Institute of Carbon Neutrality, Zhejiang Wanli University, Ningbo 315100, China
2
Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Dalian 116023, China
*
Authors to whom correspondence should be addressed.
Energies 2025, 18(17), 4462; https://doi.org/10.3390/en18174462
Submission received: 23 July 2025 / Revised: 12 August 2025 / Accepted: 20 August 2025 / Published: 22 August 2025
(This article belongs to the Section D2: Electrochem: Batteries, Fuel Cells, Capacitors)

Abstract

Commercialization of lithium-sulfur (Li-S) batteries is critically hampered by the severe lithium polysulfide shuttle effect. Hence, designing multifunctional materials that synergistically provide physical confinement of polysulfides, chemical entrapment, and catalytic promotion is a viable route for improving Li-S battery performance. Herein, graphitic carbon nitride (g-C3N4) with abundant nitrogen atoms was used as the chemical adsorption material to realize a “physical-chemical” dual confinement for polysulfides. Furthermore, the integration of CNTs with g-C3N4 is intended to substantially enhance the conductivity of the cathode material. Consequently, the synthesized g-C3N4/CNT composite, which functions as an effective polysulfide immobilizer, significantly improved the cycling stability and discharge capacity of Li-S batteries. This enhancement can be attributed to its potent adsorption and catalytic activities. Li-S cells utilizing g-C3N4/CNT cathodes exhibit exceptional discharge capacity and notable rate capability. Specifically, after 100 cycles at 0.2 C, the discharge capacity was 701 mAh g−1. Furthermore, even at a high rate of 2 C, a substantial capacity of 457 mAh g−1 was retained.

1. Introduction

Since the utilization of lithium-ion batteries (LIBs) in portable electronic devices was established in the 1990s [1,2], tremendous efforts have been made to improve their energy density. However, its theoretical density is too low to satisfy the growing energy requirements [3]. Thus, research on high-energy-density battery systems has intensified in recent years [4,5].
Lithium-sulfur (Li-S) batteries are promising next-generation energy storage systems that have attracted significant interest due to their high theoretical specific capacity (1675 mAh g−1), natural abundance of sulfur, and eco-friendliness [6,7,8]. Nevertheless, the commercialization of Li-S batteries faces significant barriers, primarily short cycle life and low practical energy density, attributable to three key factors: (i) the poor electrical conductivity of sulfur/Li2S restricts charge/mass transport, limiting active material utilization; (ii) repeated ~80% volume fluctuations during cycling—caused by the density mismatch between S (2.07 g cm−3) and Li2S (1.66 g cm−3)—induce electrode structural degradation and compromise safety; (iii) soluble lithium polysulfides (LiPSs) migrate to the anode, forming irreversible Li2S2/Li2S deposits that reduce Coulombic efficiency [9,10]. While recent breakthroughs in cathode design and electrolyte engineering have shown promising pathways to mitigate these issues [11], challenges in scalable fabrication of high-performance systems persist.
To alleviate these issues, the further development of Li-S batteries has focused on investigating the physical confinement and chemical entrapment of LiPSs, as well as the catalytic acceleration of LiPS conversion, to fabricate long-lasting Li-S batteries. LiPS anchoring materials are regarded as an effective method to inhibit LiPS dissolution [12]. Multiple carbon-based composites—such as carbon nanotubes [13], graphene [14], and porous carbon architectures [15,16]—have been utilized to encapsulate sulfur in Li-S batteries, physically confining LiPSs through Van der Waals interactions. While these materials improve conductivity, their weak adsorption often fails to suppress long-term shuttle effects [17]. To enhance chemisorption, polar materials, including metal oxides (e.g., TiO2 and MnO2) [18,19], sulfides (e.g., CoS2 and WS2) [20,21], and single-atom catalysts [22], have shown promise. However, these materials typically exhibit limited conductivity or require complex synthesis routes.
However, single materials typically possess some fundamental drawbacks that prevent simultaneous further optimization towards higher conductivity and catalytic ability to satisfy practical applications in sulfur reactions. An alternative approach is to form hybrid materials with different components that combine the advantages of individual materials and compensate for their individual drawbacks to design better Li-S batteries. Wang’s group reported enhanced polysulfide immobilization and catalytic conversion through in situ growth of NiCo2S4 nanoparticles on reduced graphene oxide [23]. Lin and co-workers reported that the highly N-doped carbon/graphene hybrid could effectively suppress the cumbersome shuttle effect [24]. Despite recent progress in binary composites (e.g., conductive carbon/polar material hybrids), the simultaneous optimization of conductivity, adsorption capacity, and scalable fabrication remains challenging. This gap necessitates the development of rationally designed multifunctional hosts that integrate complementary properties using simple methods.
Graphitic C3N4 (g-C3N4) is a polar non-metal material like layered graphite. The dominant pyridinic nitrogen can serve as LiPS adsorption sites to immobilize and interact with LiPSs to accelerate the electrochemical conversion processes [25,26]. In addition, g-C3N4 exhibits low density, environmental compatibility, and ease of synthesis. These advantages establish it as a viable cathode material for long-cycle Li-S batteries, as demonstrated by Nazar’s group [27]. However, the intrinsically low electronic conductivity of g-C3N4 severely restricts charge transfer kinetics. For instance, Huangfu et al. demonstrated through four-point probe measurements that pristine g-C3N4 nanosheets exhibit conductivity values as low as 5.6 × 10−5 S cm−1 [28], while Song et al. reported similarly suppressed conductivities (~1.3 × 10−3 S cm−1) that impair sulfur utilization in Li-S batteries [29].
While g-C3N4 was initially proposed as a polysulfide anchor, its practical application remains hindered by its low conductivity and limited pore engineering. In this study, we design a conductive hybrid scaffold by integrating CNTs within a porous g-C3N4 matrix. This synergistic structure not only enhances charge transfer but also provides hierarchical pores for sulfur confinement and catalytic sites for accelerated conversion kinetics, addressing key limitations of single-component hosts. To address these limitations, we designed a hierarchical g-C3N4/CNT composite using a scalable two-step process. Unlike complex hybrid syntheses that require multi-step modifications [30], our approach directly couples conductive CNTs with nitrogen-rich g-C3N4. This leverages: (i) g-C3N4’s inherent polysulfide affinity via pyridinic N sites [27], (ii) CNT’s 3D conductive network for charge transfer, and (iii) synergistic catalysis at their interfaces—a feature not exploited in prior g-C3N4-based hosts [27]. The sustained capacity of 701 mAh g−1 (g-C3N4/CNT) after 100 cycles at 0.2 C outperforms pure g-C3N4 cathodes (440 mAh g−1) and competes favorably with contemporary hosts such as ZIF-8-derived MOFs (515 mAh g−1) [31], VO2-VS binary composites (589 mAh g−1) [32], and NiCo2S4/graphene hybrids (650 mAh g−1) [23]. Notably, our capacity retention at 2 C (457 mAh g−1) exceeds that of typical MOF cathodes (300–400 mAh g−1) and MoS2-based systems (∼400 mAh g−1) under equivalent conditions [6,33], while utilizing a low-cost, eco-friendly carbon/nitrogen architecture. In this composite, g-C3N4 provides abundant reaction sites for polysulfide capture and catalytic transformation, while the introduced CNTs form a conductive network that supports fast charge/mass transfer, which is confirmed by the CV results at different scanning rates. Consequently, the g-C3N4/CNT composite delivers synergistic enhancement in Li-S batteries, effectively suppressing polysulfide shuttling and boosting electrochemical performance. This synergy-driven design strategy enables practical Li-S batteries and provides a scalable paradigm for related energy storage systems, such as Na/K-S batteries.

