Next Article in Journal
Influence of Content and Type of Lanthanide on the Structure of Ln2O3-Covered Carbon Nanoflakes: The EPR and XPS Study
Previous Article in Journal
Perovskite Solar Cells Modified with Conjugated Self-Assembled Monolayers at Buried Interfaces
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Design of the Ni3N/Nb4N5 Heterostructure as Bifunctional Adsorption/Electrocatalytic Materials for Lithium–Sulfur Batteries

1
Key Laboratory of Fine Chemical Application Technology of Luzhou, Sichuan Vocational College of Chemical Technology, Luzhou 646300, China
2
Key Laboratory of High-Tech Research on Power Batteries and Energy Storage of Zhenjiang, Institute for Energy Research, Jiangsu University, Zhenjiang 212013, China
3
Huizhou Research Institute, Sun Yat-sen University, Huizhou 516081, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2025, 15(13), 1015; https://doi.org/10.3390/nano15131015
Submission received: 15 May 2025 / Revised: 16 June 2025 / Accepted: 20 June 2025 / Published: 1 July 2025
(This article belongs to the Section Energy and Catalysis)

Abstract

Lithium–sulfur (Li-S) batteries are hindered by the sluggish electrochemical kinetics and poor reversibility of lithium polysulfides (LiPSs), which limits their practical energy density and cycle life. In order to address this issue, a novel Ni3N/Nb4N5 heterostructure was synthesized via electrospinning and nitridation as a functional coating for polypropylene (PP) separators. Adsorption experiments were conducted in order to ascertain the heterostructure’s superior affinity for LiPSs, thereby effectively mitigating their shuttling. Studies of Li2S nucleation demonstrated the catalytic role of the substance in accelerating the deposition kinetics of Li2S. Consequently, Li-S cells that employed the Ni3N/Nb4N5-modified separator were found to achieve significantly enhanced electrochemical performance, with the cells delivering an initial discharge capacity of 1294.4 mAh g−1 at 0.2 C. The results demonstrate that, after 150 cycles, the cells retained a discharge capacity of 796.2 mAh g−1, corresponding to a low capacity decay rate of only 0.25% per cycle. In addition, the rate capability of the cells was found to be improved in comparison to control cells with NiNb2O6-modified or pristine separators.

1. Introduction

The pursuit of high-energy-density, sustainable energy storage technologies has prompted extensive research into metal–air batteries (MABs) and lithium–sulfur (Li-S) batteries as promising alternatives to conventional lithium-ion batteries (LIBs). Although theoretical analyses indicate that MABs, including lithium–air (Li-air) and sodium–air (Na-air) systems, possess the potential to exhibit exceptional energy densities, their practical deployment remains constrained by critical challenges [1,2,3,4].
In contrast, Li-S batteries offer a compelling balance of theoretical energy density (2600 Wh/kg, six times higher than lithium-ion (Li-ion) batteries), material abundance (sulfur is inexpensive and abundant in the earth’s crust), and environmental benignity [5,6]. However, several critical challenges remain unresolved. Among these, the design of sulfur-containing host materials has been demonstrated as an effective strategy to enhance cycling stability, as evidenced by extensive studies [7,8,9,10]. Numerous host materials, such as metal oxides [11,12], sulfides [13], and nitrides [14,15,16], as well as metallic species with distinct dispersion states (e.g., nanoparticles [17,18,19] and single-atom catalysts [20,21]), have exhibited catalytic effects in facilitating the rapid and efficient conversion of long-chain lithium polysulfide to Li2S. It is well known that the size of the electrocatalyst significantly affects its catalytic activity in non-homogeneous catalysis, there is no theory to explain and predict the behavior of electrocatalysts with different particle sizes in different reactions, and the dependence of the kinetics of sulfur redox electrocatalysis on the degree of aggregation of the electrocatalyst atoms is yet to be explored [22], which is also crucial for understanding the underlying basis of electrocatalysis in Li-S batteries.
Since Manthiram et al. introduced the concept of employing functional interlayers or coatings between the cathode and separator to suppress the polysulfide shuttle effect and enhance sulfur utilization in 2012 [23], significant efforts have been dedicated to developing advanced interlayer materials. Transition metal compounds [24,25,26,27,28,29,30,31,32,33,34] (e.g., oxides, nitrides, sulfides, and selenides) have emerged as promising candidates due to their strong polar-polar interactions with LiPSs and superior chemisorption capabilities [35], thereby effectively improving cell stability.
In some cases, materials with identical chemical compositions may exhibit significantly distinct charge transfer mechanisms or thermodynamic properties, primarily governed by particle size variations [23]. Electrodes composed of smaller particles generally demonstrate higher electrical conductivity, which is critical for minimizing ohmic losses during cycling [36]. Furthermore, reducing the particle size of electrode materials typically enhances capacity retention at elevated current rates. This phenomenon can be rationalized through Fick’s first law of diffusion and stochastic process theory [37], where the diffusion length is expressed as L∝ D × t , with D representing the diffusion coefficient of charge carriers and t denoting the characteristic diffusion time. Decreasing the particle size effectively reduces the diffusion time required for lithium-ion intercalation into the electrode bulk, thereby improving charge–discharge kinetics.
In recent years, heterojunction engineering has emerged as a promising strategy for modulating the electronic structure of catalytic materials. This approach leverages the unique capability of heterojunctions to regulate band structures, accelerate charge transfer, and enhance catalytic conversion from LiPS (lithium polysulfides) to Li2S. Table 1 lists some recent heterojunction applications of the interlayers in Li-S batteries. As seen, Li et al. [38] synthesized a WS2−x/ZnS@C composite, where the constructed heterojunction not only facilitates charge transfer and ion diffusion but also improves reaction kinetics, thereby boosting overall electrochemical performance. Similarly, Li et al. [39] reported a Co9S8@MoS2 heterojunction material with pronounced energy-level separation. The impurity energy levels within the bandgap of Co9S8 were optimized, which enhanced its metallic characteristics and electrical conductivity. Beyond electronic modulation, heterojunctions also contribute to suppressing the shuttle effect, strengthening interfacial interactions, and enabling synergistic functionalities. Zhou et al. [40] developed a TiO2-TiN heterojunction as a dense and ultrathin separator coating, which effectively restricts LiPS shuttling even under high sulfur loading, enabling seamless LiPS capture–diffusion–conversion across interfaces. Ren et al. [41] engineered 2D/2D Ti3C2Tx/Ni-Co MOF heterojunction nanosheets coated on separators, achieving dual functions of shuttle suppression and synergistic adsorption–electrocatalysis. Yao et al. [42] integrated SnS (with strong adsorption and high conductivity) and ZnS (with superior catalytic activity) into a ZnS-SnS@NC heterojunction encapsulated by an N-doped carbon shell. This architecture exhibited enhanced lithiophilic and sulfiphilic properties, significantly suppressing the shuttle effect while improving sulfur utilization and cycle stability. Lei et al. [43] designed a MoS2-MoO3 heterojunction nanosheet that shortens Li⁺/electron transport pathways. Its strong donor–acceptor interactions enhance interfacial electrochemical reactions, while the built-in electric field accelerates charge transfer and LiPS conversion. Moreover, the synergistic interplay between MoO3 and MoS2 provides abundant adsorption/catalytic sites, further mitigating the shuttle effect and boosting sulfur utilization. Despite these advances, individual components in heterojunctions often struggle to integrate multiple advantages, such as abundant catalytic active sites, robust chemical adsorption, high conductivity, and rapid Li⁺ diffusion. Theoretical insights suggest that rational heterojunction design could holistically unify these merits, paving the way for next-generation electrocatalytic systems with unprecedented performance [44].
In this study, we investigated the catalytic and adsorption mechanisms of metal nitride heterostructure toward lithium polysulfides (LiPSs). The Ni3N/Nb4N5 heterostructure exhibits a uniform dispersion and porous architecture, which facilitates ion transport and provides abundant active sites for LiPS conversion. The visualization of adsorption tests and X-ray photoelectron spectroscopy (XPS) analysis confirmed the strong chemisorption capability of Ni3N/Nb4N5 toward LiPSs. Furthermore, the constructed heterogeneous interface facilitates accelerated charge transfer kinetics, while the synergistic catalytic effect between bimetallic nitrides promotes the homogeneous nucleation and growth of Li2S. Electrochemical evaluations demonstrate that the Ni3N/Nb4N5-PP-based cell achieved enhanced cycling stability and rate capability, delivering discharge capacities of 1403.6, 1032.4, 909.5, and 793.9 mAh g−1 at 0.1 C, 0.25 C, 0.5 C, and 1.0 C, respectively. Notably, the Li-S cell retained a discharge capacity of 796.2 mAh g−1 after 150 cycles at 0.2 C.