2. Materials and Methods

The interconnected macroporous structure with thin, curled g-C3N4 nanosheets was synthesized via a two-step protocol: (1) 1.5 g melamine was dispersed in 90 mL deionized water under continuous stirring at 90 °C for 1 h; (2) the resultant mixture underwent hydrothermal treatment in a 150 mL Teflon-lined stainless steel autoclave (Xi’an Changyi Instrument Equipment Co., Ltd., Xi’an, China) at 185 °C for 24 h. The product was collected by filtration, washed, dried, and subsequently calcined at 520 °C for 4 h in an air atmosphere.
The g-C3N4/CNT composite was synthesized as follows: (1) 0.2 g g-C3N4 was dispersed in 150 mL deionized water via 60-min ultrasonic treatment to form a homogeneous suspension, ultrasonication was performed using a KQ-300DE instrument (Kunshan Ultrasonic Instruments Co., Ltd., Kunshan, China) at 40 kHz and 300 W for 60 min; (2) 0.14 g multilayered CNTs (XFNANO Materials Tech Co., Ltd., Nanjing, China, ~1 µm length, ~20 nm diameter) were introduced into the solution; (3) the mixture was magnetically stirred for 4 h followed by drying at 60 °C to yield the composite powder.
Preparation of sulfur cathodes. The S@g-C3N4/CNT composite was prepared using a traditional melt-diffusion approach by mixing g-C3N4/CNT and pure sulfur powder at a mass ratio of 1:3. Afterwards, the sample was ground and calcined at 155 °C in Ar for 12 h to obtain S@g-C3N4/CNT.
Physical characterization was performed as follows: Phase analysis was performed using X-ray diffraction (XRD, Bruker D8 Focus, Bruker Corporation, Karlsruhe, Germany) with Cu Kα radiation (λ = 1.5406 Å). Morphological and structural characterizations were performed using scanning electron microscopy (SEM, ZEISS GeminiSEM 500, Carl Zeiss AG, Oberkochen, Germany) and transmission electron microscopy (TEM, JEOL JEM-2100F, JEOL Ltd., Akishima, Tokyo). Surface properties, including specific surface area and pore size distribution, were determined through nitrogen adsorption-desorption measurements at 77 K (BET, V-Sorb 2800P system, Gold APP Instruments Corporation China, Beijing, China). Chemical state analysis was conducted via X-ray photoelectron spectroscopy (XPS, PHI 5000 Versa Probe, ULVAC-PHI Incorporated, Kanagawa, Japan), and thermal stability was evaluated by thermogravimetric analysis (TGA, TA Q600 thermobalance, TA Instruments, New Castle, DE, USA).
Electrochemical characterization was conducted using CR2032 coin cells with S@g-C3N4/CNT cathodes. The cathode slurry was formulated by homogenizing active material, Super P conductive carbon, and PVDF binder at an 8:1:1 mass ratio in N-methyl-2-pyrrolidone solvent, followed by doctor-blade coating onto carbon-coated aluminum foil. After drying at 60 °C, electrodes were punched into 9.0 mm disks. Cells employed 1.0 M LiTFSI and 0.1 M LiNO3 in DOL/DME (1:1 v/v) electrolyte. Electrochemical evaluation comprised cyclic voltammetry (CHI-660C workstation, CH Instruments, Inc., Shanghai, China, 0.1 mV s−1 scan rate), electrochemical impedance spectroscopy (Versa STAT4 system, AMETEK Inc., Berwyn, PA, USA, 100 kHz–0.01 Hz), and galvanostatic cycling (Neware BTS4000 tester, Neware Technology Ltd., Shenzhen, China, 1.7–2.8 V voltage window).
Symmetric cells were constructed with identical g-C3N4/CNT (or pristine g-C3N4) electrodes as both working and counter electrodes. The Li2S6 electrolyte (0.1 M) was prepared by stoichiometrically reacting Li2S and sulfur (1:5 mass ratio) in a DOL/DME (1:1 v/v) solvent. Subsequently, 40 µL of the polysulfide solution was injected into the cell. Specifically, the E/S ratio used in coin cell assembly was ~10 µL mg−1. The areal mass loading of sulfur in the cathodes was ~2 mg cm−2, unless otherwise stated. For all electrochemical measurements, pure lithium metal foil was used as both counter and reference electrodes, ensuring an excess lithium supply; therefore, the N/P ratio was not applicable in this configuration.
Linear Sweep Voltammetry (LSV): A 9:1 (w/w) mixture of g-C3N4/CNT (or pristine g-C3N4) and PVDF binder in NMP was homogenized and drop-cast onto a glassy carbon working electrode (0.07 cm2). Measurements employed 0.1 M Li2S in methanol electrolyte. Reference electrodes were test-specific: Ag/AgCl for LSV in methanol, none for symmetric cells, and Li metal for full-cell CVs; all potentials were referenced to their native scales.
Li2S Nucleation Testing: Cells configured with g-C3N4/CNT (or pristine g-C3N4) cathodes and lithium foil anodes used 0.25 M Li2S8 in a tetraethylene glycol dimethyl ether (TEGDME) electrolyte. After galvanostatic discharge to 2.06 V, potentiostatic deposition was performed at 2.05 V.
Density functional theory (DFT) calculations employed the projector-augmented wave method and a plane-wave basis set within the Vienna Ab Initio Simulation Package (VASP) to investigate Li2S4 adsorption on g-C3N4, g-C3N4/CNT, and CNT substrates [34]. The Perdew-Burke-Ernzerhof (PBE) generalized gradient approximation was used to describe the exchange-correlation effects, with Grimme’s D3 dispersion correction applied for adsorption energy (Eads) calculations [35]. Based on the XRD and TEM results, the C6N7 ring served as the structural unit for the g-C3N4 lattice, defined by the following parameters: a = 12.52 Å, b = 12.71 Å, c = 24.51 Å, α = 90°, β = 90°, and γ = 120°. Multi-walled carbon nanotubes were used to construct slab models for both g-C3N4/CNT and CNT, adopting the same lattice parameters as those of g-C3N4. Identical Li2S4 molecule configurations were bonded to each of the surface models. All models incorporated a 15 Å vacuum layer that was perpendicular to the surface. Computational settings included a plane-wave cutoff energy of 400 eV and a 2 × 2 × 1 Monkhorst-Pack k-point mesh. Self-consistent field iterations converged at 1 × 10−4 eV, and ionic relaxation ceased when maximal atomic forces fell below 0.02 eV Å−1. Dipole corrections were applied to all slabs. Adsorption energy was calculated using the following formula:
E ads = E LS + E sur E LS - sur
where ELS and Eₛᵤᵣ represent the total energies of the isolated Li2S4 molecule and the clean substrate surface (g-C3N4, g-C3N4/CNT, or CNT), respectively, and E LS - sur is the total energy of the Li2S4-adsorbed system. In this study, a positive Eads value indicates a stable adsorption configuration relative to the separated adsorbate and substrate.

3. Results and Discussions

XRD spectra in Figure 1 characterize CNT, g-C3N4, g-C3N4/CNT composite, and S@g-C3N4/CNT. The commercial CNT exhibits characteristic (002) and (100) reflections at 26.2° and ~43°, confirming graphitic crystallinity [36]. The two peaks located at 2θ of about 27.8° and 13.1° in g-C3N4 correspond to (002) and (100) planes [37]. The XRD profile of g-C3N4/CNT composites predominantly reflects g-C3N4 characteristics, with CNT diffraction signatures masked by their low loading concentration. The g-C3N4 (002) peak shifts from 27.8° to 28.1° in the composite, corresponding to a d-spacing reduction from 0.321 nm to 0.317 nm (Table S3). This compression reflects strong interfacial coupling with CNTs, which facilitates charge transfer between components. Concurrently, crystallite size decreases from 18.2 nm to 15.7 nm, indicating CNT-induced fragmentation increases the number of accessible active sites. Characteristic peaks of orthorhombic sulfur (α-S8, JCPDS 08-0247) at 2θ = 23.1° (222), 25.9° (026), and 26.8° (206) are clearly identified in the S@g-C3N4/CNT pattern, confirming successful sulfur infusion into the composite host [38]. The absence of sulfur peaks in g-C3N4/CNT further validates that sulfur loading occurred post-scaffold synthesis.
Figure 2a shows an interconnected macroporous architecture with a thin, curled nanosheet morphology in pristine g-C3N4. Upon CNT integration (Figure 2b,c), the composite evolves into a sponge-like matrix in which the g-C3N4 and CNT phases interpenetrate homogeneously. This hierarchical porosity accommodates sulfur infiltration while mitigating volume expansion during cycling. Concurrently, the percolating CNT network enhances bulk conductivity, facilitating rapid charge/mass transport. Uniform elemental distribution is further evidenced by C/N mappings (Figure 2e,f) correlating with the composite’s SEM morphology (Figure 2d). Homogeneous sulfur distribution within the S@g-C3N4/CNT composite is further confirmed by EDX mapping (Supplementary Materials Figure S1), consistent with its effective infiltration into the porous host.
TEM was used to further explore the properties of the g-C3N4/CNT samples. Figure 3a,b displays the TEM results of g-C3N4/CNT, which demonstrate that both g-C3N4 and CNTs preserve their initial morphology, and CNTs are uniformly distributed throughout the g-C3N4 host structure. The interconnected CNT network facilitates rapid charge/mass transfer, while the g-C3N4 nanosheets provide abundant reaction sites for polysulfide capture and transformation. The crystalline features of g-C3N4/CNT are investigated by HRTEM (Figure 3c,d), which show the typical (002) graphitic plane for CNTs with stripe intervals of 0.340 nm [39,40].
Figure 4a presents the N2 adsorption-desorption isotherm of the g-C3N4/CNT composite, which exhibits a characteristic Type IV profile [41]. The material achieves a high specific surface area of 66.485 m2 g−1, confirming its porous architecture. Corresponding pore size distribution (Figure 4b) reveals a dominant mesoporous peak centered at 2.5 nm. These engineered mesopores facilitate polysulfide adsorption and mitigate shuttle effects through physical confinement. The BET surface area (66.485 m2/g) and pore volume (0.203 cm3/g) of g-C3N4/CNT exceed those of most reported g-C3N4 hosts (typically < 30 m2/g) [24], primarily due to the CNT-induced 3D hierarchical porosity. The dominant mesopores (centered at 2.5 nm) synergize with the visible macroporous framework (Figure 2c) to confine polysulfides while enabling efficient, unhindered ion transport. This architecture contributes to high sulfur utilization (74 wt%) and stable cycling.
XPS was used to study the surface properties of the materials. Only C, N, and O were observed in the survey spectrum shown in Figure 5a, which demonstrates the purity of the synthesized products. The C1s spectrum of g-C3N4/CNT (Figure 5b) reveals two dominant peaks at 284.8 eV (C-C) and 289.3 eV (C-N) [42], with an additional C-OH component at 285.8 eV, indicating oxygen-containing functionalities on CNTs [43]. Deconvolution of the N1s region (Figure 5c) identifies three nitrogen configurations: pyridinic N (398.7 eV), pyrrolic N (400.4 eV), and graphitic N (401.2 eV) originating from g-C3N4 [44]. Thermogravimetric analysis (Figure 5d) quantifies sulfur loading at 74 wt% in both S@g-C3N4 and S@g-C3N4/CNT composites. The presence of C-OH species (285.8 eV) on CNTs further augments polysulfide immobilization. Oxygen functional groups engage LiPS via hydrogen bonding and Li+ coordination, complementing g-C3N4’s nitrogen-mediated chemisorption. This cooperative effect between oxygen (physical-chemical adsorption) and nitrogen (chemical adsorption) sites establishes a multi-modal trapping network, reducing shuttle effects and enhancing cycle stability relative to non-functionalized carbons [45]. To further validate the proposed physical-chemical dual-confinement mechanism, we performed DFT calculations (Figure S2 in the Supplementary Materials). The DFT results quantitatively compare the adsorption energies of Li2S4 on g-C3N4, g-C3N4/CNT, and CNT surfaces, demonstrating that the g-C3N4/CNT composite exhibits strong adsorption (4.23 eV), which is beneficial for the polysulfide reduction reaction.
The synthesized composite was designed for robust polysulfide adsorption and accelerated conversion of sulfur species. Linear sweep voltammetry (LSV, Figure 6a) reveals enhanced current response kinetics in g-C3N4/CNT compared to pristine g-C3N4, validating the superior catalytic conversion of Li2S to polysulfides. This enhancement can be attributed to the improved electrical conductivity and active material utilization. Complementary cyclic voltammetry (CV) of symmetric cells with Li2S6 electrolyte (Figure 6b) demonstrates sharper redox peaks at ±0.36 V for the composite, indicating accelerated polysulfide interconversion. The reduced ΔE (0.36 V vs. 0.51 V for pure g-C3N4) quantitatively confirms the superior electrocatalytic activity of g-C3N4/CNT, where minimized polarization accelerates LiPS redox cycling. Excellent reaction reversibility is confirmed by overlapping CV profiles during cycling (Figure 6c). The Li2S deposition capacity (Figure 6d,e) is 233.3 mAh g−1 for g-C3N4/CNT, indicating efficient polysulfide-to-Li2S conversion catalysis. The multifunctional mechanism (Figure 6f) integrates strong chemisorption with catalytic cycling kinetics, enabling rapid sulfur redox reactions.
Cyclic voltammetry of S@g-C3N4/CNT cathodes (Figure 7a) exhibits two distinct redox couples. The cathodic peak at 2.30 V (C) signifies reduction of elemental sulfur to long-chain polysulfides (Li2Sₓ, 4 ≤ x ≤ 8), while the 2.00 V peak (B) corresponds to further reduction to solid Li2S2/Li2S. Anodic activity at 2.40 V (A) reflects sulfur regeneration. Notably, the dual oxidation peaks at ~2.30/2.40 V vs. Li+/Li maintain their intensity over cycles, demonstrating rapid Li2S/Li2S2 decomposition kinetics and exceptional electrochemical reversibility. In contrast, S@g-C3N4 cathodes (Figure 7b) display a single broad oxidation peak and kinetic hysteresis at elevated scan rates, indicating sluggish sulfur electrochemistry. Minimal increase in peak separation (0.10 → 0.40 mV s−1) further confirms the superior reaction reversibility of S@g-C3N4/CNT. The catalytic superiority of g-C3N4 is further evidenced by scan-rate-dependent peak shifts (Table S2), where smaller polarization changes indicate stabilized transition states during sulfur reduction. The attenuated peak (C) shift (34 mV vs. 106 mV) in S@g-C3N4/CNT confirms g-C3N4’s catalytic function in reducing reaction hysteresis, synergistically enhanced by CNT-enabled rapid charge transfer. Comparative linear fits of peak currents (Figure 7c,d) enable quantitative analysis of Li+ diffusion kinetics via the Randles–Ševčík equation [46]:
I = 269000 × n1.5 × A × D0.5 × C × v0.5
Within the Randles–Ševčík framework, the peak current (I) correlates with the electron transfer number (n), electrode geometric area (A), Li+ diffusion coefficient (D), bulk Li+ concentration (C), and scan rate (v).
As shown in the results, the S@g-C3N4/CNT cathode demonstrates a higher slope rate than the S@g-C3N4 cathode, confirming the accelerated Li+ diffusion rate. The minor y-intercepts in peak current plots (Figure 7c,d) originate from pseudocapacitive charge storage at conductive interfaces, with reduced values for S@g-C3N4/CNT confirming enhanced electrode kinetics—consistent with its lower ΔE in symmetric-cell CVs (Figure 6b).
Cycling stability at 0.2C (Figure 8a) reveals superior performance of S@g-C3N4/CNT over S@g-C3N4 cathodes. The composite cathode delivers an initial capacity of 1227 mAh g−1 (vs. 1127 mAh g−1 for pristine), maintaining 701 mAh g−1 after 100 cycles with minimal decay. In contrast, the control cathode retains only 440 mAh g−1. Galvanostatic charge/discharge profiles of S@g-C3N4/CNT (Figure 8b) align with CV data: the 2.30 V plateau corresponds to S8 → Li2Sₙ (4 ≤ n ≤ 8) reduction, while the extended 2.00 V plateau signifies Li2Sₙ → Li2S2/Li2S conversion. Oxidation plateaus at ~2.36/2.45 V reflect Li2S2/Li2S → S8/Li2S8 reactions. This stable voltage plateau configuration enhances cycling efficiency by mitigating polarization during sulfur redox transitions.
Rate capability analysis (Figure 8c) reveals superior performance of the S@g-C3N4/CNT cathode, delivering specific capacities of 1264 (0.2C), 961 (0.5C), 659 (1C), and 457 mAh g−1 (2C)—surpassing S@g-C3N4 at all rates. Upon reverting to 0.2C, the composite cathode recovers 1055 mAh g−1, demonstrating exceptional rate resilience. The superior performance of the S@g-C3N4/CNT composite compared to representative cathodes reported in the recent literature is comprehensively summarized in Table S1 (Supplementary Materials). As evident from Table S1, our optimized S@g-C3N4/CNT composite achieves a remarkable Cycle performance, significantly exceeding values reported for state-of-the-art cathode materials. This enhancement is primarily attributed to the synergistic effect between g-C3N4 and CNTs, which facilitates efficient charge separation and transfer which also proved by the reduced ΔE by g-C3N4/CNT in Figure 6b. Electrochemical impedance spectroscopy (Figure 8d) corroborates enhanced reaction kinetics, where reduced charge-transfer resistance in S@g-C3N4/CNT facilitates rapid charge transport. Long-term cycling (Figure 8e) shows that the composite maintains 383 mAh g−1 after 500 cycles from an initial 639 mAh g−1. This stability originates from synergistic effects: (1) N-rich sites in g-C3N4 and hierarchical porosity enable effective polysulfide confinement, and (2) CNT networks substantially elevate electrical conductivity.