2. Experimental Procedures

2.1. Materials

Niobium oxalate hydrate: Alfa Aesar Chemical Company (Shanghai, China); N,N-dimethylformamide: Sinopharm Chemical Reagent Company (Shanghai, China); nickel acetate tetrahydrate: Sinopharm Chemical Reagent Company (Shanghai, China); peroxyacetyl nitrate: Aladdin ReagentCompany (Shanghai China); electrolyte: DuoDuo Chemical Reagent (Suzhou, China); polyvinylidene fluoride: Aldrich Chemical Company (Shanghai, China); N-methylpyrrolidone: Sigma Chemical Reagent Company (Shanghai, China).

2.2. Fabrication of Ni₃N/Nb₄N₅ Heterostructure

The fabrication of the Ni3N/Nb4N5 heterostructure was achieved through a sequential process that entailed the combination of electrospinning and controlled thermal treatments. The synthesis protocol that was utilized is outlined as follows:
Precursor solution preparation: The formation of a homogeneous solution was achieved through the dissolution of 2 mmol of niobium oxalate hydrate (C10H5NbO20) in 4 mL of DMF, under constant stirring at 60 °C. Subsequently, 1 mmol of nickel acetate tetrahydrate (Ni(CH3COO)2·4H2O) was added, and the mixture was stirred for a further 8 h.
Polymer matrix preparation: Concurrently, 1 g of polyacrylonitrile (PAN) was dissolved in 6 mL of DMF, with the solution being subjected to mechanical stirring at ambient temperature for a duration of 8 h.
Electrospinning process: The two solutions were thoroughly blended prior to processing through the electrospinning apparatus.
Thermal conversion: The electrospun fibers were then subjected to oxidative calcination at 800 °C (with a 5 °C/min ramp rate) for a duration of three hours in air, with the objective of producing the NiNb2O6 intermediate.
Nitridation: The final transformation was achieved by subjecting the NiNb2O6 precursor to ammoniating at 800 °C for 3 h (with a heating rate of 5 °C/min) under NH3/Ar flow, thereby yielding the desired Ni3N/Nb4N5 nanocomposite.

2.3. Preparation of Modified Separators

The electrode slurry was prepared by means of homogenously blending active materials (Ni3N/Nb4N5 heterostructure or alternatively NiNb2O6), a conductive additive (Super P carbon black), and a polymeric binder (PVDF) in an 8:1:1 mass ratio through mechanical grinding. The subsequent addition of N-methyl-2-pyrrolidone (NMP) as a dispersion medium enabled uninterrupted magnetic stirring for a period of three hours, thereby achieving a well-dispersed slurry. The slurry was uniformly coated onto a polypropylene (PP, Celgard 2400) separator using a 200 μm coating blade with a coating machine. Subsequently, the modified separator was dried in a vacuum oven at 50 °C for 12 h. Finally, the separator was punched into 19 mm diameter disks, with an active material (Ni3N/Nb4N5 or NiNb2O6) loading of 1.3–1.9 mg cm−2. The Ni3N/Nb4N5 and NiNb2O6 modified separators were abbreviated as Ni3N/Nb4N5-PP and NiNb2O6-PP.

2.4. Visualized Adsorption of Polysulfides

This section will deal with the Li2S6 solution synthesis. The solvent system was initially prepared by combining 1,3-dioxolane (DOL) and dimethoxyethane (DME) at a volumetric ratio of 1:1. Subsequently, stoichiometric amounts of Li2S and elemental sulfur (in a 1:5 molar ratio) were dissolved in the DOL/DME mixture. The reaction system was maintained at a temperature of 60 °C for a period of 24 h, with the protection of argon, thereby yielding a solution of Li2S6 at a concentration of 5 mmol L−1. This solution was then diluted to a concentration of 2 mmol L−1. For the adsorption tests, 5 mL aliquots of the diluted solution were separately mixed with 10 mg portions of Ni3N/Nb4N5 and NiNb2O6 powders in an inert environment. The mixtures were maintained for a duration of 0~24 h to monitor their adsorption behavior, and images were captured at designated time intervals for further analysis.

2.5. Fabrication and Characterization of Symmetric Cells

The active material (Ni3N/Nb4N5 or NiNb2O6), conductive carbon black, and PVDF binder were then homogenously mixed in ethanol with a weight ratio of 8:1:1 under ultrasonication for 30 min to form a uniform electrode slurry. The homogenized slurry was uniformly deposited onto 12 mm diameter carbon fiber paper (Toray TGP-H-060,Toray Investment Co., Ltd., Shanghai, China)) using a micropipette, followed by drying at 50 °C for 12 h in a vacuum oven.
A CR2025 coin cell was assembled using a polypropylene (Celgard-2400) separator and 50 μL of Li2S6 electrolyte. The electrolyte was formulated by dissolving 1.0 M Li2S6 and 1.0 M LiTFSI in a 1:1 (v/v) binary solvent system comprising DOL and DME. The present study investigates the electrochemical behavior of LiPSs. To this end, cyclic voltammetry (CV) measurements were performed on the symmetric cell at scan rates of 5 and 50 mV s−1. The potential window was set between −0.8 and 0.8 V.

2.6. Experimental Procedure for Li₂S Nucleation Studies

The nucleation behavior was investigated using CR2025-type coin cells. Carbon paper substrates were coated with Ni3N/Nb4N5 or NiNb2O6 active materials, achieving mass loadings between 1.0 and 1.3 mg cm−2. Lithium foil acted as the counter electrode. In the context of the electrolyte system, the cathodic compartment was constituted of 0.25 M Li2S6 in conjunction with 1.0 M LiTFSI within a tetraethylene glycol dimethyl ether solvent, while the anodic compartment utilized a 1.0 M LiTFSI solution within an equivalent solvent. Precise 20 μL aliquots of each electrolyte solution were dispensed into their respective cell chambers. Electrochemical conditioning was initiated with a constant-current discharge at 2.06 V (0.112 mA). This was followed by a potentiostatic hold at 2.05 V, which was maintained until the current stabilized below 10−5 A. This process was found to promote Li2S crystallization [45].

2.7. Material Characterization

The crystalline structure was characterized by X-ray diffraction (XRD) analysis performed on a Bruker D8 Advance diffractometer equipped with a Cu Kα radiation source (wavelength = 1.54056 Å). The morphology and microstructure of the samples were characterized by scanning electron microscopy (SEM, JSM-7800F) and transmission electron microscopy (TEM, FEI Talos F200X G2, FEI Czech Ltd., Hillsboro, OR, USA), both equipped with energy-dispersive spectroscopy (EDS). The chemical composition of the sample was investigated through the implementation of X-ray photoelectron spectroscopy (XPS) analysis.