4. Conclusions

This work develops a facile two-step synthesis of hierarchically porous g-C3N4/CNT composites with an enhanced specific surface area (66.5 m2 g−1). As a multifunctional polysulfide reservoir, the designed architecture implements physicochemical dual confinement through (i) physical entrapment via mesopores (2.5 nm dominant) and (ii) chemical anchoring at N sites. The resulting S@g-C3N4/CNT cathodes deliver exceptional rate capability (457 mAh g−1 at 2C) and cycling stability (383 mAh g−1 after 500 cycles @1C). This strategy establishes a cost-effective pathway toward high-performance Li-S batteries via synergistic pore engineering and conductive-network optimization.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/en18174462/s1, Figure S1. (a) SEM image of S@g-C3N4/CNT and (b–d) corresponding elemental mapping. Figure S2. Adsorption configurations and adsorption energy of Li2S4 molecules on g-C3N4, g-C3N4/CNT and CNT surfaces. Table S1. Electrochemical performance comparison of Li-S batteries with different hosts of cathodes. Table S2. The scan-rate-dependent peak shifts of S@g-C3N4/CNT and S@g-C3N4 cathodes. Table S3. Structural parameters from XRD refinement. Table S4. The fitted charge-transfer resistance (Rct) and interfacial capacitance values for both S@g-C3N4 and S@g-C3N4/CNT cathodes. References [47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64] are cited in Supplementary Materials.

Author Contributions

Investigation, Z.C. (Zhen Chen), H.M., and J.W.; Data curation, Z.C. (Zhen Chen) and J.W.; Writing—original draft, Z.C. (Zhen Chen) and J.W.; Writing—review and editing, Z.C. (Zhen Chen), H.M., J.W., and L.Y.; Supervision, J.W., X.W., and Z.C. (Zhongwei Chen); Project administration, J.W., X.W., and Z.C. (Zhongwei Chen); Funding acquisition, J.W., X.W., and Z.C. (Zhongwei Chen). All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (No. 52402319, 22379047), Zhejiang Provincial Natural Science Foundation of China (under Grant No. LQN25B030003), Ningbo Yongjiang Talent Programme (under Grant No. 2023A395G), Entrepreneurial and innovative team project of Ningbo Yinzhou District (X.W.), and Ningbo Key Laboratory of High Energy Density Batteries, General scientific research project of Zhejiang Education Department (Y202353940), Talent research start-up project of Zhejiang Wanli University (SC1032380180530).

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Acknowledgments

The authors acknowledge the Beijing Super Cloud Computing Center (BSCC) for providing the HPC resources that contributed to the research results reported in this paper. URL: http://www.blsc.cn/, accessed on 1 September 2023.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Du, H.; Wang, Y.; Kang, Y.; Zhao, Y.; Tian, Y.; Wang, X.; Tan, Y.; Liang, Z.; Wozny, J.; Li, T.; et al. Side Reactions/Changes in Lithium-Ion Batteries: Mechanisms and Strategies for Creating Safer and Better Batteries. Adv. Mater. 2024, 36, 2401482. [Google Scholar] [CrossRef]
  2. Zheng, M.; You, Y.; Lu, J. Understanding materials failure mechanisms for the optimization of lithium-ion battery recycling. Nat. Rev. Mater. 2025, 10, 355. [Google Scholar] [CrossRef]
  3. Sharma, H.; Sharma, S.; Mishra, P.K. A critical review of recent progress on lithium-ion batteries: Challenges, applications, and future prospects. Microchem. J. 2025, 212, 113494. [Google Scholar] [CrossRef]
  4. Yamada, A. Hidden Negative Issues and Possible Solutions for Advancing the Development of High-Energy-Density in Lithium Batteries: A Review. Adv. Sci. 2024, 11, 2401739. [Google Scholar] [CrossRef]
  5. Fei, Y.; Li, G. Unveiling the Pivotal Parameters for Advancing High Energy Density in Lithium-Sulfur Batteries: A Comprehensive Review. Adv. Funct. Mater. 2024, 34, 2312550. [Google Scholar] [CrossRef]
  6. Li, S.; Chen, Z.; Chen, J.; Luo, X.; Qiu, X.; Qian, Y. Engineering of Lignocellulose Pulp Binder for Ah-Scale Lithium–Sulfur Batteries. Adv. Energy Mater. 2025, 15, 2405461. [Google Scholar] [CrossRef]
  7. Min, Y.; Zou, X.; Lu, Q.; Cai, W.; Bu, Y. Advances in the Catalytic Mechanism of Metal Oxides for Lithium–Sulfur Batteries. Small 2025, 21, 2411794. [Google Scholar] [CrossRef]
  8. Chen, J.; Fu, Y.; Guo, J. Development of Electrolytes under Lean Condition in Lithium–Sulfur Batteries. Adv. Mater. 2024, 36, 2401263. [Google Scholar] [CrossRef] [PubMed]
  9. Pathak, A.D.; Cha, E.; Choi, W. Towards the commercialization of Li-S battery: From lab to industry. Energy Storage Mater. 2024, 72, 103711. [Google Scholar] [CrossRef]
  10. Li, B.Q.; Zhang, S.Y.; Kong, L.; Peng, H.J.; Zhang, Q. Porphyrin organic framework hollow spheres and their applications in lithium-sulfur batteries. Adv. Mater. 2018, 30, 1707483. [Google Scholar] [CrossRef]
  11. Shi, C.; Takeuchi, S.; Alexander, G.V.; Hamann, T.; O’Neill, J.; Dura, J.A.; Wachsman, E.D. High Sulfur Loading and Capacity Retention in Bilayer Garnet Sulfurized-Polyacrylonitrile/Lithium-Metal Batteries with Gel Polymer Electrolytes. Adv. Energy Mater. 2023, 13, 2301656. [Google Scholar] [CrossRef]
  12. Feng, J.; Zhang, C.; Liu, W.; Yu, S.; Wang, L.; Wang, T.; Shi, C.; Zhao, X.; Chen, S.; Chou, S.; et al. Enabling Efficient Anchoring-Conversion Interface by Fabricating Double-Layer Functionalized Separator for Suppressing Shuttle Effect. Angew. Chem. Int. Ed. 2024, 136, e202407042. [Google Scholar] [CrossRef]
  13. Fu, Y.Z.; Su, Y.S.; Manthiram, A. Highly reversible lithium/dissolved polysulfide batteries with carbon nanotube electrodes. Angew. Chem. Int. Ed. 2013, 52, 6930–6935. [Google Scholar] [CrossRef] [PubMed]
  14. Kaiser, M.R.; Ma, Z.; Wang, X.; Han, F.; Gao, T.; Fan, X.; Wang, J.Z.; Liu, H.K.; Dou, S.; Wang, C. Reverse microemulsion synthesis of sulfur/graphene composite for lithium/sulfur batteries. ACS Nano 2017, 11, 9048–9056. [Google Scholar] [CrossRef] [PubMed]
  15. Wang, X.; Chen, H.; Yu, F.; Wei, S.; Liu, J.; Que, L.; Wu, Y.; Pang, Q.; Lu, C.; Xie, Y. Two-in-one linkage reaction of aloe vera peel-derived porous carbon in lithium–sulfur batteries for synergistic enhancement of performance and stability. Chem. Eng. J. 2025, 513, 162707. [Google Scholar] [CrossRef]
  16. Nong, S.; Huang, D.; Li, Y.; Yang, R.; Xie, J.; Li, J.; Huang, H.; Liang, X.; Li, G.; Lan, Z.; et al. Multifunction-balanced porous carbon and its application in sulfur-loading host and separator modification for lithium–sulfur batteries. J. Energy Storage 2024, 81, 110296. [Google Scholar] [CrossRef]
  17. Zhang, L.; Liu, D.; Muhammad, Z.; Wan, F.; Xie, W.; Wang, Y.; Song, L.; Niu, Z.; Chen, J. Single Nickel Atoms on Nitrogen-Doped Graphene Enabling Enhanced Kinetics of Lithium–Sulfur Batteries. Adv. Mater. 2019, 31, 1903955. [Google Scholar] [CrossRef]