2.8. Material Preparation and Electrochemical Characterization

The sulfur/acetylene black (S/AB) composite was fabricated through mechanical ball-milling of sublimed sulfur and AB (3:2 mass ratio) for 2 h, followed by thermal treatment in a sealed reactor at 155 °C for 12 h. This facilitated sulfur melting and subsequent homogeneous distribution. Subsequent to cooling to ambient temperature, cathode slurries were prepared by blending the composite with PVDF binder (9:1 mass ratio) in NMP solvent. The homogeneous mixture was applied to carbon-modified Al current collectors by doctor-blading, followed by vacuum drying at 50 °C for 12 h. The electrodes were punched into 12 mm diameter disks, with sulfur loadings ranging from 1.0 to 1.5 mg cm−2. The assembly of coin cells was conducted within an Ar atmosphere glovebox, utilizing lithium metal anodes, prepared S/AB cathodes, and Ni3N/Nb4N5-PP (or an alternative) separators. The electrolyte formulation consisted of 1.0 M LiTFSI and 2 wt% LiNO3 in DME (1:1 v/v) solvent, with strict control of electrolyte volume at 30 μL per mg sulfur. Galvanostatic charge–discharge tests were performed on a LAND battery testing system within a voltage window of 1.7–2.8 V. Cyclic voltammetry (CV) was conducted at a scan rate of 0.1 mV s−1, and electrochemical impedance spectroscopy (EIS) was performed over a frequency range of 0.01 Hz to 100 kHz.