  18. Ma, L.; Zhang, Y.; Zhang, S.; Wang, L.; Zhang, C.; Chen, Y.; Wu, Q.; Chen, L.; Zhou, L.; Wei, W. Integrating Energy Band Alignment and Oxygen Vacancies Engineering of TiO2 Anatase/Rutile Homojunction for Kinetics-Enhanced Li–S Batteries. Adv. Funct. Mater. 2023, 33, 2305788. [Google Scholar] [CrossRef]
  19. Jiang, Q.; Xu, H.; Hui, K.S.; Ye, Z.; Zha, C.; Lin, Z.; Zheng, M.; Lu, J.; Hui, K.N. Breaking the Passivation Effect for MnO2 Catalysts in Li−S Batteries by Anion-Cation Doping. Angew. Chem. Int. Ed. 2024, 63, e202408474. [Google Scholar] [CrossRef]
  20. Liang, J.; Sun, J.; Cao, X.; Li, X.; Chen, X.; Xing, R.; Kong, J. Enhanced Reaction Kinetics in Sodium-Ion Batteries Achieved by 3D Heterostructure CoS2/CoS with Self-Induced Internal Electric Field. Adv. Sci. 2025, 12, 2502241. [Google Scholar] [CrossRef]
  21. Ye, Z.; He, L.; Liao, J.; Huang, Y.; Jiang, Q. Preparation of multilayer core-shell WS2@CZIF-8@CZIF-67 and application in lithium-sulfur batteries. J. Alloys Compd. 2025, 1022, 179754. [Google Scholar] [CrossRef]
  22. Li, S.; Li, L.; Zhao, Y.; Yang, H.; Tong, H.; Fan, S.; Wang, Z.; Xu, W. Modulating the d-band center of single-atom catalysts for efficient Li2S2-Li2S conversion in durable lithium-sulfur batteries. Energy Storage Mater. 2024, 70, 103477. [Google Scholar] [CrossRef]
  23. Huang, C.; Sun, T.; Shu, H.; Chen, M.; Lian, Q.; Zhou, Y.; Gao, P.; Xu, S.; Yang, X.; Wu, M.; et al. Multifunctional reaction interfaces for capture and boost conversion of polysulfide in lithium-sulfur batteries. Electrochim. Acta 2020, 334, 135658. [Google Scholar] [CrossRef]
  24. Xu, H.; Jiang, Q.; Zhang, B.; Chen, C.; Lin, Z. Integrating conductivity, immobility, and catalytic ability into high-n carbon/graphene sheets as an effective sulfur host. Adv. Mater. 2020, 32, 1906357. [Google Scholar] [CrossRef] [PubMed]
  25. Chen, Y.; Wu, Y.; Li, L.; Liao, Y.; Luo, S.; Wu, Y.; Qing, Y. Polysulfides manipulation: Constructing g-C3N4 networks encapsulated into natural wood fibers for high-performance lithium–sulfur batteries. Chem. Eng. J. 2023, 461, 141988. [Google Scholar] [CrossRef]
  26. Li, Y.; Deng, Y.; Yang, J.-L.; Tang, W.; Ge, B.; Liu, R. Bidirectional Catalyst with Robust Lithiophilicity and Sulfiphilicity for Advanced Lithium–Sulfur Battery. Adv. Funct. Mater. 2023, 33, 2302267. [Google Scholar] [CrossRef]
  27. Khasim, S.; Pasha, A.; Lakshmi, M.; Panneerselvam, C.; Ullah, M.F.; Darwish, A.A.A.; Hamdalla, T.A.; Alfadhli, S.; Al-Ghamdi, S.A. Synthesis of g-C3N4/CuO Nanocomposite as a Supercapacitor with Improved Electrochemical Performance for Energy Storage applications. Int. J. Electrochem. Sci. 2022, 17, 220838. [Google Scholar] [CrossRef]
  28. Huang, F.Y.; Zheng, T.; Zhang, K.; She, X.; Xu, H.; Fang, Z.; Xie, K. Facile fabrication of permselective g-C3N4 separator for improved lithium-sulfur batteries. Electrochim. Acta 2018, 272, 60–67. [Google Scholar] [CrossRef]
  29. Liu, G.; Yuan, H.; Wang, Z.; Qin, N.; Huang, Y.; Cao, Y.; Li, Y.; Lu, W.; Zeng, C.; Liu, Q.; et al. Cu single atoms regulating nitrogen active-sites of g-C3N4 for sodium ion storage. Energy Storage Mater. 2024, 71, 103608. [Google Scholar] [CrossRef]
  30. Zhang, Y.; Guo, C.; Zhou, J.; Yao, X.; Li, J.; Zhuang, H.; Chen, Y.; Chen, Y.; Li, S.L.; Lan, Y.Q. Anisotropically Hybridized Porous Crystalline Li-S Battery Separators. Small 2023, 19, 2206616. [Google Scholar] [CrossRef]
  31. Jin, L.; Li, B.; Zhang, K.; Qian, X.; Zhao, S.; Xu, H. Sn Nanoparticles Encapsulated in Nanoporous Nitrogen-Rich Carbon Derived from Sn-MOF@ZIF-8 for Lithium–Sulfur Battery Separators. ACS Appl. Nano Mater. 2024, 7, 19538–19547. [Google Scholar] [CrossRef]
  32. Deng, Y.; Tang, W.; Zhu, Y.; Ma, J.; Zhou, M.; Shi, Y.; Yan, P.; Liu, R. Catalytic VS2–VO2 Heterostructure that Enables a Self-Supporting Li2S Cathode for Superior Lithium–Sulfur Batteries. Small Methods 2023, 7, 2300186. [Google Scholar] [CrossRef]
  33. Yang, W.; Yang, W.; Dong, L.; Gao, X.; Wang, G.; Shao, G. Enabling immobilization and conversion of polysulfides through a nitrogen-doped carbon nanotubes/ultrathin MoS2 nanosheet core-shell architecture for lithium-sulfur batteries. J. Mater. Chem. A 2019, 7, 13103–13112. [Google Scholar] [CrossRef]
  34. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  35. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [PubMed]
  36. Reyhani, R.; Zadhoush, A.; Tabrizi, N.S.; Nazockdast, H.; Naeimirad, M. Synthesis and characterization of powdered CNT-doped carbon aerogels. J. Non-Cryst. Solids 2021, 571, 121058. [Google Scholar] [CrossRef]
  37. Fang, S.; Xia, Y.; Lv, K.; Li, Q.; Sun, J.; Li, M. Effect of carbon-dots modification on the structure and photocatalytic activity of g-C3N4. Appl. Catal. 2016, 185, 225–232. [Google Scholar] [CrossRef]
  38. Wei, Z.Z.; Zhang, N.X.; Feng, T.; Wu, F.; Zhao, T.; Chen, R.J. A copolymer microspheres-coated separator to enhance thermal stability of lithium-sulfur batteries. Chem. Eng. J. 2022, 430, 10. [Google Scholar] [CrossRef]