3. Results and Discussion

Figure 1 illustrates the preparation process of the NiNb2O6 nanorods and the Ni3N/Nb4N5 heterostructure. Initially, NiNb2O6 nanorods, composed of uniformly sized nanoparticles, were synthesized via the electrospinning method, demonstrating high homogeneity. Subsequently, the NiNb2O6 nanorods underwent nitridation treatment in an ammonia atmosphere to form the Ni3N/Nb4N5 heterostructure. After nitridation, the nanorods exhibited a well-dispersed morphology, leading to the formation of homogeneous Ni3N/Nb4N5 nanoparticles.
The crystal structure of the as-prepared NiNb2O6 and Ni3N/Nb4N5 heterostructure were analyzed and verified by XRD. As shown in Figure 2a, several typical diffraction peaks in the XRD pattern of NiNb2O6 are in good agreement with the reference spectrum of NiNb2O6 nanorods (JCPDS No. 76-2354), confirming the successful synthesis of crystalline NiNb2O6. In addition, Figure 2b shows the XRD pattern of the Ni3N/Nb4N5 heterostructure, where the major diffraction peaks correspond to Nb4N5 (JCPDS No. 74-0606) and Ni3N (JCPDS No. 89-5144). The presence of non-impurity peaks indicates that the pure Ni3N/Nb4N5 heterostructure was successfully synthesized through the nitridation of NiNb2O6.
The microstructure of NiNb2O6 and Ni3N/Nb4N5 heterostructure, as well as the cross-section of the modified separator, were observed using scanning electron microscopy (SEM). As shown in Figure 3a, the NiNb2O6 exhibits a regular nanorod structure. In contrast, the nitrided Ni3N/Nb4N5 heterostructure primarily displays a nanoparticle structure (Figure 3b), with uniform dispersion and good homogeneity. Moreover, the nanoparticle stacking structure features a large surface area, allowing the space between nanoparticles to be effectively infiltrated and stored with electrolyte, thereby enhancing the electron exchange rate [46]. A slurry was prepared to coat the commercial PP separator, and the thickness of the coated layer was approximately 23.65 μm, as observed in the cross-sectional view (Figure 3c). Furthermore, the TEM and EDS mapping in Figure 3d–h confirm the presence of Nb, Ni, and N elements in the Ni3N/Nb4N5 heterostructure, indicating the successful incorporation of these elements. As demonstrated in Figure 3i, high-resolution TEM analysis clearly identified two characteristic interplanar distances measuring 0.250 nm and 0.203 nm, respectively. These were indexed to the (211) crystallographic plane of Nb4N5 and the (111) plane of Ni₃N, thus confirming the coexistence of both nitride phases. Additionally, a lattice spacing of 0.215 nm was observed, which corresponds to the (002) plane present in both Ni3N and Nb4N5. These findings are consistent with the reference data for Nb4N5 (JCPDS No. 74-0606) and Ni3N (JCPDS No. 89-5144) at 42° in the XRD spectra. Therefore, the Ni3N/Nb4N5 heterostructure was successfully synthesized through the nitridation of NiNb2O6.
In order to undertake a systematic investigation into the immobilization performance of LiPSs of the Ni3N/Nb4N5 composite, comparative visual adsorption experiments were performed under controlled conditions. The quantitative results were then documented visually, as illustrated in Figure 4a–c. After 12 h of standing, the yellow color of the Li2S6 solution containing Ni3N/Nb4N5 powder partially faded, whereas the solution with NiNb2O6 powder exhibited almost no color change. After 24 h, the yellow color of the Li2S6 solution with Ni3N/Nb4N5 powder completely disappeared, while the solution containing NiNb2O6 showed only slight fading [47,48]. These results demonstrate that Ni3N/Nb4N5 exhibits a significantly stronger adsorption capability for LiPSs compared to NiNb2O6.
To further investigate the interaction mechanism between Ni3N/Nb4N5 and LiPSs, XPS analysis was conducted on the samples before and after Li2S6 adsorption. Figure 5 presents the high-resolution XPS spectra of N 1s, Nb 3d, and Ni 2p before and after Li2S6 adsorption by Ni3N/Nb4N5. In Figure 5a,d, the N 1s spectra exhibit three characteristic peaks corresponding to the Ni-N bond, Nb-N bond, and graphitic nitrogen. Following the adsorption of Li2S6, a shift in the peak positions of the three characteristic peaks was observed. The original peak positions of 397.03 eV/396.14 eV/398.84 eV shifted to higher binding energies of 0.11 eV, 0.31 eV, and 0.22 eV, respectively. In Figure 5b–e, the Nb 3d spectra display three pairs of characteristic peaks, assigned to Nb-O, Nb-N-O, and Nb-N bonds. The oxygen signal likely originates from surface oxidation due to exposure to air [37]. After the adsorption of Li2S6, the binding energies of Nb-O and Nb-N showed negligible changes; however, the Nb-N-O binding energy shifted to the right by 0.32 eV from the original 208.18 eV/205.38 eV. Previous studies have demonstrated that surface oxide layers on metal compounds can activate surface metal sites through strong chemical bonds for polysulfide bonding [49]. Similarly, the Ni 2p spectra (Figure 5c,f) show three pairs of characteristic peaks, including satellite peaks, Ni-O bonds, and Ni-N bonds. After Li2S6 adsorption, the binding energies of the Ni-O and Ni-N bonds increased by 0.45 eV and 0.29 eV, respectively.
Compared to the Ni3N/Nb4N5 heterostructure, the XPS high-resolution spectra of Nb 3d, O 1s [50,51], and Ni 2p [52,53] before/after the NiNb2O6 adsorption of Li2S6 are shown in Figure 6. No significant change in binding energy was observed before and after adsorption, indicating that Ni3N/Nb4N5 exhibited stronger chemical adsorption capability for lithium LiPSs. This enhanced adsorption effectively suppressed the shuttle effect in Li-S batteries and improved the utilization of sulfur.
In addition, the high-resolution S 2p spectra of both Ni3N/Nb4N5 and NiNb2O6 (Figure 7a,b) exhibit three pairs of characteristic peaks. The binding energy of terminal sulfur S T - 1 in Ni3N/Nb4N5 (162.00 eV) was slightly higher than that in NiNb2O6 (161.94 eV). Similarly, the binding energy of bridging sulfur ( S B 0 ) in Ni3N/Nb4N5 (163.70 eV) was higher than that in NiNb2O6 (163.63 eV). Additionally, the peak observed at 169.02 eV for Ni3N/Nb4N5 can be assigned to thiosulfate species [54]. Compared with NiNb2O6, strong thiosulfate signals were detected in Ni3N/Nb4N5, which could promote the conversion of higher-order LiPSs into short-chain lithium sulfides. This further demonstrates that Ni3N/Nb4N5 plays a critical role in enhancing the kinetics of sulfide conversion, whereas only weak thiosulfate signals were observed in NiNb2O6 [55]. As demonstrated by visual adsorption tests and XPS characterization, the heterogeneous interface of Ni3N/Nb4N5 facilitates charge transfer and exhibits enhanced thiophilicity, which promotes the adsorption of lithium LiPSs and effectively suppresses the shuttle effect in Li-S batteries.
To investigate the effect of the Ni3N/Nb4N5-modified separator on the electrochemical performance of Li-S batteries, the samples were coated onto a commercial separator for modification and the Li-S cells were assembled for electrochemical testing. Figure 8a presents the cyclic voltammetry (CV) curves of symmetric cells with different electrodes at a scan rate of 5 mV s−1. It can be observed that the Ni3N/Nb4N5 electrode exhibits higher redox current peaks and smaller redox peak separation compared to the NiNb2O6 electrode, whereas the CV curve of the NiNb2O6 cell shows greater potential hysteresis. The initial cycle of CV tests conducted at a scan rate of 0.1 mV s−1 for the Li-S cells assembled with Ni3N/Nb4N5-PP, NiNb2O6-PP, and PP separators are shown in Figure 8b. The two reduction peaks observed at approximately 2.27 V and 1.97 V correspond to the conversion of S8 to long-chain LiPSs and to the final products Li2S2/Li2S, respectively. Additionally, the anodic peaks around 2.36 V and 2.48 V correspond to the conversion of Li2S2/Li2S to long-chain LiPSs and S8. The Ni3N/Nb4N5-PP separator exhibits three distinct electrochemical advantages in comparison to both NiNb2O6-PP and PP separators. The following observations were made: firstly, there was a shift in the reduction peaks toward higher potentials; secondly, there was a decrease in the overpotential gap; and thirdly, there was a shift in the oxidation peaks to lower potentials. This combination of effects led to significantly lower charge–discharge overpotentials. This electrochemical behavior indicates that the Ni3N/Nb4N5 heterostructure effectively promotes LiPS redox kinetics [56]. In order to conduct a more in-depth investigation into the interfacial characteristics, we undertook electrochemical impedance spectroscopy (EIS) measurements. The objective of this was to ascertain the charge transfer resistance (Rct) between the electrode and the electrolyte. As illustrated in Figure 8c, the resulting Nyquist plots for all cell configurations are presented. The results revealed that the Rct of the Ni3N/Nb4N5-PP cell (45.09 Ω) was significantly lower than that of the NiNb2O6-PP (72.12 Ω) and PP (242.1 Ω) cells, suggesting that Ni3N/Nb4N5 accelerates electron transfer and reaction kinetics at the electrode–electrolyte interface [57].
To investigate the capability of Ni3N/Nb4N5 nanoparticles in catalyzing the conversion of LiPSs Li2S, nucleation experiments were conducted. As illustrated in Figure 9a,b, the initial current drop is attributed to the reduction of higher-order LiPSs in the solution to Li2S4 [58]. The Ni3N/Nb4N5 sample exhibited a significantly shorter nucleation time (4800 s) compared to NiNb2O6 (9300 s), indicating faster kinetics of Li2S nucleation. Additionally, Ni3N/Nb4N5 demonstrated a stronger Li2S nucleation capability than NiNb2O6. This comparison suggests that the catalytic effect of Ni3N/Nb4N5 can lower the energy barrier for Li2S growth, thereby enhancing the conversion of soluble LiPSs to insoluble solid Li2S [59].
Furthermore, the cycling stability of Li-S cells assembled with Ni3N/Nb4N5-PP, NiNb2O6-PP, and PP separators was evaluated. The materials under consideration were Ni3N/Nb4N5-PP, NiNb2O6-PP, and standard PP. Initial cycling tests conducted at 0.1 C rate revealed substantial differences in discharge capacities, with the Ni3N/Nb4N5-PP -PP cell achieving 1403.6 mAh g−1, outperforming both the NiNb2O6-PP (1224 mAh g−1) and unmodified PP (948.6 mAh g−1) counterparts. The corresponding charge–discharge profiles for the initial three cycles are illustrated in Figure 10a–c. It is important to note the observed potential difference at 800 mAh g−1 capacity, where the Ni3N/Nb4N5-PP system exhibited a minimal 0.17 V polarization, which is substantially lower than the 0.2 V and 0.24 V values recorded for the NiNb2O6-PP and PP systems, respectively. The reduced polarization suggests that the electrocatalytic activity of Ni3N/Nb4N5 decreases the kinetic resistance of Li-S batteries, with Li-S cells with Ni3N/Nb4N5-PP demonstrating low polarization and superior reversibility [14]. From the long-term cycling performance depicted in Figure 10d, it is evident that the Li-S cell with Ni3N/Nb4N5-PP exhibited an initial discharge capacity of 1294.4 mAh g−1 at 0.2 C, and after 150 cycles, the discharge capacity was retained at 796.2 mAh g−1, with a capacity decay rate of 0.25% per cycle. Electrochemical performance comparisons revealed significant differences among the tested separators. The NiNb2O6-PP configuration demonstrated an initial discharge capacity of 1011.9 mAh g−1, which is considerably lower than that of the Ni3N/Nb4N5-PP system. This observation serves to confirm the enhanced cycling stability of the latter. This performance discrepancy may originate from the inferior conductivity of NiNb2O6, as reported in previous studies [60,61]. It has been demonstrated that the Ni3N/Nb4N5-PP cells exhibit superior cycling stability. This can be attributed to the formation of heterostructured interfaces, which enhance electron transport kinetics and reduce polarization loss during cycling. Furthermore, the high electrical conductivity of Ni3N promotes the formation of a more homogeneous SEI layer, which inhibits the continuous decomposition of the electrolyte [62]. Testing the rate capability of the material across the range of 0.1–1.5 C (Figure 10e) demonstrated the exceptional performance of the Ni3N/Nb4N5-PP, with capacities of 1403.6 mAh g−1 (0.1 C), 1032.4 mAh g−1 (0.25 C), 909.5 mAh g−1 (0.5 C), and 793.9 mAh g−1 (1.0 C) being maintained. It is noteworthy that even at the maximum 1.5 C rate, this system demonstrated a capacity of 695.4 mAh g−1, which surpassed the capacities of both NiNb2O6-PP (634.5 mAh g−1) and PP (235.7 mAh g−1) systems. Upon returning to 0.1 C, the Ni3N/Nb4N5-PP cell exhibited a recovery of 949.4 mAh g−1, thereby further substantiating its superior rate adaptability and capacity retention in comparison to alternative configurations.