  39. Saini, R.; Farooq, U.; Imam, F. Investigation of the Synergistic Effect of Doping and Composite Formation on Electrochemical OER Performance of g-C3N4 for Sustainable Energy. Energy Fuels 2025, 39, 1327. [Google Scholar] [CrossRef]
  40. Kaya, S. Synthesis, characterization and 1-propanol electrooxidation application of carbon nanotube supported bimetallic catalysts. Int. J. Hydrogen Energy 2023, 48, 18398–18404. [Google Scholar] [CrossRef]
  41. Yan, Z.; Sun, M.; Wei, C.; Yuan, Z.; Gu, X.; Wang, J.; Wu, L.; Hu, Y.; Yang, L.; Wang, X.; et al. Design of self-supporting porous soft-hard carbon hybrid materials for ultra-long cycling lithium-ion storage. J. Power Sources 2025, 649, 237249. [Google Scholar] [CrossRef]
  42. Li, Y.; Xu, H.; Ouyang, S.; Lu, D.; Wang, X.; Wang, D.; Ye, J. In situ surface alkalinized g-C3N4 toward enhancement of photocatalytic H2 evolution under visible-light irradiation. J. Mater. Chem. A 2016, 4, 2943–2950. [Google Scholar] [CrossRef]
  43. Jiang, D.; Ma, W.; Xiao, P.; Shao, L.; Li, D.; Chen, M. Enhanced photocatalytic activity of graphitic carbon nitride/carbon nanotube/Bi2WO6 ternary Z-scheme heterojunction with carbon nanotube as efficient electron mediator. J. Colloid Interface Sci. 2018, 512, 693–700. [Google Scholar] [CrossRef] [PubMed]
  44. Yu, H.; Shi, R.; Zhao, Y.; Bian, T.; Zhao, Y.; Zhou, C.; Waterhouse, G.I.N.; Wu, L.-Z.; Tung, C.-H.; Zhang, T. Alkali-assisted synthesis of nitrogen deficient graphitic carbon nitride with tunable band structures for efficient visible-light-driven hydrogen evolution. Adv. Mater. 2017, 29, 1605148–1605155. [Google Scholar] [CrossRef]
  45. Cai, D.Q.; Yang, J.L.; Liu, T.; Zhao, S.X.; Cao, G. Interfaces-dominated Li2S nucleation behavior enabled by heterostructure catalyst for fast kinetics Li-S batteries. Nano Energy 2021, 89, 106452. [Google Scholar] [CrossRef]
  46. Sun, R.; Qu, M.; Peng, L.; Yang, W.; Wang, Z.; Bai, Y.; Sun, K. Regulating electrochemical kinetics of CoP by incorporating oxygen on surface for high-performance Li–S batteries. Small 2023, 19, 2302092. [Google Scholar] [CrossRef] [PubMed]
  47. Talapaneni, S.N.; Hwang, T.H.; Je, S.H.; Buyukcakir, O.; Choi, J.W.; Coskun, A. Elemental-sulfur-mediated facile synthesis of a covalent triazine framework for high-performance lithium–sulfur batteries. Angew. Chem. Int. Ed. 2016, 55, 3106–3111. [Google Scholar] [CrossRef]
  48. Xu, F.; Yang, S.; Jiang, G.; Ye, Q.; Wei, B.; Wang, H. Fluorinated, sulfur-rich, covalent triazine frameworks for enhanced confinement of polysulfides in lithium–sulfur batteries. ACS Appl. Mater. Interfaces 2017, 9, 37731–37738. [Google Scholar] [CrossRef]
  49. Je, S.H.; Kim, H.J.; Kim, J.; Choi, J.W.; Coskun, A. Perfluoroaryl-elemental sulfur SNAr chemistry in covalent triazine frameworks with high sulfur contents for lithium–sulfur batteries. Adv. Funct. Mater. 2017, 27, 1703947. [Google Scholar] [CrossRef]
  50. Meng, Y.; Lin, G.; Ding, H.; Liao, H.; Wang, C. Impregnation of sulfur into a 2D pyrene-based covalent organic framework for high-rate lithium–sulfur batteries. J. Mater. Chem. A 2018, 6, 17186–17191. [Google Scholar] [CrossRef]
  51. Wang, D.G.; Tan, L.; Wang, H.; Song, M.; Wang, J.; Kuang, G.C. Multiple Covalent Triazine Frameworks with Strong Polysulfide Chemisorption for Enhanced Lithium-Sulfur Batteries. ChemElectroChem 2019, 6, 2777–2781. [Google Scholar] [CrossRef]
  52. Xu, F.; Yang, S.; Chen, X.; Liu, Q.; Li, H.; Wang, H.; Wei, B.; Jiang, D. Energy-storage covalent organic frameworks: Improving performance via engineering polysulfide chains on walls. Chem. Sci. 2019, 10, 6001–6006. [Google Scholar] [CrossRef]
  53. Hu, X.; Jian, J.; Fang, Z.; Zhong, L.; Yuan, Z.; Yang, M.; Ren, S.; Zhang, Q.; Chen, X.; Yu, D. Hierarchical assemblies of conjugated ultrathin COF nanosheets for high-sulfur-loading and long-lifespan lithium–sulfur batteries: Fully-exposed porphyrin matters. Energy Storage Mater. 2019, 22, 40–47. [Google Scholar] [CrossRef]
  54. Liu, X.F.; Chen, H.; Wang, R.; Zang, S.Q.; Mak, T.C. Cationic Covalent-Organic Framework as Efficient Redox Motor for High-Performance Lithium–Sulfur Batteries. Small 2020, 16, 2002932. [Google Scholar] [CrossRef]
  55. Gao, G.; Jia, Y.; Gao, H.; Shi, W.; Yu, J.; Yang, Z.; Dong, Z.; Zhao, Y. New covalent triazine framework rich in nitrogen and oxygen as a host material for lithium–sulfur batteries. ACS Appl. Mater. Interfaces 2021, 13, 50258–50269. [Google Scholar] [CrossRef]
  56. Wang, D.G.; Li, N.; Hu, Y.; Wan, S.; Song, M.; Yu, G.; Jin, Y.; Wei, W.; Han, K.; Kuang, G.C. Highly fluoro-substituted covalent organic framework and its application in lithium–sulfur batteries. ACS Appl. Mater. Interfaces 2018, 10, 42233–42240. [Google Scholar] [CrossRef]
  57. Yang, Z.; Peng, C.; Meng, R.; Zu, L.; Feng, Y.; Chen, B.; Mi, Y.; Zhang, C.; Yang, J. Hybrid anatase/rutile nanodots-embedded covalent organic frameworks with complementary polysulfide adsorption for high-performance lithium–sulfur batteries. ACS Cent. Sci. 2019, 5, 1876–1883. [Google Scholar] [CrossRef] [PubMed]
  58. Kim, J.; Elabd, A.; Chung, S.Y.; Coskun, A.; Choi, J.W. Covalent triazine frameworks incorporating charged polypyrrole channels for high-performance lithium–sulfur batteries. Chem. Mater. 2020, 32, 4185–4193. [Google Scholar] [CrossRef]