4. Conclusions

In this study, the Ni3N/Nb4N5 heterostructure was synthesized using electrospinning followed by nitridation treatment and applied as a modified separator for Li-S batteries. The Ni3N/Nb4N5 heterostructure was uniformly dispersed and possessed abundant porosity, which facilitated ion transport. The visual adsorption tests and XPS results demonstrated that the Ni3N/Nb4N5 had a strong adsorption capability for LiPSs, and the heterointerface promoted accelerated charge transfer. The Li2S nucleation tests confirmed that the Ni3N/Nb4N5 enhanced the nucleation and growth of Li2S. The electrochemical measurements revealed that the Li-S cell with Ni3N/Nb4N5-PP could deliver an initial discharge capacity of 1294.4 mAh g−1 at 0.2 C, with a capacity decay rate of 0.25% per cycle after 150 cycles, demonstrating superior cycling stability and rate performance compared to the Li-S cells with NiNb2O6-PP and pristine separator. This study introduces a novel method for fabricating the Ni3N/Nb4N5 heterostructure, which markedly enhances the electrochemical performance of Li-S batteries.

Author Contributions

X.L.: writing—original draft preparation, writing—review and editing. W.S.: validation. S.Z.: writing—original draft preparation, data collection. C.X.: conceptualization, supervision. J.L.: project administration, writing—review and editing. G.L.: resources, supervision, funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Natural Science Foundation of Guangdong Province (2023A1515010894), the Science and Technology Plan Project of Luzhou (2024RCX211), the PhD Innovation Studio and Special Research Project of Sichuan Vocational College of Chemical Technology (SCHYCXGZS-2024-02/YJKT2025040305), the Development Project for Metal Ion Removal from Wastewater Using Microreactors, and the research project cooperated with Jiangyin Longjicheng Pipe Jacking Engineering Co., Ltd.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no competing financial interests.

References

  1. Li, Y.; Lu, J. Metal–air batteries: Will they be the future electrochemical energy storage device of choice. ACS Energy Lett. 2017, 2, 1370–1377. [Google Scholar] [CrossRef]
  2. Jiang, L.; Luo, X.; Wang, D.W. A review on system and materials for aqueous flexible metal–air batteries. Carbon Energy 2023, 5, e284. [Google Scholar] [CrossRef]
  3. Yaqoob, L.; Noor, T.; Iqbal, N. An overview of metal-air batteries, current progress, and future perspectives. J. Energy Storage 2022, 56, 106075. [Google Scholar] [CrossRef]
  4. Ahmed, S.; Ali, A.; Asif, M.; Shim, J.; Park, G. Exploring innovative trends and advancements in rechargeable zinc-air batteries. Inorg. Chem. Commun. 2024, 170, 113288. [Google Scholar] [CrossRef]
  5. Zhang, C.; Biendicho, J.J.; Zhang, T.; Du, R.; Li, J.; Yang, X.; Arbiol, J.; Zhou, Y.; Morante, J.R.; Cabot, A. Combined high catalytic activity and efficient polar tubular nanostructure in urchin-like metallic NiCo2Se4 for high-performance lithium–sulfur batteries. Adv. Funct. Mater. 2019, 29, 1903842. [Google Scholar]
  6. Peng, H.; Huang, J.; Cheng, X.; Zhang, Q. Review on High-Loading and High-Energy Lithium–Sulfur Batteries. Adv. Energy Mater. 2017, 7, 1700260. [Google Scholar] [CrossRef]
  7. Zhao, Y.; Wu, W.; Li, J.; Xu, Z.; Guan, L. Encapsulating MWNTs into hollow porous carbon nanotubes: A tube-in-tube carbon nanostructure for high-performance lithium-sulfur batteries. Adv. Mater. 2014, 26, 5113–5118. [Google Scholar] [CrossRef]
  8. Kim, S.C.; Gao, X.; Liao, S.L.; Su, H.; Chen, Y.; Zhang, W.; Greenburg, L.; Pan, J.; Zheng, X.; Ye, Y.; et al. Solvation-property relationship of lithium-sulphur battery electrolytes. Nat. Commun. 2024, 15, 1268. [Google Scholar] [CrossRef]
  9. Guo, D.; Yuan, M.; Zheng, X.; Li, M.; Nan, C.; Sun, G.; Huang, X.; Li, H. Ultralong cycle life enabled by in situ growth of CoMo1-xP/Mo heterostructure for lithium-sulfur batteries. J. Energy Chem. 2022, 73, 5–12. [Google Scholar] [CrossRef]
  10. Preefer, M.B.; Oschmann, B.; Hawker, C.J.; Seshadri, R.; Wudi, F. High sulfur content material with stable cycling in lithium-sulfur batteries. Angew. Chem. Int. Ed. 2017, 129, 15314–15318. [Google Scholar] [CrossRef]
  11. Wang, Y.; Zhang, R.; Chen, J.; Wu, H.; Lu, S.; Wang, K.; Li, H.; Harris, C.J.; Xi, K.; Kumar, R.V.; et al. Enhancing catalytic activity of titanium oxide in lithium-sulfur batteries by band engineering. Adv. Energy Mater. 2019, 9, 1900953. [Google Scholar] [CrossRef]
  12. Du, M.; Geng, P.; Pei, C.; Jiang, X.; Shan, Y.; Hu, W.; Ni, L.; Pang, H. High-entropy Prussian blue analogues and their oxide family as sulfur hosts for lithium-sulfur batteries. Angew. Chem. Int. Ed. 2022, 61, e202209350. [Google Scholar] [CrossRef] [PubMed]
  13. Razzaq, A.A.; Yuan, X.; Chen, Y.; Hu, J.; Mu, Q.; Ma, Y.; Zhao, X.; Miao, L.; Ahn, J.H.; Peng, Y.; et al. Anchoring MOF-derived CoS2 on sulfurized polyacrylonitrile nanofibers for high areal capacity lithium-sulfur batteries. J. Mater. Chem. A 2020, 8, 1298–1306. [Google Scholar] [CrossRef]
  14. Shen, Z.; Zhang, Z.; Li, M.; Yuan, Y.; Zhao, Y.; Zhang, S.; Zhong, C.; Zhu, J.; Lu, J.; Zhang, H. Rational design of a Ni3N0.85 electrocatalyst to accelerate polysulfide conversion in lithium-sulfur batteries. ACS Nano 2020, 14, 6673–6682. [Google Scholar] [CrossRef] [PubMed]
  15. Wang, R.; Jiao, J.; Liu, D.; He, Y.; Yang, Y.; Sun, D.; Pan, H.; Fnag, F.; Wu, R. High-entropy metal nitride embedded in concave porous carbon enabling polysulfide conversion in lithium-sulfur batteries. Small 2024, 20, 2405148. [Google Scholar] [CrossRef] [PubMed]
  16. Cheong, J.L.; Hu, C.; Liu, W.; Ng, M.F.; Sullivan, M.B.; Ying, J.Y. 3D carbonaceous nanostructured transition metal nitride, carbonitride and carbide as polysulfide regulators for lithium-sulfur batteries. Nano Energy 2022, 102, 107659. [Google Scholar] [CrossRef]
  17. Liu, R.; Liu, Z.; Liu, W.; Liu, Y.; Lin, X.; Li, Y.; Li, P.; Huang, Z.; Feng, X.; Yu, L.; et al. TiO2 and Co nanoparticle-decorated carbon polyhedra as efficient sulfur host for high-performance lithium-sulfur batteries. Small 2019, 15, 1804533. [Google Scholar] [CrossRef]
  18. Liu, Y.; Chatterjee, A.; Rusch, P.; Wu, C.; Nan, P.; Peng, M.; Bettels, F.; Li, T.; Ma, C.; Zhang, C.; et al. Monodisperse molybdenum nanoparticles as highly efficient electrocatalysts for Li-S batteries. ACS Nano 2021, 15, 15047–15056. [Google Scholar] [CrossRef]
  19. Cai, D.; Liu, B.; Zhu, D.; Chen, D.; Lu, M.; Cao, J.; Wang, Y.; Huang, W.; Shao, Y.; Tu, H.; et al. Ultrafine Co3Se4 nanoparticles in nitrogen-doped 3D carbon matrix for high-stable and long-cycle-life lithium sulfur batteries. Adv. Energy Mater. 2020, 10, 1904273. [Google Scholar] [CrossRef]
  20. Zhang, D.; Wang, S.; Hu, R.; Gu, J.; Cui, Y.; Li, B.; Chen, W.; Liu, C.; Shang, J.; Yang, S. Catalytic conversion of polysulfides on single atom zinc implanted MXene toward high-rate lithium-sulfur batteries. Adv. Funct. Mater. 2020, 30, 2002471. [Google Scholar] [CrossRef]
  21. Liu, K.; Wang, X.; Gu, S.; Yuan, H.; Jiang, F.; Li, Y.; Tan, W.; Long, Q.; Chen, J.; Xu, Z.; et al. N, S-coordinated Co single atomic catalyst boosting adsorption and conversion of lithium polysulfides for lithium-sulfur batteries. Small 2022, 18, 2204707. [Google Scholar] [CrossRef]
  22. Zhou, X.; Meng, R.; Zhong, N.; Yin, S.; Ma, G.; Liang, X. Size-dependent cobalt catalyst for lithium sulfur batteries: From single atoms to nanoclusters and nanoparticles. Small Methods 2021, 5, 2100571. [Google Scholar] [CrossRef] [PubMed]
  23. Zhang, J.; Li, Z.; Fu, Q.; Xue, Y.; Cui, Z. The size-dependence of electrochemical thermodynamics of metal nanoparticles electrodes in theory and experiment. J. Electrochem. Soc. 2017, 164, 828–835. [Google Scholar] [CrossRef]
  24. Zhang, L.; Chen, X.; Wan, F.; Niu, Z.; Wang, Y.; Zhang, Q.; Chen, J. Enhanced electrochemical kinetics and polysulfide traps of indium nitride for highly stable lithium-sulfur batteries. ACS Nano 2018, 12, 9578–9586. [Google Scholar] [CrossRef] [PubMed]
  25. Luo, D.; Zhang, Z.; Li, G.; Cheng, S.; Li, S.; Li, J.; Gao, R.; Li, M.; Sy, S.; Deng, Y.; et al. Revealing the rapid electrocatalytic behavior of ultrafine amorphous defective Nb2O5-x nanocluster toward superior Li-S performance. ACS Nano 2020, 14, 4849–4860. [Google Scholar] [CrossRef] [PubMed]
  26. Sun, T.; Zhao, X.; Li, B.; Shu, H.; Luo, L.; Xia, W.; Chen, M.; Zeng, P.; Yang, X.; Gao, P.; et al. NiMoO4 nanosheets anchored on N-S doped carbon clothes with hierarchical structure as a bidirectional catalyst toward accelerating polysulfides conversion for Li-S Battery. Adv. Funct. Mater. 2021, 31, 2101285. [Google Scholar] [CrossRef]
  27. Tian, D.; Song, X.; Wang, M.; Wu, X.; Qiu, Y.; Guan, B.; Xu, X.; Fang, L.; Zhang, N.; Sun, K. MoN supported on graphene as a bifunctional interlayer for advanced Li-S batteries. Adv. Energy Mater. 2019, 9, 1901940. [Google Scholar] [CrossRef]
  28. Zhang, H.Y.; Dai, R.Q.; Zhu, S.; Zhou, L.Z.; Xu, Q.J.; Min, Y.L. Bimetallic nitride modified separator constructs internal electric field for high-performance lithium-sulfur battery. Chem. Eng. J. 2022, 429, 132454. [Google Scholar] [CrossRef]
  29. Yan, Y.; Li, H.; Cheng, C.; Yan, T.; Gao, W.; Mao, J.; Dai, K.; Zhang, L. Boosting polysulfide redox conversion of Li-S batteries by one-step-synthesized Co-Mo bimetallic nitride. J. Energy Chem. 2021, 61, 336–346. [Google Scholar] [CrossRef]
  30. He, J.R.; Chen, Y.F.; Manthiram, A. Vertical Co9S8 hollow nanowall arrays grown on a Celgard separator as a multifunctional polysulfide barrier for high-performance Li-S batteries. Energy Environ. Sci. 2018, 11, 2560–2568. [Google Scholar] [CrossRef]
  31. Sun, L.; Liu, Y.; Xie, J.; Fan, L.; Wu, J.; Jiang, R.; Jin, Z. Polar Co9S8 anchored on pyrrole-modified graphene with in situ growth of CNTs as multifunctional self-supporting medium for efficient lithium-sulfur batteries. Chem. Eng. J. 2023, 451, 138370. [Google Scholar] [CrossRef]
  32. Wang, D.; Guo, Q.; Hong, X.; Xie, K.; Ma, Z.; Zheng, C.; Xiong, S. Monodispersed FeS2 electrocatalyst anchored to nitrogen-doped carbon host for lithium-sulfur batteries. Adv. Funct. Mater. 2022, 32, 2205471. [Google Scholar]
  33. Jiang, X.; Zhang, S.; Zou, B.; Li, G.; Yang, S.; Zhao, Y.; Lian, J.; Li, H.; Ji, H. Electrospun CoSe@ NC nanofiber membrane as an effective polysulfides adsorption-catalysis interlayer for Li-S batteries. Chem. Eng. J. 2022, 430, 131911. [Google Scholar] [CrossRef]
  34. Zhao, M.; Chen, X.; Li, X.Y.; Li, B.Q.; Huang, J.Q. Lithium-sulfur batteries: An organodiselenide comediator to facilitate sulfur redox kinetics in lithium-sulfur batteries. Adv. Mater. 2021, 33, 2170100. [Google Scholar] [CrossRef]
  35. Wang, X.; Zhang, L.; Zhang, C.; Ma, X.; Jin, Q.; Li, L.; Zhang, Z.; Zhang, X.; Wu, L. Strengthened dipole-dipole interaction on high-entropy oxide electrocatalysts for high-rate and excellently stable lithium-sulfur batteries. J. Energy Chem. 2025, 105, 292–301. [Google Scholar] [CrossRef]
  36. Rai, A.K.; Anh, L.T.; Gim, J.; Mathew, V.; Kang, J.; Paul, B.J.; Song, J.; Kim, J. Simple synthesis and particle size effects of TiO2 nanoparticle anodes for rechargeable lithium ion batteries. Electrochim. Acta 2013, 90, 112–118. [Google Scholar] [CrossRef]
  37. Jung, S.K.; Hwang, I.; Chang, D.; Park, K.Y.; Kim, S.J.; Seong, W.M.; Eum, D.; Park, J.; Kim, B.; Kim, J.; et al. Nanoscale phenomena in lithium-ion batteries. Chem. Rev. 2020, 120, 6684–6737. [Google Scholar] [CrossRef]
  38. Li, Y.; Qian, J.; Zhang, M.; Wang, S.; Wang, Z.; Li, M.; Bai, Y.; An, Q.; Xu, H.; Wu, F.; et al. Co-construction of sulfur vacancies and heterojunctions in tungsten disulfide to induce fast electronic/ionic diffusion kinetics for sodium-ion batteries. Adv. Mater. 2020, 32, e2005802. [Google Scholar] [CrossRef]
  39. Li, B.; Su, Q.; Yu, L.; Zhang, J.; Du, G.; Wang, D.; Han, D.; Zhang, M.; Ding, S.; Xu, B. Tuning the band structure of MoS2 via Co9S8@MoS2 core-shell structure to boost catalytic activity for lithium-sulfur batteries. ACS Nano 2020, 14, 17285–17294. [Google Scholar] [CrossRef]
  40. Zhou, T.; Lv, W.; Li, J.; Zhou, G.; Zhao, Y.; Fan, S.; Liu, B.; Li, B.; Kang, F.; Yang, Q.H. Twinborn TiO2-TiN heterostructures enabling smooth trapping-diffusion-conversion of polysulfides towards ultralong life lithium-sulfur batteries. Energy Environ. Sci. 2017, 10, 1694–1703. [Google Scholar] [CrossRef]
  41. Ren, Y.; Zhai, Q.; Wang, B.; Hu, L.; Ma, Y.; Dai, Y.; Tang, S.; Meng, X. Synergistic adsorption-electrocatalysis of 2D/2D heterostructure toward high performance Li-S batteries. Chem. Eng. J. 2022, 439, 135535. [Google Scholar] [CrossRef]
  42. Yao, W.; Zheng, W.; Xu, J.; Tian, C.; Han, K.; Sun, W.; Xiao, S. ZnS-SnS@NC heterostructure as robust lithiophilicity and sulfiphilicity mediator toward high-rate and long-life lithium-sulfur batteries. ACS Nano 2021, 15, 7114–7130. [Google Scholar] [CrossRef] [PubMed]
  43. Lei, D.; Shang, W.; Zhang, X.; Li, Y.; Qiao, S.; Zhong, Y.; Deng, X.; Shi, X.; Zhang, Q.; Hao, C.; et al. Facile synthesis of heterostructured MoS2-MoO3 nanosheets with active electrocatalytic sites for high-performance lithium-sulfur batteries. ACS Nano 2021, 15, 20478–20488. [Google Scholar] [CrossRef] [PubMed]
  44. Jamnik, J.; Maier, J. Nanocrystallinity effects in lithium battery materials Aspects of nano-ionics. Phys. Chem. Chem. Phys. 2003, 5, 5215–5220. [Google Scholar] [CrossRef]
  45. Xu, Z.L.; Lin, S.; Onofrio, N.; Zhou, L.; Shi, F.; Lu, W.; Kang, K.; Zhang, Q.; Lau, S.P. Exceptional catalytic effects of black phosphorus quantum dots in shuttling-free lithium sulfur batteries. Nat. Commun. 2018, 9, 4164. [Google Scholar] [CrossRef]
  46. Li, Z.; Ma, Z.; Wang, Y.; Chen, R.; Wu, Z.; Wang, S. LDHs derived nanoparticle-stacked metal nitride as interlayer for long-life lithium sulfur batteries. Sci. Bull. 2018, 63, 169–175. [Google Scholar] [CrossRef]
  47. ABhargav, J.R.; He, A.; Gupta, A. Manthiram, Lithium-sulfur batteries: Attaining the critical metrics. Joule 2020, 4, 285–291. [Google Scholar] [CrossRef]
  48. Lim, W.G.; Kim, S.; Jo, C.; Lee, J. A comprehensive review of materials with catalytic effects in Li-S batteries: Enhanced redox kinetics. Angew. Chem. Int. Ed. 2019, 58, 18746–18757. [Google Scholar] [CrossRef]
  49. Fan, S.; Huang, S.; Pam, M.E.; Chen, S.; Wu, Q.; Hu, J.; Wang, Y.; Ang, L.K.; Yan, C.; Shi, Y.; et al. Design multifunctional catalytic interface: Toward regulation of polysulfide and Li. Small 2019, 15, e1906132. [Google Scholar] [CrossRef]
  50. Tan, L.; Zhou, S.; Jin, Y.; Zhu, H.; Zhang, Q.; Guo, J.; Li, X.; Dong, Z.; Yang, N.; Cong, Y. Spontaneous electrochemical reconstruction of NiNb2O6@C for high-rate lithium-ion batteries. Chem. Eng. J. 2024, 495, 153397. [Google Scholar] [CrossRef]
  51. Zhao, S.; Lian, J.; Zhang, S.; Cui, Y.; Li, G.; Wang, Y.; Li, H. Molten salt synthesis of submicron NiNb2O6 anode material with ultra-high rate performance for lithium-ion batteries. Chem. Eng. J. 2023, 461, 141997. [Google Scholar] [CrossRef]
  52. Chen, H.; Cheng, H.; Liu, H.; Hu, Y.; Yuan, T.; Dai, S.; Liu, M.; Hu, H. Design of phase interface and defect in niobium-nickel oxide for ultrafast Li-ion storage. J. Mater. Sci. Technol. 2023, 147, 145–152. [Google Scholar] [CrossRef]
  53. Bharti, B.; Kumar, S.; Lee, H.N.; Kumar, R. Formation of oxygen vacancies and Ti3+ state in TiO2 thin film and enhanced optical properties by air plasma treatment. Sci. Rep. 2016, 6, 32355. [Google Scholar] [CrossRef]
  54. He, M.; Li, X.; Li, W.; Zheng, M.; Wang, J.; Ma, S.; Yin, G.; Zuo, P.; Sun, X. Immobilization and kinetic promotion of polysulfides by molybdenum carbide in lithium-sulfur batteries. Chem. Eng. J. 2021, 411, 128563. [Google Scholar] [CrossRef]
  55. Luo, Z.H.; Zheng, M.; Zhou, M.X.; Sheng, X.T.; Chen, X.L.; Shao, J.J.; Wang, T.S.; Zhou, G. 2D Nanochannel interlayer realizing high-performance lithium-sulfur batteries. Adv. Mater. 2025, 37, 2417321. [Google Scholar] [CrossRef] [PubMed]
  56. Li, W.; Gong, Z.; Yan, X.; Wang, D.; Liu, J.; Guo, X.; Zhang, Z.; Li, G. In situ engineered ZnS-FeS heterostructure in N-doped carbon nanocages accelerating polysulfide redox kinetics for lithium sulfur batteries. J. Mater. Chem. A 2020, 8, 433–442. [Google Scholar] [CrossRef]
  57. Chung, S.H.; Manthiram, A. A polyethylene glycol-supported microporous carbon coating as a polysulfide trap for utilizing pure sulfur cathodes in lithium-sulfur batteries. Adv. Mater. 2014, 26, 7352–7357. [Google Scholar] [CrossRef]
  58. Fan, F.Y.; Carter, W.C.; Chiang, Y.M. Mechanism and kinetics of Li2S precipitation in lithium-sulfur batteries. Adv. Mater. 2015, 27, 5203–5209. [Google Scholar] [CrossRef]
  59. Kong, Z.; Lin, Y.; Hu, J.; Wang, Y.; Zhan, L. Phosphorus doped hierarchical porous carbon nanosheet array as an electrocatalyst to enhance polysulfides anchoring and conversion. Chem. Eng. J. 2022, 436, 132719. [Google Scholar] [CrossRef]
  60. Huang, S.; Wang, Z.; Von Lim, Y.; Wang, Y.; Li, Y.; Zhang, D.; Yang, H.Y. Recent advances in heterostructure engineering for lithium-sulfur batteries. Adv. Energy Mater. 2021, 11, 2003689. [Google Scholar] [CrossRef]
  61. Li, R.; Zhou, X.; Shen, H.; Yang, M.; Li, C. Conductive holey MoO2-Mo3N2 heterojunctions as job-synergistic cathode host with low surface area for high-loading Li-S batteries. ACS Nano 2019, 13, 10049–10061. [Google Scholar] [CrossRef] [PubMed]
  62. Liang, W.; Zhou, X.; Zhang, B.; Zhao, Z.; Song, X.; Chen, K.; Wang, L.; Ma, Z.; Liu, J. The versatile establishment of charge storage in polymer solid electrolyte with enhanced charge transfer for LiF-rich SEI generation in lithium metal batteries. Angew. Chem. 2024, 136, e202320149. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the preparation of Ni3N/Nb4N5 heterostructure.
Figure 1. Schematic diagram of the preparation of Ni3N/Nb4N5 heterostructure.
Nanomaterials 15 01015 g001
Figure 2. (a) XRD patterns of NiNb2O6 nanorods and (b) Ni3N/Nb4N5 heterostructure.
Figure 2. (a) XRD patterns of NiNb2O6 nanorods and (b) Ni3N/Nb4N5 heterostructure.
Nanomaterials 15 01015 g002
Figure 3. SEM image of (a) NiNb2O6 and (b) Ni3N/Nb4N5 heterostructure; (c) cross-sectional SEM of Ni3N/Nb4N5-PP; (d) low-magnification TEM image of Ni3N/Nb4N; (eh) STEM image and EDS elemental mappings; (i) high-resolution TEM image of Ni3N/Nb4N.
Figure 3. SEM image of (a) NiNb2O6 and (b) Ni3N/Nb4N5 heterostructure; (c) cross-sectional SEM of Ni3N/Nb4N5-PP; (d) low-magnification TEM image of Ni3N/Nb4N; (eh) STEM image and EDS elemental mappings; (i) high-resolution TEM image of Ni3N/Nb4N.
Nanomaterials 15 01015 g003
Figure 4. (ac) Visual adsorption experiment of Li2S6 by pristine Ni3N/Nb4N5 and NiNb2O6.
Figure 4. (ac) Visual adsorption experiment of Li2S6 by pristine Ni3N/Nb4N5 and NiNb2O6.
Nanomaterials 15 01015 g004
Figure 5. High-resolution XPS spectra of (a) N 1s, (b) Nb 3d, and (c) Ni 2p before Ni3N/Nb4N5 adsorption of Li2S6, and (d) N 1s, (e) Nb 3d, and (f) Ni 2p after Ni3N/Nb4N5 adsorption of Li2S6.
Figure 5. High-resolution XPS spectra of (a) N 1s, (b) Nb 3d, and (c) Ni 2p before Ni3N/Nb4N5 adsorption of Li2S6, and (d) N 1s, (e) Nb 3d, and (f) Ni 2p after Ni3N/Nb4N5 adsorption of Li2S6.
Nanomaterials 15 01015 g005
Figure 6. High-resolution XPS spectra of (a) Nb 3d, (b) Ni 2p, and (c) O 1s before NiNb2O6 adsorption of Li2S6, and high-resolution XPS spectra of (d) Nb 3d, (e) Ni 2p, and (f) O 1s after NiNb2O6 adsorption of Li2S6.
Figure 6. High-resolution XPS spectra of (a) Nb 3d, (b) Ni 2p, and (c) O 1s before NiNb2O6 adsorption of Li2S6, and high-resolution XPS spectra of (d) Nb 3d, (e) Ni 2p, and (f) O 1s after NiNb2O6 adsorption of Li2S6.
Nanomaterials 15 01015 g006
Figure 7. High-resolution XPS spectra of S 2p after adsorption of Li2S6 by (a) Ni3N/Nb4N5 heterojunction and (b) NiNb2O6.
Figure 7. High-resolution XPS spectra of S 2p after adsorption of Li2S6 by (a) Ni3N/Nb4N5 heterojunction and (b) NiNb2O6.
Nanomaterials 15 01015 g007
Figure 8. CV curves of (a) symmetric cells with Ni3N/Nb4N5 and NiNb2O6; (b) the initial CV curve and (c) EIS curve of Li-S cells with Ni3N/Nb4N5-PP, NiNb2O6-PP, and PP separators.
Figure 8. CV curves of (a) symmetric cells with Ni3N/Nb4N5 and NiNb2O6; (b) the initial CV curve and (c) EIS curve of Li-S cells with Ni3N/Nb4N5-PP, NiNb2O6-PP, and PP separators.
Nanomaterials 15 01015 g008
Figure 9. Potentiostatic discharge curves of (a) Ni3N/Nb4N5 and (b) NiNb2O6 electrode.
Figure 9. Potentiostatic discharge curves of (a) Ni3N/Nb4N5 and (b) NiNb2O6 electrode.
Nanomaterials 15 01015 g009
Figure 10. Charge–discharge curves of the first three cycles of Li-S cells with (a) Ni3N/Nb4N5-PP, (b) NiNb2O6-PP, and (c) PP at 0.1 C; (d) cyclic performance of Li-S cells with Ni3N/Nb4N5-PP, NiNb2O6-PP, and PP at 0.2 C; (e) rate performance.
Figure 10. Charge–discharge curves of the first three cycles of Li-S cells with (a) Ni3N/Nb4N5-PP, (b) NiNb2O6-PP, and (c) PP at 0.1 C; (d) cyclic performance of Li-S cells with Ni3N/Nb4N5-PP, NiNb2O6-PP, and PP at 0.2 C; (e) rate performance.
Nanomaterials 15 01015 g010
Table 1. Some recent heterojunction applications for Li-S battery interlayers.
Table 1. Some recent heterojunction applications for Li-S battery interlayers.
TypologyCurrent DensityNumber of CyclesCapacity Decay RateCapacity/Retention Rate
1Co9S8@MoS2/CNFinterlayer1.0 C4000.09%[39]
2TiO2–TiNinterlayer0.3 C300/927 mAh g−1 [40]
3Ti3C2Tx/Ni-Co MOFmodified separators0.2 C2000.06%1100 mAh g−1/87.3% [41]
4ZnS-SnS@NCmodified separators4.0 C20000.01%632 mAh g−1/74.9% [42]
5MoS2-MoO3modified separators1.0 C6000.01%92.00% [43]
6This workmodified separators0.2 C2000.25%796.2 mAh g−1
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

Li, X.; Shang, W.; Zhang, S.; Xu, C.; Lian, J.; Li, G. The Design of the Ni3N/Nb4N5 Heterostructure as Bifunctional Adsorption/Electrocatalytic Materials for Lithium–Sulfur Batteries. Nanomaterials 2025, 15, 1015. https://doi.org/10.3390/nano15131015

AMA Style

Li X, Shang W, Zhang S, Xu C, Lian J, Li G. The Design of the Ni3N/Nb4N5 Heterostructure as Bifunctional Adsorption/Electrocatalytic Materials for Lithium–Sulfur Batteries. Nanomaterials. 2025; 15(13):1015. https://doi.org/10.3390/nano15131015

Chicago/Turabian Style

Li, Xialei, Wen Shang, Shan Zhang, Chun Xu, Jiabiao Lian, and Guochun Li. 2025. "The Design of the Ni3N/Nb4N5 Heterostructure as Bifunctional Adsorption/Electrocatalytic Materials for Lithium–Sulfur Batteries" Nanomaterials 15, no. 13: 1015. https://doi.org/10.3390/nano15131015

APA Style

Li, X., Shang, W., Zhang, S., Xu, C., Lian, J., & Li, G. (2025). The Design of the Ni3N/Nb4N5 Heterostructure as Bifunctional Adsorption/Electrocatalytic Materials for Lithium–Sulfur Batteries. Nanomaterials, 15(13), 1015. https://doi.org/10.3390/nano15131015

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