  59. Haldar, S.; Wang, M.; Bhauriyal, P.; Hazra, A.; Khan, A.H.; Bon, V.; Isaacs, M.A.; De, A.; Shupletsov, L.; Boenke, T. Porous dithiine-linked covalent organic framework as a dynamic platform for covalent polysulfide anchoring in lithium–sulfur battery cathodes. J. Am. Chem. Soc. 2022, 144, 9101–9112. [Google Scholar] [CrossRef]
  60. Wang, H.; Jiang, J.; Wan, T.; Luo, Y.; Liu, G.; Li, J. A COF-coated ordered porous framework as multifunctional polysulfide barrier towards high-performance lithium-sulfur batteries. J. Colloid Interface Sci. 2023, 638, 542–551. [Google Scholar] [CrossRef]
  61. Yan, R.; Mishra, B.; Traxler, M.; Roeser, J.; Chaoui, N.; Kumbhakar, B.; Schmidt, J.; Li, S.; Thomas, A.; Pachfule, P. A Thiazole-linked Covalent Organic Framework for Lithium-Sulphur Batteries. Angew. Chem. Int. Ed. 2023, 62, e202302276. [Google Scholar] [CrossRef]
  62. Liu, X.; Ding, X.; Zheng, T.; Jin, Y.; Wang, H.; Yang, X.; Yu, B.; Jiang, J. Single Cobalt Ion-Immobilized Covalent Organic Framework for Lithium–Sulfur Batteries with Enhanced Rate Capabilities. ACS Appl. Mater. Interfaces 2024, 16, 4741–4750. [Google Scholar] [CrossRef] [PubMed]
  63. Xu, S.; Zhu, M.; Jin, Z.; Zhang, Y.; Yan, X.; Peng, H.; Song, Y.; Dong, Y.; Zhang, J. Regulation of covalent organic framework pore structure for boosting electrochemical performances of lithium-sulfur battery. J. Colloid Interface Sci. 2025, 138154. [Google Scholar] [CrossRef] [PubMed]
  64. Pathak, D.D.; Neem, M.; Kumar, G.; Modak, B.; Neogi, S.; Mandal, B.P. Diamondoid covalent organic framework-MXene composite for cathode host in lithium sulfur battery. J. Energy Storage 2025, 117, 116176. [Google Scholar] [CrossRef]
Figure 1. XRD analysis of pristine CNT, g-C3N4, g-C3N4/CNT composite, and sulfur-loaded S@g-C3N4/CNT.
Figure 1. XRD analysis of pristine CNT, g-C3N4, g-C3N4/CNT composite, and sulfur-loaded S@g-C3N4/CNT.
Energies 18 04462 g001
Figure 2. (a) SEM image of g-C3N4; (b,c) SEM images of g-C3N4/CNT; (df) SEM image of g-C3N4/CNT and corresponding elemental mapping.
Figure 2. (a) SEM image of g-C3N4; (b,c) SEM images of g-C3N4/CNT; (df) SEM image of g-C3N4/CNT and corresponding elemental mapping.
Energies 18 04462 g002
Figure 3. TEM (a,b) and HRTEM (c,d) images of the g-C3N4/CNT composite.
Figure 3. TEM (a,b) and HRTEM (c,d) images of the g-C3N4/CNT composite.
Energies 18 04462 g003
Figure 4. Structural porosity analysis: (a) N2 sorption isotherm and (b) mesopore distribution profile (g-C3N4/CNT).
Figure 4. Structural porosity analysis: (a) N2 sorption isotherm and (b) mesopore distribution profile (g-C3N4/CNT).
Energies 18 04462 g004
Figure 5. Surface and thermal analysis: (a) XPS survey, (b) C 1 s, (c) N 1 s, (d) TGA (S@g-C3N4 vs. S@g-C3N4/CNT).
Figure 5. Surface and thermal analysis: (a) XPS survey, (b) C 1 s, (c) N 1 s, (d) TGA (S@g-C3N4 vs. S@g-C3N4/CNT).
Energies 18 04462 g005
Figure 6. (a) LSV profiles; (b,c) CV curves; (d,e) Li2S deposition profiles; (f) Illustration of the working mechanism of g-C3N4/CNT.
Figure 6. (a) LSV profiles; (b,c) CV curves; (d,e) Li2S deposition profiles; (f) Illustration of the working mechanism of g-C3N4/CNT.
Energies 18 04462 g006
Figure 7. Cyclic voltammetry at multiple scan rates: (a) S@g-C3N4/CNT and (b) S@g-C3N4 cathodes. Corresponding peak current linear regressions: (c) S@g-C3N4/CNT, (d) S@g-C3N4 cells.
Figure 7. Cyclic voltammetry at multiple scan rates: (a) S@g-C3N4/CNT and (b) S@g-C3N4 cathodes. Corresponding peak current linear regressions: (c) S@g-C3N4/CNT, (d) S@g-C3N4 cells.
Energies 18 04462 g007
Figure 8. Electrochemical performance of S@g-C3N4/CNT vs. S@g-C3N4: (a) Cycling stability at 0.2 C (100 cycles), (b) galvanostatic profiles (cycles 1/10/20/50/100 @0.2C), (c) rate capability (0.2–2 C), (d) EIS plots, (e) Long-term cycling at 1C.
Figure 8. Electrochemical performance of S@g-C3N4/CNT vs. S@g-C3N4: (a) Cycling stability at 0.2 C (100 cycles), (b) galvanostatic profiles (cycles 1/10/20/50/100 @0.2C), (c) rate capability (0.2–2 C), (d) EIS plots, (e) Long-term cycling at 1C.
Energies 18 04462 g008
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chen, Z.; Meng, H.; Wang, J.; Yang, L.; Wang, X.; Chen, Z. High-Efficiency Polysulfide Trapping with g-C3N4/CNT Hybrids for Superior Lithium-Sulfur Batteries. Energies 2025, 18, 4462. https://doi.org/10.3390/en18174462

AMA Style

Chen Z, Meng H, Wang J, Yang L, Wang X, Chen Z. High-Efficiency Polysulfide Trapping with g-C3N4/CNT Hybrids for Superior Lithium-Sulfur Batteries. Energies. 2025; 18(17):4462. https://doi.org/10.3390/en18174462

Chicago/Turabian Style

Chen, Zhen, Hao Meng, Jiayi Wang, Lin Yang, Xin Wang, and Zhongwei Chen. 2025. "High-Efficiency Polysulfide Trapping with g-C3N4/CNT Hybrids for Superior Lithium-Sulfur Batteries" Energies 18, no. 17: 4462. https://doi.org/10.3390/en18174462

APA Style

Chen, Z., Meng, H., Wang, J., Yang, L., Wang, X., & Chen, Z. (2025). High-Efficiency Polysulfide Trapping with g-C3N4/CNT Hybrids for Superior Lithium-Sulfur Batteries. Energies, 18(17), 4462. https://doi.org/10.3390/en18174462

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop