Next Article in Journal
In Situ Engineered Plastic–Crystal Interlayers Enable Li-Rich Cathodes in PVDF-HFP-Based All-Solid-State Polymer Batteries
Previous Article in Journal
Experience-Driven NeuroSymbolic System for Efficient Robotic Bolt Disassembly
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Transition Metal-Based Catalysts Powering Practical Room-Temperature Na-S Batteries: From Advances to Further Perspectives

1
State Key Laboratory of Advanced Glass Materials, Wuhan 430070, China
2
Wuhan Belt and Road Joint Laboratory on Near-Zero Carbon Materials, School of Chemistry, Chemical Engineering and Life Sciences, Wuhan University of Technology, Wuhan 430070, China
3
School of Safety Science and Emergency Management, Wuhan University of Technology, Wuhan 430070, China
4
State Key Laboratory of Materials Processing and Die & Mould Technology, School of Materials Science and Engineering, Huazhong University of Science and Technology, Wuhan 430074, China
*
Author to whom correspondence should be addressed.
Batteries 2025, 11(9), 333; https://doi.org/10.3390/batteries11090333
Submission received: 18 July 2025 / Revised: 23 August 2025 / Accepted: 1 September 2025 / Published: 5 September 2025
(This article belongs to the Special Issue 10th Anniversary of Batteries: Interface Science in Batteries)

Abstract

Room-temperature sodium–sulfur (RT Na-S) batteries hold great potential in the field of large-scale energy storage due to their high theoretical energy density and low cost of raw materials. However, the inherent low conductivity, notorious shuttling, and sluggish kinetics of cathode materials cause the loss of active substances and capacity delay, hindering the practical application of RT Na-S batteries. Owing to their low cost, variable oxidation states, and unsaturated d orbitals, transition metal (TM)-based catalysts have been extensively studied in circumventing the above shortcomings. Herein, the review first elaborates on the reaction mechanisms and current challenges of RT Na-S batteries. Subsequently, the role and function mechanism of TM-based catalysts (including single/dual atoms, nanoparticles, compounds, and heterostructures) in RT Na-S batteries are described. Specifically, based on the theories of electronic transfer and atomic orbital hybridization, the interaction mechanism between TM-based catalysts and polysulfides, as well as the catalytic performance, are systematically discussed and summarized. Finally, a discussion on the challenges and future research perspectives associated with TM-based catalysts for RT Na-S batteries is provided.

Graphical Abstract

1. Introduction

With the rapid advancement of the global energy transition, the demand for renewable energy sources (wind, geothermal, and solar) is steadily increasing. However, constrained by their geographical availability, intermittency, and instability, the development of energy storage technologies is crucial for storing these unstable energy sources [1,2]. Battery energy storage systems have emerged as critical grid-stabilizing solutions due to their rapid power regulation capabilities. Among them, lithium-ion batteries have achieved scalable deployment by leveraging fundamental advantages, including high energy density, long-term cycling stability, and a low capacity fade rate [3]. Furthermore, the implementation of electrospinning technology has substantially accelerated the industrial application of lithium-ion batteries through precise electrode microstructure engineering [4]. Although lithium-ion batteries have dominated the energy storage market for the past three decades, their poor energy density makes it more difficult for them to meet societal demands. Owing to their enormous capacity and low anode electrochemical potential, lithium metal batteries have recently attracted considerable interest from both academia and industry. Nevertheless, the scarcity and uneven distribution of lithium resources result in high battery costs, hindering the development and application of lithium metal batteries. Sodium, which is a member of the same periodic table group as lithium, is abundant and inexpensive, making sodium metal anodes a highly promising alternative to lithium metal anodes [5,6]. Coupling an abundant sulfur cathode, which exhibits a high theoretical specific capacity (1675 mAh/g) and low cost, with sodium constitutes sodium–sulfur (Na-S) batteries. These systems achieve a theoretical energy density of up to 1274 Wh/kg and are recognized as a highly promising next-generation technology for large-scale energy storage applications. High-temperature Na-S batteries, operating at 300–350 °C, were deployed in energy storage systems as early as the last century [7]. Nevertheless, sustaining such elevated temperatures necessitates substantial auxiliary energy input, imposing a significant economic burden. Moreover, under these conditions, the molten electrodes (anode and cathode) are partitioned solely by a solid-state β″-alumina electrolyte; fracture of this ceramic separator may permit direct electrode contact, presenting critical safety hazards including combustion or explosion risks [8,9]. Intermediate-temperature Na-S batteries (130–180 °C) similarly utilize β″-alumina solid-state electrolytes. The reduced operational temperature, however, markedly diminishes the ionic conductivity of sodium ions (Na+), thereby compromising overall electrochemical performance [10,11]. Given the dual constraints of cost and safety, the vigorous development of room-temperature sodium–sulfur (RT Na-S) batteries represents a critical imperative. Under ambient conditions, sulfur undergoes complete conversion to sodium sulfide (Na2S), enabling the theoretical capacity to be achieved [12]. These inherent advantages highlight the necessity and significant promise of RT Na-S battery systems.
Since the utilization of polyvinylidene fluoride (PVDF)-based gel polymer electrolytes in 2006, research on RT Na-S batteries has advanced significantly [13]. During battery operation, the sulfur cathode undergoes a stepwise conversion process. The initially formed sodium polysulfide (NaPS) intermediates with varying chain lengths are ultimately reduced to sodium sulfide (Na2S). Theoretically, the complete reduction of elemental sulfur to Na2S can deliver a high theoretical specific capacity of 1675 mAh/g. However, this reaction system faces multiple challenges: First, long-chain NaPSs readily dissolve in ether-based electrolytes [14] or undergo nucleophilic side reactions in ester-based electrolytes [15], resulting in active material loss. Second, short-chain NaPSs (such as sodium persulfide, Na2S2) and the final product Na2S exhibit extremely low solubility in electrolytes, typically depositing as poorly conductive solids. This significantly reduces reaction conversion kinetics, leading to battery performance degradation. Furthermore, elemental sulfur suffers from intrinsically poor conductivity, causing inefficient electron transfer between sulfur species during charge/discharge cycles, which severely limits sulfur utilization. These issues become particularly pronounced when the sulfur loading in the cathode exceeds 3.0 mg/cm2, making stable battery operation challenging [16].
To address the aforementioned challenges and enhance sulfur loading in the cathode as well as energy density, numerous studies have employed conductive porous materials as hosts or separator modifiers [17,18]. These materials serve a dual purpose, enhancing the intrinsic electronic conductivity of the cathode (as sulfur hosts) or acting as secondary current collectors (as separator coatings); concurrently, they effectively suppress the uncontrolled diffusion and shuttle effect of NaPSs through physical confinement or chemical adsorption. For instance, utilizing porous carbon materials (e.g., graphene [19,20], carbon nanotubes [21], carbon nanofibers [22,23,24,25,26]) and porous organic materials (e.g., metal–organic frameworks [27], covalent organic frameworks [28,29], polyacrylonitrile [30]) as hosts or separator modifiers has improved the performance of RT Na-S batteries. However, porous materials can only partially mitigate, not fundamentally resolve, the issues of the NaPS shuttle effect and sluggish conversion kinetics between sulfides of different chain lengths. Drawing inspiration from the development path of lithium–sulfur batteries, researchers have gradually shifted their focus to catalysts for RT Na-S batteries, aiming to accelerate reaction conversion rates and reduce the residence time of NaPS intermediates in the electrolyte [31,32,33,34]. Compared with other catalysts, e.g., precious metal-based catalysts, transition metal (TM)-based catalysts stand out due to their abundance and low cost. In addition, TM-based catalysts exhibit variable oxidation states and partially filled d orbitals, enabling their exceptional performance in catalysis, materials science, energy storage, and related fields.
The classification and application of TM-based catalysts for RT Na-S batteries have been extensively reviewed [35,36,37]. For instance, synthesis strategies for TM-based catalysts were systematically categorized according to metal element species [35]. However, the fundamental understanding of their intrinsic nature (particularly the electronic structures and orbital interactions governing catalytic behavior) remains insufficiently deep and systematic. Herein, this review systematically elucidates the interaction mechanisms between various TM-based catalysts (e.g., metal single/dual atoms [38], metal nanoparticles, metal compounds, heterostructures [39]) and NaPSs, as illustrated in Figure 1. It conceptualizes the adsorption–conversion principle, major catalyst categories, and key performance-determining factors. Furthermore, by establishing structure–performance relationships between the micro-scale electronic/orbital characteristics and macro-scale electrochemical behavior, it provides a theoretical foundation for clarifying the electron transfer mechanisms as well as the “adsorption–catalysis” reaction pathway. Finally, further directions for the rational design of next-generation, high-efficiency, stable TM-based catalysts are highlighted, which are expected to promote the practical implementation of RT Na-S batteries.

2. Working Mechanisms, Challenges, and Design Principles of RT Na-S Batteries

2.1. Working Mechanisms

The redox reaction of the sulfur cathode involves multi-electron transfer processes, inevitably generating NaPS intermediates. Notably, NaPSs exhibit significant differences in dissolution behavior and chemical stability between ester-based and ether-based electrolytes [41]. Consequently, when these two types of electrolytes are applied to RT Na-S batteries, the electrochemical reaction mechanisms of the sulfur cathode also demonstrate marked distinctions.
In ether-based electrolytes, the solubility of NaPSs decreases dramatically with chain length shortening. During the charge/discharge of RT Na-S batteries, sulfur undergoes a multi-step “solid–liquid–solid” conversion pathway [42]. As illustrated in Figure 2a, in the initial discharge stage, S8 combines with Na+ to form soluble Na2S8 (solid-to-liquid conversion), attributed to the voltage plateau at 2.30 V (Process (1) in Figure 2b). Subsequently, Na2S8 is further reduced to Na2S6 and Na2S4 (liquid-to-liquid conversion), with the voltage decreasing along a slope between 2.30 V and 1.60 V (Process (2) in Figure 2b). This stage involves multiple equilibrium conversion steps (Processes (3)–(5) in Figure 2b). Next, Na2S4 continues to be reduced to Na2S3, Na2S2, or Na2S (liquid-to-solid conversion), forming a low-voltage plateau at 1.60 V (Processes (6)–(8) in Figure 2b). Finally, Na2S2 is reduced to the end product Na2S (solid-to-solid conversion, Process (9) in Figure 2b). However, the poor conductivity of the discharge products (Na2S2 and Na2S) results in relatively sluggish reduction kinetics, often leading to incomplete conversion. Synchronously, this causes a phenomenon during charging where the voltage rises rapidly while the capacity shows a negligible increase, namely polarization.
In ester-based electrolytes, during the initial cycles, S8 undergoes Processes (1) and (3) shown in Figure 2b, generating long-chain NaPSs (Na2Sx, 4 < x ≤ 8). However, long-chain NaPSs exhibit significant thermodynamic instability in ester solvents. Through density functional theory (DFT) calculation, the potential energy difference (ΔE) of the reaction between EC and Na2S6 was found to be negative. This indicates that the reaction is thermodynamically favorable and tends to proceed spontaneously because the energy of the reactants is greater than that of the products. It was discovered that EC and long-chain NaPSs undergo irreversible adverse reactions [45]. The disappearance of the high-voltage plateau (approximately 2.30 V) during subsequent discharge (Figure 2c) confirms that long-chain Na2Sx cannot participate in further reversible conversion [47]. Moreover, within the voltage region below 2.0 V, the reaction corresponds to the stepwise conversion of long-chain NaPSs (which do not directly contact the electrolyte) into short-chain NaPSs (Na2Sx, 2 ≤ x ≤ 4), ultimately yielding Na2S. Notably, the ΔE between DEGDME and Na2S6 consistently exceeds 1 eV (Figure 2d), indicating negligible interaction between the ether solvent and NaPSs. Consequently, long-chain NaPSs exhibit higher solubility in ether-based electrolytes. By contrast, the significantly reduced solubility of NaPSs in ester-based electrolytes leads to sluggish sulfur electrode reaction kinetics and the concurrent occurrence of multiple reaction steps, ultimately manifesting as weakened plateau features in the charge–discharge curves.
Ester-based electrolytes (e.g., EC, PC, DEC, DMC) exhibit high dielectric constants and wide electrochemical stability windows, enabling stable operation at higher voltages. Additionally, their low solubility for NaPSs effectively suppresses NaPS dissolution and the shuttle effect, thereby extending cycle life [48]. By contrast, ether-based electrolytes (e.g., TEGDME, glymes, DEGDME) display high NaPS solubility, which facilitates the “solid–liquid–solid” conversion reaction, endowing it with superior rate capability.
Consequently, the reduction of sulfur cathodes generates long-chain NaPSs in both ether-based and ester-based electrolytes. For an ether-based system, suppressing NaPS shuttling requires cathode host design or separator modification. For ester-based systems, constructing stable interfaces is essential to prevent direct contact between NaPSs and the electrolyte, thereby avoiding irreversible side reactions. Notably, both electrolytes necessitate catalytic materials to accelerate the conversion kinetics of short-chain NaPSs (Na2S2/Na2S) to achieve high specific capacity and long-term cycling stability.

2.2. Challenges and Design Principles

During practical cycling of RT Na-S batteries, the sulfur cathode faces multiple challenges (Figure 2e) that severely constrain its capacity utilization and cycling stability. Firstly, both the initial active material sulfur (S8) and the discharged products (Na2S and Na2S2) exhibit extremely low electronic and ionic conductivity, leading to impeded electron/ion transport and sluggish electrochemical reaction kinetics [49]. This not only reduces active material utilization but also causes significant polarization. Secondly, during the solid–liquid–solid phase transformation, sluggish kinetics prolong the residence time of NaPS intermediates in the electrolyte, enabling their diffusion to the anode side via concentration gradients. This results in irreversible active material loss, capacity decay, and reduced Coulombic efficiency. In addition, the conversion of sulfur to Na2S involves substantial volume expansion, causing the cathode to endure repeated expansion–contraction stress cycles, ultimately leading to electrode structure pulverization [50]. Concurrently, continuous volume changes compromise the integrity of the cathode–electrolyte interphase (CEI) layer, exacerbating electrolyte infiltration and triggering NaPS dissolution and side reactions, which eventually cause battery failure.
Increasing the sulfur loading is crucial for enhancing the energy density of RT Na-S batteries, but the thick electrode structure introduces more severe challenges. Under high sulfur loading, increased sulfur particle stacking density disrupts the conductive network, hindering electron transport pathways. Synchronously, thick electrodes significantly extend Na+ diffusion paths, intensifying concentration polarization and reducing active material utilization. Additionally, sluggish solid–liquid–solid phase transformation kinetics prolong NaPS residence time, and under high-loading conditions, the shuttle effect intensifies, leading to rapid battery failure.
To address challenges such as the shuttle effect and volume expansion, researchers have pursued multi-dimensional material and structural optimizations. On the one hand, constructing three-dimensional porous conductive scaffolds or introducing polar host materials effectively enhances sulfur loading and suppresses NaPS dissolution and shuttling while mitigating volume changes during cycling [51]. On the other hand, introducing interlayers between the cathode and separator or modifying the separator surface creates physical/chemical barriers that effectively adsorb or block NaPS migration, significantly suppressing the shuttle effect [52]. These approaches have improved sulfur utilization and reduced active material loss to some extent, thereby enhancing electrochemical performance.
However, although the aforementioned strategies based on physical confinement and chemical adsorption can partially alleviate NaPS diffusion and shuttling, they are fundamentally passive restriction mechanisms. They struggle to achieve long-term, effective immobilization of high-concentration, soluble NaPS intermediates under high sulfur loading and fail to fundamentally resolve the core issue of slow NaPS conversion. By contrast, introducing highly efficient catalytic materials (e.g., metals and their compounds) demonstrates significant advantages [53]. These catalysts can not only anchor NaPSs via chemical adsorption but also lower the reaction energy barrier for converting NaPSs into the final discharge products [54], fundamentally accelerating the kinetics for NaPS conversions, improving sulfur utilization, and consequently enhancing battery cycling stability and capacity (Figure 2f). The classification and mechanistic roles of catalysts will be comprehensively reviewed and discussed in the next section.

3. Classifications and Function Mechanisms of Catalytic Materials

Currently, the catalysts applied in RT Na-S batteries are predominantly transition metal-based materials. Leveraging their advantages, including tunable electronic structures (d-orbital characteristics), multivalent flexibility, and strong interfacial interactions, these catalysts can be categorized into four main types: metal single/dual atoms, metal nanoparticles, metal compounds, and heterostructure materials. As summarized in Table 1, these materials effectively lower the energy barriers of sulfur redox reactions and accelerate reaction kinetics through diverse function mechanisms, such as electron transfer driven by strong Lewis acid–base interactions, formation of metal–sulfur chemical bonds, and regulation of reaction pathways. As directly demonstrated in Figure 3, single/dual atoms exhibit notable cycling stability at low current densities, while heterostructure catalysts display superior long-term cycling stability under high-current-density conditions.

3.1. TM Single/Dual-Atom Catalysts

TM single/dual-atom catalysts, leveraging their unique atomically dispersed characteristics and tunable electronic structures, offer novel solutions to address NaPS shuttling and sluggish redox kinetics. However, their inherent high surface energy endows them with agglomeration, necessitating the introduction of electron-rich heteroatoms (e.g., N, P, O, S) to construct Lewis base sites. These sites promote the formation of stable M-N coordination bonds with Lewis acidic metal centers (M), enabling robust anchoring within carbon matrices [111,112]. Such coordination not only significantly enhances metal atom stability but, more critically, allows precise modulation of the electronic structure of central metals by tailoring the coordination microenvironment of M-Nx sites (including ligand types, coordination number, and spatial configuration). This optimization ultimately governs their interaction mechanisms with NaPSs.

3.1.1. TM Single-Atom Catalysts

TM single-atom catalysts (SACs) establish strong interactions with NaPSs through isolated metal sites (e.g., Mn, Fe, Co). This interaction not only physically anchors NaPSs to suppress their dissolution in electrolytes but also reduces reaction energy barriers. More critically, electron transfer from the TM single-atom center to sulfur species alters the electronic state of sulfur. This charge transfer process fundamentally arises from electron redistribution between atoms/molecules, originating directly from orbital hybridization that modifies electron cloud density and energy states—a quantum process where energetically proximate valence atomic orbitals (e.g., s, p, d) reconfigure into degenerate hybrid orbitals with new spatial orientations. This orbital reorganization occurs during molecular formation, driven by interatomic interactions that mix disparate orbital types into isoenergetic hybrid sets. Hybrid orbitals affect the adsorption of polysulfides by altering the electron occupation state of the eg orbitals. Meanwhile, they lower the dissociation energy barrier of the S-S bond and enhance the M-S chemical bond, jointly reducing the system energy and accelerating the kinetics of sulfur redox. The degree of orbital hybridization dictates charge transfer intensity: stronger hybridization enhances orbital overlap and energy-level matching, leading to more pronounced charge transfer, and vice versa. As a key regulatory mechanism, the coordination environment directly influences hybridization patterns by modulating the electronic structure and orbital energy levels of central atoms, thereby enabling precise control over charge transfer intensity. Specifically, coordination environment modulation can be achieved through the following four mechanisms.
Regulation of central metal atoms: In RT Na-S batteries, SACs potentiate the adsorption–catalysis synergy of sulfur species by modulating orbital hybridization between metal centers and NaPSs. The electronic structure of metal sites (e.g., d-band center position, occupied state density) directly governs interfacial electron interactions with NaPSs. Tian et al. [63] systematically demonstrated that Mn-N4 sites, with abundant unoccupied antibonding orbitals, significantly enhance d-p orbital hybridization, thereby improving electrochemical performance. Lei et al. [73] further revealed that Mn single-atom sites exhibit d-band states closer to the Fermi level (indicating fewer electrons in antibonding orbitals) and the lowest d-band center, which boosts electron supply capacity to cleave S-S bonds and accelerate short-chain NaPS/Na2S formation. Their strong electron-trapping capability also facilitates reversible Na2S conversion, ultimately enhancing cycling stability. To efficiently screen ideal SACs for high-energy-density Na-S batteries, Bai et al. [74] innovatively integrated natural language processing (NLP, Figure 4a) with catalytic descriptors (the Gibbs free energy of capacity-determining step and d-band center of absorption-determining ability, Figure 4b). This strategy employs NLP for precise central-atom screening, constructs binary descriptors to optimize candidate materials, and ultimately identifies cobalt SACs anchored by nitrogen/sulfur (SA Co-N/S). This approach pioneers new pathways for catalyst selection and provides novel insights into interfacial catalytic kinetics in Na-S systems.
Regulation of the coordination number: In single-atom catalyst studies, metal centers are typically anchored to carbon substrates via coordination with four nitrogen atoms (M-N4). While this symmetric coordination ensures uniform electron distribution, it constrains catalytic activity. Precise modulation of the coordination number emerges as an effective strategy to overcome this limitation: reduced coordination induces asymmetric structures that enhance charge localization and adsorption strength. Yao et al. [40] demonstrated that Zn-N2 sites accelerate sulfur conversion and lower Na2S decomposition barriers, enabling the Zn-N2/CF/S cathode to achieve a minimal capacity decay rate of 0.006% per cycle after 4000 cycles at 10 A g−1. Song et al. [64] established a dual-descriptor model: the geometric descriptor γ (lNa-S/lFe-N) negatively correlates with adsorption strength, while the electronic descriptor φ (eg/t2g) positively correlates with Na2S decomposition barriers. The Fe-N1 configuration with minimal γ (0.824) and maximal φ (0.664) exhibits optimal catalytic activity. Li et al. [70] confirmed Mn-S bond formation in Mn-N2 via Mn 2p XPS (Figure 4c), where enhanced d-p hybridization improves S8 chemisorption energy, delivering a faster response (114 s) and higher peak current (0.104 mA) of Na2S precipitation (Figure 4d).
Notably, p-block metals are conventionally considered catalytically inert due to filled d orbitals, but combining with electron-rich atoms can induce activity [113]. Guo et al. [114] engineered In-N3 sites (εd = −5.256 eV) with elevated d-band centers versus In-N4d = −5.427 eV), raising p-d hybridized antibonding orbital levels and reducing bonding energy, thereby lowering NaPS reaction barriers. When steric hindrance impedes metal–sulfur hybridization, alkali metal cations’ (Li+/Na+) s orbitals interact with catalyst d/p orbitals, altering reaction pathways [115]. Wu et al. [67] introduced a fifth axial N atom to planar InN4, forming InN5 configurations that strengthen s-orbital overlap with Na, significantly enhancing charge transfer.
Doping of Coordination Atoms: While reducing the coordination number enhances active site exposure, it may compromise metal center stability. Introducing heteroligands (e.g., O, S) provides an effective alternative for electronic structure modulation. Xiao et al. [59] coordinated Cu atoms with two N and two O atoms, forming Cu sites that weaken S-S bonds in cyclic S8 molecules, catalyzing the conversion of elemental sulfur to short-chain S2−4 within pores and facilitating mutual transformation between short-chain sulfur species and Na2S. Zheng et al. [71] synthesized asymmetrically coordinated Zn single-atom sites via template sacrifice (Figure 4e). Compared to symmetric Zn-N4 sites, Zn-N3O exhibits superior catalytic activity for NaPSs, maintaining 1155 mAh g−1 reversible capacity after 100 cycles at 0.1C. UV-vis spectra show weakened absorption peaks and clarified Na2S6 solution (Figure 4f), confirming strong NaPS adsorption attributed to N/O ligands, inducing localized charges at Zn centers. This triggers intense Zn-S d-p orbital hybridization, forming bonding orbitals below the Fermi level (EF) and antibonding orbitals above EF (Figure 4g). Such electronic restructuring elongates Na-S bonds in adsorbed Na2Sx by 0.18 Å, reducing reaction barriers during charge/discharge and accelerating sulfur conversion kinetics. Similarly, Hu et al. [62] derived N/O dual-coordinated Co single-atom catalysts from MOF precursors. Oxygen incorporation modulates Co d-orbital electron density, strengthening Co-Na2Sx d-p hybridization. This significantly lowers the Gibbs free energy change (ΔG) of the rate-limiting step (Na2S4→Na2S2) and reduces Na2S decomposition barriers by 0.17 eV, demonstrating that enhanced d-p hybridization effectively accelerates sulfur redox kinetics.
Spatial configuration engineering: Spatial configuration engineering serves as a critical strategy for optimizing the electronic structure of metal single-atom catalysts. Beyond modulating the primary coordination shell, secondary coordination shell design proves equally essential. Bai et al. [72] engineered a single-atom catalyst (SA Fe-N/S@CNF) anchored on 3D interconnected carbon nanofibers, featuring an Fe-N4S2 active center where two S atoms reside in the secondary coordination shell—bonded to primary-shell N atoms without direct Fe coordination. These secondary-shell S atoms regulate the electronic structure of Fe via d-p orbital hybridization, elevating the Fe d-band center by Δα toward the Fermi level (Figure 4h). This shift raises antibonding orbital energies above the Fermi level, triggering electron back-donation to bonding orbitals and consequently enhancing NaPS binding and catalytic capability. Furthermore, introducing axial ligands overcomes planar coordination constraints to optimize metal centers. Zheng et al. [66] incorporated an axial N ligand to planar Ca-O4 sites, inducing charge localization that strengthens p-p hybridization between Ca 3p and S 3p orbitals in NaPSs. The higher adsorption energy of NaPSs on Ca-O4N-C-SAC versus Ca-O5-C-SAC originates from Ca-S bond formation, which effectively traps NaPS intermediates, suppresses shuttling, and boosts both NaPS affinity and conversion kinetics.

3.1.2. Dual-Atom Catalysts

Although single-atom catalysts demonstrate distinct advantages in enhancing NaPS adsorption, suppressing shuttle effects, and improving conversion kinetics through precise coordination environment modulation, their catalytic performance remains constrained by the inherent limitation of single active sites in simultaneously achieving strong adsorption and efficient catalytic conversion of NaPSs. To overcome this fundamental constraint, researchers have shifted focus to bimetallic dual-atom catalysts (BDACs). These diatomic sites establish bifunctional active centers via electronic coupling and geometric synergy between adjacent metal atoms, thus providing expanded atomic-level design flexibility for synergistic optimization of adsorption–catalysis performance toward NaPSs.
Li et al. [61] developed an Fe-Co heteronuclear dual-atom catalyst embedded in nitrogen-doped hollow carbon nanospheres (Fe-Co/NC). Aberration-corrected HAADF-STEM characterization (Figure 4i) confirmed the existence of Fe-Co atomic pairs. Compared to the monometallic systems, the Fe-Co/NC cathode has a lower concentration (43%), showing a lower polarization voltage and a larger redox peak current (Figure 4j), indicating a significantly enhanced sulfur conversion kinetics at the diatomic position. During the sulfur reduction process, the Fe-Co/NC cathode exhibits the strongest adsorption of NaPSs, and the change in the free energy of the rate-determining step (Na2S4→Na2S) is also the smallest (Figure 4k), further demonstrating that the conversion of NaPSs is accelerated. Moreover, Fe-Co/NC demonstrates a remarkably low Na2S decomposition energy barrier of 0.89 eV (Figure 4l), confirming superior catalytic efficiency for Na2S reduction. Mechanistic studies reveal that Fe introduction induces Co 3d-Fe 3d orbital coupling, narrowing the energy gap between Co dxy and dz2 orbitals by Δo. This promotes electron transfer to eg orbitals, shifting Co(II) from a low-spin to a high-spin state due to limited eg occupancy and reduced dz2 occupation. Consequently, enhanced hybridization between Co dz2 orbitals and antibonding π* orbitals of sulfur in NaPSs occurs, substantially boosting catalytic performance.

3.2. TM Nanoparticle Catalysts

Different from atomically dispersed single/dual-atom catalysts, metal nanoparticles (e.g., Fe [116], Co [78,117], Ni [118]) comprise aggregates of tens to thousands of atoms. These nanoparticle catalysts effectively reduce energy barriers for sulfur redox reactions, enhance NaPS interactions, and accelerate conversion kinetics. During catalytic processes in RT Na-S batteries, electron transfer between metal surfaces and adsorbed NaPSs induces localized charge polarization at metal–sulfur interfaces. This charge redistribution provides the driving force for metal–sulfur bond formation. The anchoring and catalytic capabilities of metal nanoparticles toward NaPSs are governed by two critical factors: (1) support structures dictate catalyst–NaPS contact efficiency and mass transport, and (2) intrinsic activity, determined by electronic structure, is primarily regulated by the charge state of metal centers. Strategic approaches, including doping engineering, defect engineering, and alloying, enable precise modulation of nanoparticle electronic structures, thereby optimizing metal charge states for enhanced performance.

3.2.1. Regulation of Catalyst Carriers

The structure of catalyst carriers plays dual roles in regulating NaPS conversion: physically confining NaPS diffusion through adsorption while enhancing active site utilization by optimizing catalyst dispersion. Precise support design significantly boosts catalytic performance: Du et al. [78] embedded cobalt nanoparticles in N-doped porous carbon nanofibers (Co@NPCNFs), where the structure facilitates electron transfer between Co and Na2Sx, accelerating NaPS–sodium-ion binding. This enables the Co@NPCNFs/S cathode to achieve an ultralow capacity decay rate of 0.038% per cycle over 800 cycles at 0.1C. Yang et al. [51] developed a 3D porous MG-Co aerogel (Figure 5a) integrating MXene nanosheets, reduced graphene oxide (rGO), and Co nanoparticles. Compared to MXene, it exhibits a larger pore volume (0.32 cm3 g−1) and higher specific surface area (143 m2 g−1), significantly enhancing Co utilization. The MG-Co@S cathode maintains 360 mAh g−1 after 200 cycles at 0.5C (Figure 5b). Electrochemical tests reveal that symmetric cells with MG-Co electrodes in Na2S6 electrolyte show a higher current response and larger CV-integrated area (Figure 5c), confirming that Co has a significant catalytic effect on the redox reactions of sulfur species. EIS analysis demonstrates a reduction in charge transfer resistance (Rct) for MG-Co versus MG (Figure 5d), indicating optimized interfacial kinetics. This enhancement facilitates more efficient sulfur conversion, reduces polarization, and ultimately improves the energy conversion efficiency and reaction reversibility of sodium–sulfur batteries. XPS verifies Ti-S (458.6 eV and 464.5 eV) and Co-S (785.4 eV and 802.1 eV) bond formations, strengthening chemical adsorption of NaPSs.

3.2.2. Doping Engineering

The electronic structure modulation of metal nanoparticles via doping engineering can effectively optimize their catalytic performance. Tang et al. [81] employed a boron (B) doping strategy to reconstruct the electronic structure of cobalt nanoparticles, successfully inducing the formation of electron-deficient Co sites (Figure 5e). DFT calculations revealed that the Bader charge of Co in Co@BNC reached 1.84, significantly higher than that of undoped Co@NC (1.52), theoretically confirming the formation of electron-deficient Co centers. XPS characterization further elucidated the charge transfer mechanism: Upon Na2S6 adsorption, the Co 2p3/2 binding energy of Co@BNC shifted by 0.6 eV toward lower energy, indicating electron transfer from Na2S6 to the electron-deficient Co sites. Meanwhile, the emergence of a new Co-S bond peak at 777.4 eV (with a 1.8 eV increase in bond energy) further verified the formation of strong interfacial chemical interactions (Figure 5f).

3.2.3. Defect Engineering

Defect engineering optimizes the electronic distribution of active sites by constructing unsaturated bonds in the carbon skeleton to induce localized charge enrichment. Yang et al. [75] fabricated Fe1.88C0.12@CNTs catalysts (Figure 5g) that significantly enhanced d-p orbital hybridization through precisely controlled Fe defects, facilitating Fe-S bond formation and accelerating reaction kinetics. The cathode material maintained a reversible capacity of 802.93 mAh g−1 after 200 cycles at 0.2C. A solid–liquid conversion pathway of Na3FeS3↔Na6Fe1.88S6 was confirmed by Raman spectroscopy (Figure 5h), which substantially shortened the reaction pathway compared to conventional conversion mechanisms.

3.2.4. Alloying

The alloying strategy effectively modulates the electronic structure of active sites through electron transfer between heterometallic elements. Wang et al. [79] employed an Fe-Ni alloying approach, where the electronegativity difference (Fe < Ni) induced electron transfer from Fe to Ni, achieving dual functional regulation: (1) the increased electron density in Ni 3d orbitals enhanced charge transfer efficiency, while (2) the decreased electron density in Fe 3d orbitals improved its capability to accept electrons from NaPSs. Enhanced characteristic peak intensities, which indicated efficient electron transfer from Ni to S species, were significantly revealed by Ni L-edge X-ray absorption near-edge structure (XANES) spectroscopy (Figure 5i). This electronic structure modulation endowed the material with an exceptional precipitation capacity of 166 mAh g−1 (Figure 5j), much higher than that of Ni@HC (129 mAh g−1).

3.3. TM Compound Catalysts

Despite the high activity of single-atom metal sites, they tend to aggregate and deactivate due to excessive surface energy. Meanwhile, metal nanoparticles suffer from low atomic utilization efficiency and uncontrolled sulfidation/oxidation in sulfur-containing environments. By contrast, metal compounds (e.g., sulfides [36,119], oxides [85,94], carbides [120,121]) exhibit superior stability, enabling precise charge state modulation through multi-dimensional strategies including crystal phase engineering, morphological control, defect construction, and composition optimization. This comprehensive modulation not only significantly enhances the chemisorption capability toward NaPSs but also dramatically improves the reaction kinetics of NaPS conversion, thereby comprehensively boosting the electrochemical performance of RT Na-S batteries.

3.3.1. Crystal Phase Engineering

Crystal phase engineering serves as a pivotal strategy for enhancing room-temperature RT Na-S battery performance by optimizing electronic conductivity and catalytic active sites. Huang et al. [92] demonstrated orthorhombic Nb2O5 as an efficient bidirectional redox catalyst that significantly improved battery reaction kinetics (Figure 6a). XPS analysis revealed that at a discharge potential of 1.8 V, Nb2O5 achieved the dual chemical anchoring of NaPSs through the formation of Nb-S and Na-O bonds, exhibiting exceptional sulfurphilic/sodiophilic characteristics. DFT calculations illustrated continuous charge density distribution between Nb2O5 and sulfur species (Na2S/Na2S6/Na2S8), establishing efficient electron transport channels. The accumulated charges on Nb-S and Na-O bonds substantially enhanced chemisorption of sulfur species and promoted multisite catalytic conversion from NaPSs to Na2S (Figure 6b). Crucially, Nb2O5 effectively weakened the Na-S bonds in adsorbed Na2S (0.04 Å and 0.12 Å bond length increase), reducing the decomposition activation energy and thereby dramatically accelerating the oxidation of Na2S to S8 (Figure 6c). This unique bidirectional catalytic mechanism concurrently accelerated NaPS reduction and sulfide oxidation, enabling highly efficient multisite catalysis for sulfur species conversion.

3.3.2. Morphology Engineering

Optimizing nanostructure design, particularly through precise construction of specific geometric configurations, represents a crucial strategy for enhancing RT Na-S battery performance by simultaneously maximizing active site exposure and significantly shortening ion/electron transport pathways. Zhang et al. [123] developed a three-dimensional N-doped porous-carbon-nanosheet-supported Fe3Se4 nanoparticle interlayer, where Fe active sites provided electrons to NaPSs, achieving a 1.5-fold enhancement in the conversion rate from long-chain NaPSs to Na2S. Dong et al. [124] employed a 2H-MoSe2/N-doped hollow carbon sphere/graphene oxide (2H-MoSe2/N-HCS/GO) composite separator coating that reduced the NaPS diffusion energy barrier by 58%. Aslam et al. [84] designed a yolk–shell-structured Fe2N-coated N-doped carbon (S/YS-Fe2N@NC) that combined high sulfur loading (3.1 mg cm−2) with effective volume expansion buffering, delivering a high reversible capacity after 200 cycles at 1C (Figure 6d). XPS analysis revealed a 0.8 eV negative shift in S 2p binding energy after Na2S6 adsorption (Figure 6e), confirming strong electron interaction. Symmetric cell tests demonstrated a big oxidation–reduction response (Figure 6f), representing a catalytic activity improvement. The performance enhancement fundamentally relies on nanostructures integrating abundant adsorption sites with high catalytic activity. Li et al. [89] developed a 2D/3D hybrid host (2D N-doped carbon nanosheets with Co4N-embedded 3D N-doped carbon polyhedrons) that achieved faster electron transfer and reduced the NaPS conversion barrier with high shuttle-effect-suppression efficiency.

3.3.3. Vacancy Defect Engineering

Precise modulation of defects (e.g., doping engineering, vacancy creation) in metal compounds can effectively alter the local electronic structure, creating electron-rich or electron-deficient regions. This electronic structure tailoring significantly enhances chemisorption capability through Lewis acid–base interactions with polar NaPS molecules. Specifically, vacancy defects establish localized charge-enriched zones that strongly anchor NaPSs. A representative study by Ma et al. [95] demonstrated that porous carbon-modified TiO2 nanoparticles underwent in situ reduction of Ti4+ to Ti3+ during sulfur loading, accompanied by decreased Ti-O bond ratios and oxygen vacancy formation (Figure 6g). The emergence and subsequent disappearance of the Na2Sx peak (2.1 V–1.8 V) during discharge indicated the strong adsorption and conversion of NaPSs by TiO2@SPC-S (Figure 6h). Most importantly, the defect sites lowered the energy barrier for Na2S solid-phase conversion, dramatically accelerating the conversion kinetics from S8 to Na2Sx.
Beyond incorporating catalytic materials into sulfur hosts, applied as functional coatings on separators, they can simultaneously serve as secondary current collectors and catalytic centers for NaPS conversion. Sun et al. [122] developed an amorphous iron–tin oxide (A-FeSnOx) nanosheet separator coating with hierarchical oxygen vacancies (Ovs), where Sn doping precisely regulated the oxygen vacancy concentration. Inferred from the Fe2+/Fe3+ area ratio (Figure 6i), dual functional sites were confirmed for efficient NaPS chemisorption and catalytic conversion. Compared to the Na2S6 solution, the peaks corresponding to S-S bonds in the A-FeSnOx-Na2S6 suspension exhibited significant broadening and reduced visibility (Figure 6j), indicating that the abundant Ovs defects promoted the disruption of S-S bonds and facilitated the strong interaction between A-FeSnOx and Na2S6.

3.4. TM-Based Heterostructure Catalysts

Although the aforementioned catalysts demonstrate specific advantages in RT Na-S battery cathodes, they all suffer from intrinsic limitations: single-atom metals are prone to aggregation and deactivation, nanoparticles exhibit phase instability during sulfur redox reactions, and metal compounds inherently show sluggish charge transfer. To overcome these challenges in charge transport efficiency, structural stability, and active site exposure, researchers have recently developed precisely constructed heterostructures that regulate interfacial electronic coupling effects at the atomic/nano-scale, achieving synergistic enhancement across multi-dimensional catalytic components. In particular, metal heterojunction catalysts have established a new paradigm through their unique mechanisms: (1) “adsorption–catalysis division of labor” (interface-induced electron redistribution), (2) “tandem stepwise catalysis” (multi-active-site relay catalysis), and (3) “heterointerface synergy” (built-in-electric-field-enhanced charge transfer). These mechanisms collectively improve NaPS-anchoring capacity and accelerate conversion kinetics, providing groundbreaking strategies for high-performance RT Na-S battery catalysts.

3.4.1. Synergistic Adsorption–Catalysis

Heterostructure catalysts achieve efficient synergy between NaPS chemisorption and electrocatalytic conversion through spatially separated component specialization. Ye et al. [106] constructed a TiN-TiO2 heterostructure within multichannel carbon fibers (Figure 7a) that demonstrated exceptional electrochemical stability, maintaining 49% capacity retention (640 mAh g−1) after 100 cycles at 0.1 A g−1 (Figure 7b). Mechanistic studies reveal the following: (1) the TiN component exhibits superior ionic conductivity, with a 30.4% lower Na+ migration barrier (0.156 eV) compared to TiO2 (0.224 eV), significantly enhancing NaPS redox kinetics (Figure 7c), and (2) the TiO2 component shows strong chemical anchoring of intermediates (Na2S4, −3.47 eV adsorption energy) and final products (Na2S, −3.58 eV), effectively suppressing NaPS shuttling (Figure 7d). This cooperative “TiN-dominant catalysis/TiO2-responsible adsorption” mechanism endows the heterojunction catalyst with both high sulfur utilization and outstanding cycling stability during charge/discharge processes.

3.4.2. Tandem Stepwise Catalysis

Heterojunction catalysts enable stepwise NaPS conversion through component synergy, establishing an efficient “adsorption–reduction/oxidation” tandem catalytic pathway. Zhang et al. [100] designed a core–shell Fe2O3@Fe(CN)64−–polypyrrole composite (Figure 7e) with dual catalytic sites: the Fe2O3 core preferentially catalyzes long-chain NaPS conversion to Na2S4 (the Tafel slope for the C1 peak to 201–227 mV dec−1, Figure 7f), while the Fe(CN)64−–polypyrrole shell specifically reduces Na2S4 to Na2S (the Tafel slope for the C2 peak to 193 mV dec−1). At a high sulfur loading of 5.8 mg cm−2, the S/P-Fe2O3@Fe-PPy cathode delivered an initial capacity of 606 mAh g−1 at 500 mA g−1 with 62% reduced polarization. Similarly, Fang et al. [101] developed a ZnS-NC@Ni-N4 core–shell catalyst achieving efficient NaPS conversion via tandem catalysis: (1) ZnS nanocrystals reduced the S8-to-NaPS conversion Gibbs free energy to 0.23 eV; (2) Ni-N4 single-atom sites not only reduced the short-chain NaPS-to-Na2S conversion Gibbs free energy to 1.32 eV but also induced a stable CEI layer, increasing Na2S precipitation by 35% and significantly enhancing interfacial stability.

3.4.3. Heterointerface Catalysisl

Heterostructure catalysts enhance battery performance through dual synergistic mechanisms: (1) adsorption–catalysis cooperation, where distinct components separately handle NaPS chemisorption and catalytic conversion, and (2) stepwise catalysis, achieving progressive NaPS conversion via ordered interfacial reaction pathways. Crucially, strong electronic interactions at heterointerfaces precisely modulate the adsorption free energy and reaction pathways of intermediates, dramatically improving catalytic efficiency. Zhao et al. [102] demonstrated that Ni-MnO2 heterostructures on N-doped carbon spheres (1) reduce NaPS adsorption energy through MnO2-Ni electron transfer while optimizing desorption kinetics and (2) lower the reaction barriers for liquid–solid (Na2S4→Na2S2) and solid–solid (Na2S2→Na2S) conversions to −2.53 eV and −3.43 eV, respectively, boosting conversion rates. Beyond pre-constructed heterostructures, in situ formation of heterointerfaces during electrochemical cycling can simultaneously achieve strong NaPS anchoring and efficient catalytic conversion. Wei et al. [109] confined cobalt nanoparticles within hollow carbon channels, where Co-S-C heterointerfaces were electrochemically generated in situ. This interface upshifted the Co d-band center, reducing the Na2S decomposition energy barrier to 0.6 eV and improving Na2S nucleation by 52%. The Co-S-C@MC cathode demonstrated exceptional cycling stability (only 0.999% capacity decay per cycle over 2500 cycles at 0.5C) while maintaining 1300 mAh g−1 reversible capacity even under harsh conditions (6 mg cm−2 sulfur loading, 9 μL mg−1 E/S ratio). Similarly, Wang et al. [107] revealed that amorphous Ni-B dynamically evolved into NiSx catalytic phases during cycling (Figure 7g). The cathode delivered 1487 mAh g−1 initial discharge capacity at 0.2 A g−1, retaining 31.6% capacity (470 mAh g−1) after 1000 cycles at 2 A g−1. At an ultrahigh sulfur loading of 8.9 mg cm−2, it achieved 5.75 mAh cm−2 areal capacity. Mechanistic studies demonstrated that (1) the peak current (0.312 mA) and deposition capacity (546 mAh g−1) of Na2S on the amorphous Ni-B surface confirmed better Na2S nucleation, dramatically boosting liquid–solid conversion kinetics (Figure 7h), and (2) amorphous Ni-B exhibited ultralow Gibbs free energy changes (ΔG) of 2.24 eV and −0.18 eV for Na2S4→Na2S2 and Na2S2→Na2S conversions, respectively (Figure 7i).

4. Summary and Perspectives

4.1. Summary

RT Na-S batteries have attracted significant interest due to their high theoretical capacity (1675 mAh g−1), low cost (USD 0.02 kWh−1), and elemental abundance (Na: 2.57%; S: 0.05% in crust) [125]. However, practical applications are hindered by NaPS shuttling and sluggish redox kinetics (conversion barriers > 1.5 eV). Over the past decade, transition metal-based catalysts have extended cycle life by 2–3 orders of magnitude through precise d-band modulation and interfacial charge redistribution (>0.3|e|transfer), offering breakthrough solutions. This review analyzes two core challenges: (1) extremely low intrinsic conductivity (<10−30 S cm−1) of S, causing severe polarization (>300 mV) at high sulfur loading (>5 mg cm−2), and (2) capacity decay exacerbated by disproportionation reactions during solid–liquid–solid phase transitions. Establishing quantitative structure–activity relationships between electronic features (d-band center, orbital hybridization) and NaPS adsorption energy (ΔEads)/activation barriers (Ea) provides theoretical foundations for catalyst design. Four catalyst categories have emerged: (1) single/dual-atom catalysts (<40 wt.% loading) achieve 100% atomic utilization via unsaturated coordination; (2) nanoparticles (2–30 nm) leverage high-index facets; (3) metal compounds (sulfides/oxides/carbides) regulate electron transfer via anion vacancies; and (4) heterojunctions utilize built-in fields for function separation.

4.2. Future Perspectives

Despite the advances of TM-based catalysts and enhanced performance, the commercialization of RT Na-S batteries has been impeded by the following five critical bottlenecks: (1) Limitations in characterization: Conventional ex situ techniques (e.g., postmortem XPS, SEM) fail to capture transient interfacial reactions occurring at nanosecond timescales during real-time charge/discharge processes, particularly the dynamic evolution of catalyst–NaPS interfaces and metastable intermediates like Na2S3. (2) Theoretical modeling gaps: The absence of quantitative electronic–kinetic correlation models prevents precise prediction of how specific electronic structures (e.g., d-band center position, charge transfer amount) govern reaction thermodynamics (adsorption energy ΔEads) and kinetics (activation barrier Ea for Na2S decomposition). (3) Practical energy density trade-offs: Excessive catalyst loading (>20 wt.%) significantly dilutes the energy density of battery systems. (4) Electrolyte sensitivity: Under lean electrolyte conditions (E/S ratio < 3 μL mg−1), the limited ion transport pathways cause rapid performance degradation. (5) Scalability challenges: High-temperature pyrolysis (typically > 800 °C) and inert-atmosphere requirements in synthesis (e.g., for Co@NC catalysts) not only increase manufacturing costs but also induce structural heterogeneity during scale-up, as evidenced by a reduction in active site density. These process-induced defects further accelerate capacity fading rates.
To address the existing challenges and propel RT Na-S batteries toward commercialization, several breakthrough pathways are envisioned for TM-based catalysts: (1) Advanced operando characterization platforms: The development of integrated multi-modal operando analysis systems combining time-resolved Raman spectroscopy (with <10 ns temporal resolution), synchrotron-based XRD (achieving sub-Ångström spatial resolution), and real-time DFT calculations will be crucial. These platforms will enable atomic-scale observation of dynamic interfacial evolution during electrochemical processes, particularly in capturing transient reaction intermediates (e.g., Na2S3 with a lifetime of <1 ms) and potential-dependent catalyst reconstruction. (2) AI-driven catalyst design systems: Machine learning-assisted platforms utilizing >104 curated experimental and computational datasets will revolutionize catalyst discovery. These systems will incorporate key descriptors, including d-band center positions (±0.2 eV precision), Bader charge transfer values, and adsorption energy profiles, to predict optimal catalyst compositions with >90% accuracy before synthesis. (3) Precision catalyst engineering: Achieving ultralow loading (<5 wt.%) through atomic-level dispersion techniques such as atomic layer deposition (ALD) and molecular layer deposition (MLD) will be essential. This approach can maximize active site exposure while maintaining exceptional catalytic activity, potentially achieving turnover frequencies (TOFs) >103 s−1 for NaPS conversion. (4) Green synthesis paradigms: Room-temperature wet-chemical synthesis methods will replace energy-intensive processes, potentially reducing manufacturing costs by >60% while eliminating the need for inert atmospheres. Techniques like electrochemical atomic layer epitaxy (EC-ALE) could enable precise control over catalyst morphology and composition.
Despite notable progress in lithium–sulfur (Li-S) batteries, their commercialization remains impeded by the high cost and poor scalability of key components. Experience from Li-S systems underscores the crucial role of transition metal catalysts, yet research remains largely confined to lab-scale studies. A fundamental insight, therefore, is the imperative to develop cost-effective and scalable synthesis methods for such catalysts to enable the practical deployment of transition metal-based Na-S batteries. The interdisciplinary integration of these approaches for synthesizing TM-based catalysts, combining advanced characterization, computational design, and sustainable manufacturing, will accelerate the transition of RT Na-S batteries from laboratory-scale prototypes (mAh level) to grid-scale energy storage systems (kWh level). Within the next 5–7 years, we anticipate that these developments will enable practical energy densities exceeding 300 Wh kg−1 at the system level, with a cycle life surpassing 2000 cycles at a 1C rate, finally meeting the stringent requirements for commercial energy storage applications.

Author Contributions

Conceptualization, J.L., Y.X. and Y.W.; methodology, Y.W., J.L. and Y.X.; software, Y.W.; validation, Y.Y., P.L. and H.C.; formal analysis, Y.Y., P.L. and H.C.; investigation, J.L. and Y.X.; resources, Y.W. and J.L.; data curation, Y.W.; writing—original draft preparation, all authors; writing—review and editing, all authors; visualization, Y.W.; supervision, J.L. and Y.X.; project administration, J.L.; funding acquisition, J.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grant 52271226), the provincial key research and development (R&D) program of Hubei (2023BAB005), and Fundamental Research Funds for the Central Universities (104972024KFYzxk0002, 104972025KFYzxk0010, 104972025KFYjc0125). The APC was funded by Junsheng Li.

Data Availability Statement

No new data were created in this study.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Bamisile, O.; Cai, D.; Adun, H.; Dagbasi, M.; Ukwuoma, C.C.; Huang, Q.; Johnson, N.; Bamisile, O. Towards renewables development: Review of optimization techniques for energy storage and hybrid renewable energy systems. Heliyon 2024, 10, e37482. [Google Scholar] [CrossRef]
  2. Fang, R.; Chen, K.; Sun, Z.; Hu, G.; Wang, D.W.; Li, F. Realizing high-energy density for practical lithium-sulfur batteries. Interdiscip. Mater. 2023, 2, 761–770. [Google Scholar] [CrossRef]
  3. Costa, C.M.; Pinto, R.S.; Serra, J.P.; Barbosa, J.C.; Gonçalves, R.; Lanceros-Méndez, S. Next generation sustainable lithium-ion batteries: Micro and nanostructured materials and processes. Chem. Eng. J. 2025, 509, 161337. [Google Scholar] [CrossRef]
  4. Joshi, B.; Samuel, E.; Kim, Y.-I.; Yarin, A.L.; Swihart, M.T.; Yoon, S.S. Progress and potential of electrospinning-derived substrate-free and binder-free lithium-ion battery electrodes. Chem. Eng. J. 2022, 430, 132876. [Google Scholar] [CrossRef]
  5. Fang, R.; Zhao, S.; Sun, Z.; Wang, D.W.; Cheng, H.M.; Li, F. More reliable lithium-sulfur batteries: Status, solutions and prospects. Adv. Mater. 2017, 29, 1606823. [Google Scholar] [CrossRef]
  6. Bhutia, P.T.; Grugeon, S.; El Mejdoubi, A.; Laruelle, S.; Marlair, G. Safety aspects of sodium-ion batteries: Prospective analysis from first generation towards more advanced systems. Batteries 2024, 10, 370. [Google Scholar] [CrossRef]
  7. Bruce Dunn, H.K.; Jean-Marie, T. Electrical energy storage for the grid: A battery of choices. Science 2011, 334, 928–935. [Google Scholar] [CrossRef]
  8. Hueso, K.B.; Armand, M.; Rojo, T. High temperature sodium batteries: Status, challenges and future trends. Energy Environ. Sci. 2013, 6, 734–749. [Google Scholar] [CrossRef]
  9. Virkar, A.V.; Miller, G.R.; Gordon, R.S. Resistivity-microstructure relations in lithia-stabilized polycrystalline β″-alumina. J. Am. Ceram. Soc. 2006, 61, 250–252. [Google Scholar] [CrossRef]
  10. Lu, X.; Kirby, B.W.; Xu, W.; Li, G.; Kim, J.Y.; Lemmon, J.P.; Sprenkle, V.L.; Yang, Z. Advanced intermediate-temperature Na–S battery. Energy Environ. Sci. 2013, 6, 299–306. [Google Scholar] [CrossRef]
  11. Nikiforidis, G.; van de Sanden, M.C.M.; Tsampas, M.N. High and intermediate temperature sodium–sulfur batteries for energy storage: Development, challenges and perspectives. RSC Adv. 2019, 9, 5649–5673. [Google Scholar] [CrossRef]
  12. Adelhelm, P.; Hartmann, P.; Bender, C.L.; Busche, M.; Eufinger, C.; Janek, J. From lithium to sodium: Cell chemistry of room temperature sodium-air and sodium-sulfur batteries. Beilstein J. Nanotechnol. 2015, 6, 1016–1055. [Google Scholar] [CrossRef] [PubMed]
  13. Park, C.W.; Ahn, J.H.; Ryu, H.S.; Kim, K.W.; Ahn, H.J. Room-temperature solid-state sodium/sulfur battery. Electrochem. Solid State Lett. 2006, 9, A123–A125. [Google Scholar] [CrossRef]
  14. Kim, I.; Park, J.-Y.; Kim, C.; Park, J.-W.; Ahn, J.-P.; Ahn, J.-H.; Kim, K.-W.; Ahn, H.-J. Sodium polysulfides during charge/discharge of the room-temperature Na/S Battery Using TEGDME electrolyte. J. Electrochem. Soc. 2016, 163, A611–A616. [Google Scholar] [CrossRef]
  15. Yim, T.; Park, M.-S.; Yu, J.-S.; Kim, K.J.; Im, K.Y.; Kim, J.-H.; Jeong, G.; Jo, Y.N.; Woo, S.-G.; Kang, K.S.; et al. Effect of chemical reactivity of polysulfide toward carbonate-based electrolyte on the electrochemical performance of Li–S batteries. Electrochim. Acta 2013, 107, 454–460. [Google Scholar] [CrossRef]
  16. Sungjemmenla; Soni, C.B.; Vineeth, S.K.; Kumar, V. Unveiling the physiochemical aspects of the matrix in improving sulfur-loading for room-temperature sodium-sulfur batteries. Mater. Adv. 2021, 2, 4165–4189. [Google Scholar] [CrossRef]
  17. Haridas, A.K.; Huang, C. Advances in strategic inhibition of polysulfide shuttle in room-temperature sodium-sulfur batteries via electrode and interface engineering. Batteries 2023, 9, 223. [Google Scholar] [CrossRef]
  18. Zhou, C.; Dong, C.; Wang, W.; Tian, Y.; Shen, C.; Yan, K.; Mai, L.; Xu, X. An ultrathin and crack-free metal-organic framework film for effective polysulfide inhibition in lithium-sulfur batteries. Interdiscip. Mater. 2024, 3, 306–315. [Google Scholar] [CrossRef]
  19. Hussain, T.; Kaewmaraya, T.; Hu, Z.; Zhao, X.S. Efficient control of the shuttle effect in sodium-sulfur batteries with functionalized nanoporous graphenes. ACS Appl. Nano Mater. 2022, 5, 12637–12645. [Google Scholar] [CrossRef]
  20. Hu, P.; Xiao, F.P.; Wu, Y.F.; Yang, X.M.; Li, N.; Wang, H.K.; Jia, J.F. Covalent encapsulation of sulfur in a graphene/N-doped carbon host for enhanced sodium-sulfur batteries. Chem. Eng. J. 2022, 443, 136257. [Google Scholar] [CrossRef]
  21. Yu, X.W.; Manthiram, A. Room-temperature sodium-sulfur batteries with liquid-phase sodium polysulfide catholytes and binder-free multiwall carbon nanotube fabric electrodes. J. Phys. Chem. C 2014, 118, 22952–22959. [Google Scholar] [CrossRef]
  22. Guo, Q.B.; Sun, S.; Kim, K.I.; Zhang, H.S.; Liu, X.J.; Yan, C.L.; Xia, H. A novel one-step reaction sodium-sulfur battery with high areal sulfur loading on hierarchical porous carbon fiber. Carbon Energy 2021, 3, 440–448. [Google Scholar] [CrossRef]
  23. Lu, Q.Q.; Wang, X.Y.; Cao, J.; Chen, C.; Chen, K.; Zhao, Z.F.; Niu, Z.Q.; Chen, J. Freestanding carbon fiber cloth/sulfur composites for flexible room-temperature sodium-sulfur batteries. Energy Storage Mater. 2017, 8, 77–84. [Google Scholar] [CrossRef]
  24. Wang, Y.; Indubala, E.; Ma, C.; Zhang, C.; Li, C.; Zhang, W.; Chen, Y.; Zhao, Y.; Xiao, L.; Lv, B.; et al. Nitrogen doped three-dimensionally interconnected macroporous/mesoporous carbon nanofibers as free-standing electrode for room temperature sodium sulfur batteries. Diam. Relat. Mater. 2025, 152, 111997. [Google Scholar] [CrossRef]
  25. Xia, G.L.; Zhang, L.J.; Chen, X.W.; Huang, Y.Q.; Sun, D.L.; Fang, F.; Guo, Z.P.; Yu, X.B. Carbon hollow nanobubbles on porous carbon nanofibers: An ideal host for high-performance sodium-sulfur batteries and hydrogen storage. Energy Storage Mater. 2018, 14, 314–323. [Google Scholar] [CrossRef]
  26. Dong, C.W.; Zhou, H.Y.; Liu, H.; Jin, B.; Wen, Z.; Lang, X.Y.; Li, J.C.; Kim, J.; Jiang, Q. Inhibited shuttle effect by functional separator for room-temperature sodium-sulfur batteries. J. Mater. Sci. Technol. 2022, 113, 207–216. [Google Scholar] [CrossRef]
  27. Xu, H.; Xiang, Y.; Xu, X.; Liang, Y.; Li, Y.; Qi, Y.; Xu, M. A polysulfides-defensive, dendrite-suppressed, and flame-retardant separator with lean electrolyte for room temperature sodium-sulfur batteries. Adv. Funct. Mater. 2024, 34, 2403663. [Google Scholar] [CrossRef]
  28. Yin, C.C.; Li, Z.; Zhao, D.C.; Yang, J.Y.; Zhang, Y.; Du, Y.; Wang, Y. Azo-branched covalent organic framework thin films as active separators for superior sodium-sulfur batteries. ACS Nano 2022, 16, 14178–14187. [Google Scholar] [CrossRef]
  29. Chen, S.F.; Liang, L.J.; Li, Y.Y.; Wang, D.Y.; Lu, J.G.; Zhan, X.L.; Hou, Y.; Zhang, Q.H.; Lu, J. Brain capillary-inspired self-assembled covalent organic framework membrane for sodium-sulfur battery separator. Adv. Energy Mater. 2023, 13, 2204334. [Google Scholar] [CrossRef]
  30. Zhang, L.L.; Zhang, W.H.; Zhu, Z.Y.; Huang, Q.Q.; Liu, X.X.; Zhang, M.C.; Pei, W.B.; Wu, J.S. Multi-channel sulfurized polyacrylonitrile with hollow structure as cathode for room temperature sodium-sulfur batteries. J. Solid State Chem. 2021, 301, 122359. [Google Scholar] [CrossRef]
  31. Huang, X.L.; Wang, Y.X.; Chou, S.L.; Dou, S.X.; Wang, Z.M.M. Materials engineering for adsorption and catalysis in room-temperature Na-S batteries. Energy Environ. Sci. 2021, 14, 3757–3795. [Google Scholar] [CrossRef]
  32. Zhang, X.; Gao, W.; Chen, Y.; Peng, Y.; Liu, X.; Yang, X.; Xiong, X.; Wang, J.; Liu, Y.; Jia, A.; et al. Integrated adsorption-catalysis design enabling high-performance sodium-sulfur batteries. Sustain. Energy Fuels 2025, 9, 3754–3779. [Google Scholar] [CrossRef]
  33. Meng, T.; Geng, Z.; Ma, F.; Wang, X.; Zhang, H.; Guan, C. Direct ink writing of metal-based electrocatalysts for Li-S batteries with efficient polysulfide conversion. Interdiscip. Mater. 2023, 2, 589–608. [Google Scholar] [CrossRef]
  34. Joshi, A.; Mohapatra, S.; Gupta, A.; Nandan, B. Review of metal nitrides for lithium-sulfur batteries: Design, mechanisms, and prospects. Chem. A Eur. J. 2025, 31, e202500971. [Google Scholar] [CrossRef]
  35. Liu, Y.P.; Bettels, F.; Lin, Z.H.; Li, Z.H.; Shao, Y.X.; Ding, F.; Liu, S.Y.; Zhang, L. Recent advances in transition-metal-based catalytic material for room-temperature sodium-sulfur batteries. Adv. Funct. Mater. 2023, 34, 2302626. [Google Scholar] [CrossRef]
  36. Fan, M.P.; Chen, Y.M.; Ke, X.; Huang, Z.X.; Chen, Y.C.; Wu, W.L.; Qu, X.F.; Shi, Z.C.; Guo, Z.P. In situ growth of NiS2 nanosheet array on Ni foil as cathode to improve the performance of lithium/sodium-sulfur batteries. Sci. China-Technol. Sci. 2022, 65, 231–237. [Google Scholar] [CrossRef]
  37. Wang, T.; Li, W.; Fu, Y.; Wang, D.; Wu, L.; Sun, K.; Liu, D.; Ma, R.; Shi, Y.; Yang, G.; et al. A mott-schottky heterojunction with strong chemisorption and fast conversion effects for room-temperature Na-S batteries. Small 2024, 20, 2311180. [Google Scholar] [CrossRef] [PubMed]
  38. Xu, J.; Kang, Q.; Peng, B.; Zhuang, Z.; Wang, D.; Ma, L. Engineering single-atom catalysts for sulfur electrochemistry in metal-sulfur batteries. J. Energy Chem. 2025, 106, 768–790. [Google Scholar] [CrossRef]
  39. Song, H.; Li, Y.; Li, X.L.; Li, Y.; Li, D.s.; Wang, D.; Huang, S.; Yang, H.Y. Recent progress in heterostructured materials for room-temperature sodium-sulfur batteries. Interdiscip. Mater. 2024, 3, 565–594. [Google Scholar] [CrossRef]
  40. Yao, G.; Li, Z.Q.; Zhang, Y.H.; Xiao, Y.; Wei, L.Z.; Niu, H.L.; Chen, Q.W.; Yang, Y.; Zheng, F.C. Highly flexible carbon film implanted with single-atomic Zn-N2 moiety for long-life sodium-sulfur batteries. Adv. Funct. Mater. 2023, 34, 2214353. [Google Scholar] [CrossRef]
  41. Wang, Y.; Huang, X.L.; Liu, H.W.; Qiu, W.L.; Feng, C.; Li, C.; Zhang, S.H.; Liu, H.K.; Dou, S.X.; Wang, Z.M.M. Nanostructure engineering strategies of cathode materials for room-temperature Na-S batteries. ACS Nano 2022, 16, 5103–5130. [Google Scholar] [CrossRef]
  42. Yu, X.W.; Manthiram, A. Capacity enhancement and discharge mechanisms of room-temperature sodium-sulfur batteries. Chemelectrochem 2014, 1, 1275–1280. [Google Scholar] [CrossRef]
  43. Gray, E.L.; Lee, J.-I.; Li, Z.; Moloney, J.; Yang, Z.J.; Chhowalla, M. Mapping Polysulfides in Sodium–Sulfur Batteries. ACS Nano 2025, 19, 8939–8947. [Google Scholar] [CrossRef]
  44. Zhong, X.; Huang, Y.J.; Cai, J.M.; Li, Y.J.; He, Z.D.; Cai, D.Y.; Geng, Z.L.; Deng, W.T.; Zou, G.Q.; Hou, H.S.; et al. Origin of the high catalytic activity of MoS2 in Na-S batteries: Electrochemically reconstructed Mo single atoms. J. Am. Chem. Soc. 2024, 146, 32124–32134. [Google Scholar] [CrossRef]
  45. Jin, F.; Wang, R.; Liu, Y.; Zhang, N.; Bao, C.; Li, D.; Wang, D.; Cheng, T.; Liu, H.; Dou, S.; et al. Conversion mechanism of sulfur in room-temperature sodium-sulfur battery with carbonate-based electrolyte. Energy Storage Mater. 2024, 69, 103388. [Google Scholar] [CrossRef]
  46. Yang, Z.Z.; Xiao, R.; Zhang, X.Y.; Wang, X.; Zhang, D.; Sun, Z.H.; Li, F. Role of catalytic materials on conversion of sulfur species for room temperature sodium-sulfur battery. Energy Environ. Mater. 2022, 5, 693–710. [Google Scholar] [CrossRef]
  47. Liu, Y.; Li, X.; Sun, Y.; Yang, R.; Lee, Y.; Ahn, J.-H. Dual-porosity carbon derived from waste bamboo char for room-temperature sodium-sulfur batteries using carbonate-based electrolyte. Ionics 2020, 27, 199–206. [Google Scholar] [CrossRef]
  48. Chen, P.; Wang, C.Y.; Wang, T.Y. Review and prospects for room-temperature sodium-sulfur batteries. Mater. Res. Lett. 2022, 10, 691–719. [Google Scholar] [CrossRef]
  49. Wang, L.; Wang, T.; Peng, L.L.; Wang, Y.L.; Zhang, M.; Zhou, J.; Chen, M.X.; Cao, J.H.; Fei, H.L.; Duan, X.D.; et al. The promises, challenges and pathways to room-temperature sodium-sulfur batteries. Natl. Sci. Rev. 2022, 9, nwab050. [Google Scholar] [CrossRef]
  50. Manthiram, A.; Yu, X.W. Ambient temperature sodium-sulfur batteries. Small 2015, 11, 2108–2114. [Google Scholar] [CrossRef]
  51. Yang, Q.J.; Yang, T.R.; Gao, W.; Qi, Y.R.; Guo, B.S.; Zhong, W.; Jiang, J.; Xu, M.W. An MXene-based aerogel with cobalt nanoparticles as an efficient sulfur host for room-temperature Na-S batteries. Inorg. Chem. Front. 2020, 7, 4396–4403. [Google Scholar] [CrossRef]
  52. Cengiz, E.C.; Erdol, Z.; Sakar, B.; Aslan, A.; Ata, A.; Ozturk, O.; Demir-Cakan, R. Investigation of the effect of using Al2O3-nafion barrier on room-temperature Na-S batteries. J. Phys. Chem. C 2017, 121, 15120–15126. [Google Scholar] [CrossRef]
  53. Fan, B.; Chen, W.; Li, K.; Wei, Q.; He, Q.; Liu, W.; Zhou, B.; Yuan, J.; Zou, Y. Synergistic adsorption and catalytic effects of Ti3C2Tx/CoO/MoO3 composite on lithium polysulfides for high-erformance lithium-sulfur batteries. Interdiscip. Mater. 2024, 3, 726–737. [Google Scholar]
  54. Mou, J.R.; Li, Y.J.; Liu, T.; Zhang, W.J.; Li, M.; Xu, Y.T.; Zhong, L.; Pan, W.H.; Yang, C.H.; Huang, J.L.; et al. Metal-organic frameworks-derived nitrogen-doped porous carbon nanocubes with embedded Co nanoparticles as efficient sulfur immobilizers for room temperature sodium-sulfur batteries. Small Methods 2021, 5, 2100455. [Google Scholar] [CrossRef]
  55. Zhang, H.; Wang, M.; Huang, X.-L.; Lu, S.; Lu, K.; Wu, X. Atomic manganese manipulating polysulfide speciation pathway for room-temperature Na-S batteries. CCS Chem. 2024, 6, 2289–2304. [Google Scholar] [CrossRef]
  56. Zhang, B.W.; Sheng, T.; Liu, Y.D.; Wang, Y.X.; Zhang, L.; Lai, W.H.; Wang, L.; Yang, J.P.; Gu, Q.F.; Chou, S.L.; et al. Atomic cobalt as an efficient electrocatalyst in sulfur cathodes for superior room-temperature sodium-sulfur batteries. Nat. Commun. 2018, 9, 4082. [Google Scholar] [CrossRef] [PubMed]
  57. Zhang, B.W.; Li, S.N.; Yang, H.L.; Liang, X.H.; Lai, W.H.; Zhao, S.L.; Dong, J.C.; Chu, S.Q.; Gu, Q.F.; Liang, J.; et al. Atomically dispersed S-Fe-N4 for fast kinetics sodium-sulfur batteries via a dual function mechanism. Cell Rep. Phys. Sci. 2021, 2, 100531. [Google Scholar] [CrossRef]
  58. Ruan, J.; Lei, Y.-J.; Fan, Y.; Borras, M.C.; Luo, Z.; Yan, Z.; Johannessen, B.; Gu, Q.; Konstantinov, K.; Pang, W.K.; et al. Linearly interlinked Fe-Nx-Fe single atoms catalyze high-rate sodium-sulfur batteries. Adv. Mater. 2024, 36, 2312207. [Google Scholar] [CrossRef] [PubMed]
  59. Xiao, F.P.; Wang, H.K.; Xu, J.; Yang, W.Q.; Yang, X.M.; Yu, D.Y.W.; Rogach, A.L. Generating short-chain sulfur suitable for efficient sodium-sulfur batteries via atomic copper sites on a N,O-codoped carbon composite. Adv. Energy Mater. 2021, 11, 2100989. [Google Scholar] [CrossRef]
  60. Liu, R.; Feng, C.; Wu, P.; Sun, Y.; Chu, Z.; Hu, J.; Chen, W.; Guo, L.; Huang, Q.; Wang, D. Improving conversion kinetics of sodium polysulfides through electron spillover effect with V/Co dual-atomic site anchoring on N-doped MXene. Adv. Mater. 2025, 37, 2501371. [Google Scholar] [CrossRef]
  61. Li, C.; Yu, J.; Yang, D.; Li, H.; Cheng, Y.; Ren, Y.; Bi, X.; Ma, J.; Zhao, R.; Zhou, Y.; et al. Balancing electronic spin state via atomically-dispersed heteronuclear Fe-Co pairs for high-performance sodium-sulfur batteries. J. Am. Chem. Soc. 2025, 147, 8250–8259. [Google Scholar] [CrossRef]
  62. Hu, P.; Wu, Y.-F.; Gao, X.-P.; Huang, L.; Cai, B.-B.; Liu, Y.-X.; Ma, Y.; Jiang, S.; Wang, F.; Xiao, F.-P. N/O dual coordination of cobalt single atom for fast kinetics sodium-sulfur batteries. Rare Met. 2025, 44, 288–299. [Google Scholar] [CrossRef]
  63. Tian, H.; Lei, Y.; Sun, B.; Yang, C.-C.; Chen, C.-L.; Huang, T.; Zhang, X.; Chen, Y.; Song, A.; Pang, L.; et al. P-d orbital hybridization induced by transition metal atom sites for room-temperature sodium-sulfur batteries. Natl. Sci. Rev. 2025, 12, nwaf241. [Google Scholar]
  64. Song, W.; Wen, Z.; Wang, X.; Qian, K.; Zhang, T.; Wang, H.; Ding, J.; Hu, W. Unsaturation degree of Fe single atom site manipulates polysulfide behavior in sodium-sulfur batteries. Nat. Commun. 2025, 16, 2795. [Google Scholar] [CrossRef] [PubMed]
  65. Fang, D.L.; Huang, S.Z.; Xu, T.T.; Sun, P.; Li, X.L.; Lim, Y.V.; Yan, D.; Shang, Y.; Su, B.J.; Juang, J.Y.; et al. Low-coordinated Zn−N2 sites as bidirectional atomic catalysis for room-temperature Na−S batteries. ACS Appl. Mater. Interfaces 2023, 15, 26650–26659. [Google Scholar] [CrossRef] [PubMed]
  66. Zheng, F.; Zhang, Y.; Li, Z.; Yao, G.; Wei, L.; Wang, C.; Chen, Q.; Wang, H. Axial ligand induces the charge localization of Ca single-atom sites for efficient Na–S batteries. Nat. Commun. 2025, 16, 4372. [Google Scholar] [CrossRef] [PubMed]
  67. Wu, G.; Liu, T.; Lao, Z.; Cheng, Y.; Wang, T.; Mao, J.; Zhang, H.; Liu, E.; Shi, C.; Zhou, G.; et al. Optimizing s-p orbital overlap between sodium polysulfides and single-atom indium catalyst for efficient sulfur redox reaction. Angew. Chem. Int. Ed. 2025, 64, e202422208. [Google Scholar] [CrossRef]
  68. Jin, F.; Ning, Y.; Wang, B.; Ren, Z.H.; Luo, H.; Zhang, Z.K.; Zhang, N.; Wang, D.L. Cobalt, nitrogen co-doped microporous carbon matrix derived from metal organic frameworks toward high specific capacity room temperature sodium sulfur batteries. J. Power Sources 2023, 565, 232917. [Google Scholar] [CrossRef]
  69. Jiang, Y.; Yu, Z.X.; Zhou, X.F.; Cheng, X.L.; Huang, H.J.; Liu, F.F.; Yang, Y.X.; He, S.N.; Pan, H.G.; Yang, H.; et al. Single-atom vanadium catalyst boosting reaction kinetics of polysulfides in Na-S batteries. Adv. Mater. 2023, 35, 2208873. [Google Scholar] [CrossRef]
  70. Li, Z.; Chen, X.; Yao, G.; Wei, L.; Chen, Q.; Luo, Q.; Zheng, F.; Wang, H. Strengthening d-p orbital-hybridization via coordination number regulation of manganese single-atom catalysts toward fast kinetic and long-life sodium-sulfur batteries. Adv. Funct. Mater. 2024, 34, 2400859. [Google Scholar] [CrossRef]
  71. Zheng, F.; Chen, F.; Li, Z.; Yao, G.; Dong, S.; Wei, L.; Chen, Q.; Wang, C.; Wang, H. Template-sacrificing synthesis of asymmetrically coordinated Zn single-atom sites for high-performance sodium-sulfur batteries. Adv. Funct. Mater. 2025, 35, 2413084. [Google Scholar] [CrossRef]
  72. Bai, R.L.; Lin, Q.S.; Li, X.Y.; Ling, F.X.; Wang, H.J.; Tan, S.; Hu, L.X.; Ma, M.Z.; Wu, X.J.; Shao, Y.; et al. Toward complete transformation of sodium polysulfides by regulating the second-shell coordinating environment of atomically dispersed Fe. Angew. Chem.-Int. Ed. 2023, 62, e202218165. [Google Scholar] [CrossRef]
  73. Lei, Y.-J.; Lu, X.; Yoshikawa, H.; Matsumura, D.; Fan, Y.; Zhao, L.; Li, J.; Wang, S.; Gu, Q.; Liu, H.-K.; et al. Understanding the charge transfer effects of single atoms for boosting the performance of Na-S batteries. Nat. Commun. 2024, 15, 3325. [Google Scholar] [CrossRef] [PubMed]
  74. Bai, R.; Yao, Y.; Lin, Q.; Wu, L.; Li, Z.; Wang, H.; Ma, M.; Mu, D.; Hu, L.; Yang, H.; et al. Preferable single-atom catalysts enabled by natural language processing for high energy density Na-S batteries. Nat. Commun. 2025, 16, 5827. [Google Scholar] [CrossRef] [PubMed]
  75. Yang, C.S.; Zhang, H.Y. Boosting electrocatalytic activity via introducing carbon vacancies and iron defects towards high-performance room-temperature sodium-sulfur batteries. Appl. Mater. Today 2025, 42, 102575. [Google Scholar] [CrossRef]
  76. Mei, T.; Li, X.; Lin, X.; Bai, L.; Xu, M.; Qi, Y. Cobalt catalytic regulation engineering in room-temperature sodium-sulfur batteries: Facilitating rapid polysulfides conversion and delicate Na2S nucleation. Adv. Funct. Mater. 2025, 35, 2418126. [Google Scholar] [CrossRef]
  77. Zhang, R.X.; Esposito, A.M.; Thornburg, E.S.; Chen, X.Y.; Zhang, X.Y.; Philip, M.A.; Magana, A.; Gewirth, A.A. Conversion of Co nanoparticles to CoS in metal−organic framework-derived porous carbon during cycling facilitates Na2S reactivity in a Na−S battery. ACS Appl. Mater. Interfaces 2020, 12, 29285–29295. [Google Scholar] [CrossRef]
  78. Du, W.Y.; Gao, W.; Yang, T.T.; Guo, B.S.; Zhang, L.Z.; Bao, S.J.; Chen, Y.M.; Xu, M.W. Cobalt nanoparticles embedded into free-standing carbon nanofibers as catalyst for room-temperature sodium-sulfur batteries. J. Colloid Interface Sci. 2020, 565, 63–69. [Google Scholar] [CrossRef]
  79. Wang, L.F.; Wang, H.Y.; Zhang, S.P.; Ren, N.Q.; Wu, Y.; Wu, L.; Zhou, X.F.; Yao, Y.; Wu, X.J.; Yu, Y. Manipulating the electronic structure of nickel via alloying with iron: Toward high-kinetics sulfur cathode for Na-S batteries. ACS Nano 2021, 15, 15218–15228. [Google Scholar] [CrossRef]
  80. Ma, Q.R.; Zhong, W.; Du, G.Y.; Qi, Y.R.; Bao, S.J.; Xu, M.W.; Li, C.M. Multi-step controllable catalysis method for the defense of sodium polysulfide dissolution in room-temperature Na-S batteries. ACS Appl. Mater. Interfaces 2021, 13, 11852–11860. [Google Scholar] [CrossRef]
  81. Tang, K.; Peng, X.; Zhang, Z.; Li, G.; Wang, J.; Wang, Y.; Chen, C.; Zhang, N.; Xie, X.; Wu, Z. A highly dispersed cobalt electrocatalyst with electron-deficient centers induced by boron toward enhanced adsorption and electrocatalysis for room-temperature sodium-sulfur batteries. Small 2024, 20, 2311151. [Google Scholar] [CrossRef] [PubMed]
  82. Liu, Y.P.; Ma, S.Y.; Rosebrock, M.; Rusch, P.; Barnscheidt, Y.; Wu, C.Q.; Nan, P.F.; Bettels, F.; Lin, Z.H.; Li, T.R.; et al. Tungsten nanoparticles accelerate polysulfides conversion: A viable route toward stable room-temperature sodium-sulfur batteries. Adv. Sci. 2022, 9, 2105544. [Google Scholar] [CrossRef] [PubMed]
  83. Wu, Y.F.; Xu, Q.Q.; Huang, L.; Huang, B.; Hu, P.; Xiao, F.P.; Li, N. Encapsulation of sulfur in MoS2-modified metal-organic framework-derived N, O-codoped carbon host for sodium-sulfur batteries. J. Colloid Interface Sci. 2024, 654, 649–659. [Google Scholar] [CrossRef] [PubMed]
  84. Aslam, M.K.; Hussain, T.; Tabassum, H.; Wei, Z.; Tang, W.W.; Li, S.; Bao, S.J.; Zhao, X.S.; Xu, M.W. Sulfur encapsulation into yolk-shell Fe2N@nitrogen doped carbon for ambient-temperature sodium-sulfur battery cathode. Chem. Eng. J. 2022, 429, 132389. [Google Scholar] [CrossRef]
  85. Huang, X.L.; Xiang, P.; Liu, H.W.; Feng, C.; Zhang, S.H.; Tian, Z.Q.; Liu, H.K.; Dou, S.X.; Wang, Z.M. In situ implanting MnO nanoparticles into carbon nanorod-assembled microspheres enables performance-enhanced room-temperature Na-S batteries. Inorg. Chem. Front. 2022, 9, 5486–5494. [Google Scholar] [CrossRef]
  86. Shi, Y.; Zhang, L.; Wang, T.; Ma, R.; Wang, D.; Fu, Y.; Du, R.; Zhang, J.; Liu, D.; Wu, L.; et al. Optimizing adsorption-catalysis synergy to accelerate sulfur conversion kinetics in room-temperature Na-S batteries. Small 2025, 21, e2502257. [Google Scholar] [CrossRef]
  87. Zhou, X.F.; Yu, Z.X.; Yao, Y.; Jiang, Y.; Rui, X.H.; Liu, J.Q.; Yu, Y. A high-efficiency Mo2C electrocatalyst promoting the polysulfide redox kinetics for Na-S batteries. Adv. Mater. 2022, 34, 2200479. [Google Scholar] [CrossRef]
  88. Ye, C.; Jin, H.Y.; Shan, J.Q.; Jiao, Y.; Li, H.; Gu, Q.F.; Davey, K.; Wang, H.H.; Qiao, S.Z. A Mo5N6 electrocatalyst for efficient Na2S electrodeposition in room-temperature sodium-sulfur batteries. Nat. Commun. 2021, 12, 7195. [Google Scholar] [CrossRef]
  89. Li, Y.; Wang, X.Z.; Sun, M.H.; Ai, L.S.; Qin, L.; Zhao, Z.B.; Qiu, J.S. Co4N embedded nitrogen doped carbon with 2D/3D hybrid structure as sulfur host for room-temperature sodium-sulfur batteries. Electrochim. Acta 2023, 451, 142288. [Google Scholar] [CrossRef]
  90. Qi, Y.R.; Li, Q.J.; Wu, Y.K.; Bao, S.J.; Li, C.M.; Chen, Y.M.; Wang, G.X.; Xu, M.W. A Fe3N/carbon composite electrocatalyst for effective polysulfides regulation in room-temperature Na-S batteries. Nat. Commun. 2021, 12, 6347. [Google Scholar] [CrossRef]
  91. Wang, H.; Qi, Y.R.; Xiao, F.Y.; Liu, P.; Li, Y.; Bao, S.J.; Xu, M.W. Tessellated N-doped carbon/CoSe2 as trap-catalyst sulfur hosts for room-temperature sodium-sulfur batteries. Inorg. Chem. Front. 2022, 9, 1743–1751. [Google Scholar] [CrossRef]
  92. Huang, X.L.; Zhang, X.F.; Zhou, L.J.; Guo, Z.P.; Liu, H.K.; Dou, S.X.; Wang, Z.M. Orthorhombic Nb2O5 decorated carbon nanoreactors enable bidirectionally regulated redox behaviors in room-temperature Na-S batteries. Adv. Sci. 2023, 10, 2206558. [Google Scholar] [CrossRef]
  93. Ma, Q.Y.; Liu, Q.Q.; Li, Z.Y.; Pu, J.; Mujtaba, J.; Fang, Z. Oxygen vacancy-mediated amorphous GeOx assisted polysulfide redox kinetics for room-temperature sodium-sulfur batteries. J. Colloid Interface Sci. 2023, 629, 76–86. [Google Scholar] [CrossRef] [PubMed]
  94. Li, W.; Wang, T.; Yang, G.; Ma, R.; Li, W.; Fu, Y.; Wu, L.; Liu, Z.; Liu, D.; Wu, Y.; et al. Oxygen vacancy modified cerium dioxide with excellent catalytic activity for room-temperature Na-S batteries. J. Energy Storage 2024, 91, 112141. [Google Scholar] [CrossRef]
  95. Ma, R.; Zhang, L.; Shi, Y.; Fu, Y.; Wang, D.; Wang, T.; Yang, G.; Zhang, J.; Wu, L.; Liu, D.; et al. Enhanced polysulfide conversion in room-temperature sodium-sulfur batteries via nanoscale TiO2 modified porous carbon structures. J. Colloid Interface Sci. 2025, 684, 235–242. [Google Scholar] [CrossRef] [PubMed]
  96. Gao, W.J.; Song, B.Y.; Zhang, Q.Y.; He, J.R.; Wu, Y.P. 3D flower-like nanospheres constructed by transition metal telluride nanosheets as sulfur immobilizers for high-performance room-temperature Na-S batteries. Small 2023, 20, 2310225. [Google Scholar] [CrossRef]
  97. Qian, S.Y.; Yuan, Z.Y.; Li, G.S.; Li, D.D.; Li, J.Z.; Han, W. 3D layered structure Ti3C2Tx MXene/Ni(OH)2/C with strong catalytic and adsorption capabilities of polysulfides for high-capacity Sodium-Sulfur battery. Chem. Eng. J. 2023, 471, 144528. [Google Scholar] [CrossRef]
  98. Huang, C.; Yu, J.; Lei, Y.-J.; Usoltsev, O.; Gong, L.; Cui, Z.; Li, J.; Li, C.; Nan, B.; Lu, X.; et al. Generation of unpaired electrons to promote electron transfer at the cathode of room-temperature sodium sulfur batteries. Chem. Eng. J. 2025, 506, 160146. [Google Scholar] [CrossRef]
  99. Lei, Y.J.; Wu, C.; Lu, X.X.; Hua, W.B.; Li, S.B.; Liang, Y.R.; Liu, H.W.; Lai, W.H.; Gu, Q.F.; Cai, X.L.; et al. Streamline sulfur redox reactions to achieve efficient room-temperature sodium-sulfur batteries. Angew. Chem. Int. Ed. 2022, 61, e202200384. [Google Scholar] [CrossRef]
  100. Zhang, H.; Song, B.; Zhang, W.; An, B.; Fu, L.; Lu, S.; Cheng, Y.; Chen, Q.; Lu, K. Bidirectional tandem electrocatalysis manipulated sulfur speciation pathway for high-capacity and stable Na-S battery. Angew. Chem. Int. Ed. 2023, 62, e202217009. [Google Scholar] [CrossRef]
  101. Fang, D.L.; Ghosh, T.; Huang, S.Z.; Wang, Y.; Qiu, J.B.; Xu, X.H.; Yang, H.Y. Core-shell tandem catalysis coupled with interface engineering for high-performance room-temperature Na-S batteries. Small 2023, 19, 2302461. [Google Scholar] [CrossRef]
  102. Zhao, H.; Li, Z.; Liu, M.; Ge, S.; Ma, K.; Jiao, Q.; Li, H.; Xie, H.; Zhao, Y.; Feng, C. Synergistically promoting the catalytic conversion of polysulfides via in-situ construction of Ni-MnO2 heterostructure on N-doped hollow carbon spheres towards high-performance sodium-sulfur batteries. Chem. Eng. J. 2024, 494, 153006. [Google Scholar] [CrossRef]
  103. Zhang, S.; Huang, M.; Wang, Y.; Wang, Z.; Wang, H.; Liu, X. Achieving a quasi-solid-state conversion of polysulfides via building high efficiency heterostructure for room temperature Na-S batteries. Adv. Energy Mater. 2024, 14, 2303925. [Google Scholar] [CrossRef]
  104. Qin, G.; Liu, Y.; Han, P.; Cao, S.; Guo, X.; Guo, Z. High performance room temperature Na-S batteries based on FCNT modified Co3C-Co nanocubes. Chem. Eng. J. 2020, 396, 125295. [Google Scholar] [CrossRef]
  105. Zhang, S.P.; Yao, Y.; Jiao, X.J.; Ma, M.Z.; Huang, H.J.; Zhou, X.F.; Wang, L.F.; Bai, J.T.; Yu, Y. Mo2N-W2N heterostructures embedded in spherical carbon superstructure as highly efficient polysulfide electrocatalysts for stable room-temperature Na-S batteries. Adv. Mater. 2021, 33, 2103846. [Google Scholar] [CrossRef] [PubMed]
  106. Ye, X.; Ruan, J.F.; Pang, Y.P.; Yang, J.H.; Liu, Y.F.; Huang, Y.Z.; Zheng, S.Y. Enabling a stable room-temperature sodium-sulfur battery cathode by building heterostructures in multichannel carbon fibers. ACS Nano 2021, 15, 5639–5648. [Google Scholar] [CrossRef] [PubMed]
  107. Wang, B.; Wang, L.; Guo, B.; Kong, Y.; Wang, F.; Jing, Z.; Qu, G.; Mamoor, M.; Wang, D.; He, X.; et al. In situ electrochemical evolution of amorphous metallic borides enabling long cycling room-/subzero-temperature sodium-sulfur batteries. Adv. Mater. 2024, 36, 2411725. [Google Scholar] [CrossRef] [PubMed]
  108. Zhang, B.W.; Cao, L.Y.; Tang, C.; Tan, C.H.; Cheng, N.Y.; Lai, W.H.; Wang, Y.X.; Cheng, Z.X.; Dong, J.C.; Kong, Y.; et al. Atomically dispersed dual-site cathode with a record high sulfur mass loading for high-performance room-temperature sodium-sulfur batteries. Adv. Mater. 2023, 35, 2206828. [Google Scholar] [CrossRef]
  109. Wei, X.; Zhang, Z.; Luo, D.; Wang, X. Construction of heterointerfaced nanoreactor electrocatalyst via in situ evolution toward practical room-temperature sodium-sulfur batteries. Adv. Funct. Mater. 2025, 35, 2414172. [Google Scholar] [CrossRef]
  110. Wu, J.H.; Yu, Z.X.; Yao, Y.; Wang, L.F.; Wu, Y.; Cheng, X.L.; Ali, Z.; Yu, Y. Bifunctional catalyst for liquid-solid redox conversion in room-temperature sodium-sulfur batteries. Small Struct. 2022, 3, 2200020. [Google Scholar] [CrossRef]
  111. Guo, D.; Zhang, X.; Liu, M.; Yu, Z.; Chen, X.a.; Yang, B.; Zhou, Z.; Wang, S. Single Mo-N4 atomic sites anchored on N-doped carbon nanoflowers as sulfur host with multiple immobilization and catalytic effects for high-performance lithium-sulfur batteries. Adv. Funct. Mater. 2022, 32, 2204458. [Google Scholar] [CrossRef]
  112. Sun, J.-F.; Xu, Q.-Q.; Qi, J.-L.; Zhou, D.; Zhu, H.-Y.; Yin, J.-Z. Isolated single atoms anchored on N-doped carbon materials as a highly efficient catalyst for electrochemical and organic reactions. ACS Sustain. Chem. Eng. 2020, 8, 14630–14656. [Google Scholar] [CrossRef]
  113. Gu, Y.; Xi, B.J.; Zhang, H.; Ma, Y.C.; Xiong, S.L. Activation of main-group antimony atomic sites for oxygen reduction catalysis. Angew. Chem. Int. Ed. 2022, 61, e202202200. [Google Scholar] [CrossRef]
  114. Guo, Y.; Jin, Z.; Lu, J.; Wei, L.; Wang, W.; Huang, Y.; Wang, A. Engineering a deficient-coordinated single-atom indium electrocatalyst for fast redox conversion in practical 500 Wh kg−1-level pouch lithium-sulfur batteries. Energy Environ. Sci. 2023, 16, 5274–5283. [Google Scholar] [CrossRef]
  115. Lao, Z.; Han, Z.; Ma, J.; Zhang, M.; Wu, X.; Jia, Y.; Gao, R.; Zhu, Y.; Xiao, X.; Yu, K.; et al. Band structure engineering and orbital orientation control constructing dual active sites for efficient sulfur redox reaction. Adv. Mater. 2023, 36, 2309024. [Google Scholar] [CrossRef] [PubMed]
  116. Lei, H.; Wang, Z.; Yang, F.; Huang, X.; Liu, J.; Liang, Y.; Xie, J.; Javed, M.S.; Lu, X.; Tan, S.; et al. NiFe nanoparticles embedded N-doped carbon nanotubes as high-efficient electrocatalysts for wearable solid-state Zn-air batteries. Nano Energy 2020, 68, 104293. [Google Scholar] [CrossRef]
  117. Du, W.Y.; Shen, K.Q.; Qi, Y.R.; Gao, W.; Tao, M.L.; Du, G.Y.; Bao, S.J.; Chen, M.Y.; Chen, Y.M.; Xu, M.W. Efficient catalytic conversion of polysulfides by biomimetic design of “branch-leaf” electrode for high-energy sodium-sulfur batteries. Nano-Micro Lett. 2021, 13, 50. [Google Scholar] [CrossRef]
  118. Guo, B.S.; Du, W.Y.; Yang, T.T.; Deng, J.H.; Liu, D.Y.; Qi, Y.R.; Jiang, J.; Bao, S.J.; Xu, M.W. Nickel hollow spheres concatenated by nitrogen-doped carbon fibers for enhancing electrochemical kinetics of sodium-sulfur batteries. Adv. Sci. 2020, 7, 1902617. [Google Scholar] [CrossRef]
  119. Yan, Z.C.; Xiao, J.; Lai, W.H.; Wang, L.; Gebert, F.; Wang, Y.X.; Gu, Q.F.; Liu, H.; Chou, S.L.; Liu, H.K.; et al. Nickel sulfide nanocrystals on nitrogen-doped porous carbon nanotubes with high-efficiency electrocatalysis for room-temperature sodium-sulfur batteries. Nat. Commun. 2019, 10, 4793. [Google Scholar] [CrossRef]
  120. Tang, W.W.; Zhong, W.; Wu, Y.K.; Qi, Y.R.; Guo, B.S.; Liu, D.Y.; Bao, S.J.; Xu, M.W. Vanadium carbide nanoparticles incorporation in carbon nanofibers for room-temperature sodium sulfur batteries: Confining, trapping, and catalyzing. Chem. Eng. J. 2020, 395, 124978. [Google Scholar] [CrossRef]
  121. Huang, X.L.; Hussain, T.; Liu, H.W.; Kaewmaraya, T.; Xu, M.W.; Liu, H.K.; Dou, S.X.; Wang, Z.M. Dredging sodium polysulfides using a Fe3C electrocatalyst to realize improved room-temperature Na–S batteries. Inorg. Chem. Front. 2023, 10, 4241–4251. [Google Scholar] [CrossRef]
  122. Sun, W.; Hou, J.; Zhou, Y.; Zhu, T.; Yuan, Q.; Wang, S.; Manshaii, F.; Song, C.; Lei, X.; Wu, X.; et al. Amorphous FeSnOx nanosheets with hierarchical vacancies for room-temperature sodium-sulfur batteries. Angew. Chem. Int. Ed. 2024, 63, e202404816. [Google Scholar] [CrossRef]
  123. Zhang, J.; Zhou, Y.; Shu, H.; Yan, Z.; Wu, Z.; Wang, Y.; Zhu, Z.; Wang, X. The design of chemisorption and catalysis synergistic defender for efficient room temperature sodium-sulfur batteries. J. Colloid Interface Sci. 2025, 678, 292–300. [Google Scholar] [CrossRef]
  124. Dong, C.W.; Zhou, H.Y.; Jin, B.; Gao, W.; Lang, X.Y.; Li, J.C.; Jiang, Q. Enabling high-performance room-temperature sodium/sulfur batteries with few-layer 2H-MoSe2 embellished nitrogen-doped hollow carbon spheres as polysulfide barriers. J. Mater. Chem. A. 2021, 9, 3451–3463. [Google Scholar] [CrossRef]
  125. Hou, R.Q.; Yuan, S.Y.; Wang, Y.G. Research status quo and prospect of electrolytes for room-temperature sodium-sulfur battery (Ⅰ). Dianchi (Battery Bimon.) 2024, 54, 3–8. (In Chinese) [Google Scholar]
Figure 1. Mechanism, classification, and influencing factors of transition metal-based catalysts for high-performance Na-S batteries. The mechanistic diagram in the center of this figure is adapted from Yao G. et al. [40], with specific modifications to the color scheme and the addition of the label “TM-Based Catalysts RT Na-S Systems”.
Figure 1. Mechanism, classification, and influencing factors of transition metal-based catalysts for high-performance Na-S batteries. The mechanistic diagram in the center of this figure is adapted from Yao G. et al. [40], with specific modifications to the color scheme and the addition of the label “TM-Based Catalysts RT Na-S Systems”.
Batteries 11 00333 g001
Figure 2. (a) Discharge–charge profile of Na-S batteries in ether-based electrolyte (1.0 M NaCF3SO3 in triethylene glycol dimethyl ether (TEGDME)) (adapted with permission from Ref. [43], American Chemical Society 2025). (b) Reaction step equations for RT Na-S batteries. (c) Discharge–charge profile of Na-S batteries in ester-based electrolyte (1 M NaClO4 in ethylene carbonate (EC)/propylene carbonate (PC) with a 1:1 volume ratio and 5% fluoroethylene carbonate (FEC) additive) (adapted with permission from Ref. [44], American Chemical Society 2024). (d) Comparison of binding energies between Na2S6 and ester-based solvent (EC) and ether-based solvent (diethylene glycol dimethyl ether, DEGDME). Pink and blue colors represent negative and positive binding energies, respectively. (adapted with permission from Ref. [45], Elsevier 2024). (e) Key challenges of sulfur cathode for RT Na-S batteries (adapted with permission from Ref. [46], John Wiley & Sons—Books 2022). (f) Schematic diagram illustrating the mechanisms and functions of catalysts (adapted with permission from Ref. [40], John Wiley & Sons—Books 2024).
Figure 2. (a) Discharge–charge profile of Na-S batteries in ether-based electrolyte (1.0 M NaCF3SO3 in triethylene glycol dimethyl ether (TEGDME)) (adapted with permission from Ref. [43], American Chemical Society 2025). (b) Reaction step equations for RT Na-S batteries. (c) Discharge–charge profile of Na-S batteries in ester-based electrolyte (1 M NaClO4 in ethylene carbonate (EC)/propylene carbonate (PC) with a 1:1 volume ratio and 5% fluoroethylene carbonate (FEC) additive) (adapted with permission from Ref. [44], American Chemical Society 2024). (d) Comparison of binding energies between Na2S6 and ester-based solvent (EC) and ether-based solvent (diethylene glycol dimethyl ether, DEGDME). Pink and blue colors represent negative and positive binding energies, respectively. (adapted with permission from Ref. [45], Elsevier 2024). (e) Key challenges of sulfur cathode for RT Na-S batteries (adapted with permission from Ref. [46], John Wiley & Sons—Books 2022). (f) Schematic diagram illustrating the mechanisms and functions of catalysts (adapted with permission from Ref. [40], John Wiley & Sons—Books 2024).
Batteries 11 00333 g002
Figure 3. Comparison of cycling performance of different TM-based catalysts. Data for TM single/dual atom are from [40,55,56,57,59,60,62,63,64,65,66,67,68,69,70,71,72]; data for TM nanoparticles are from [51,54,75,76,77,78,79,80,81,82]; data for TM compounds are from [83,84,85,86,87,88,89,91,92,93,94,95,96,97]; data for TM heterostructures are from [44,98,99,100,102,103,104,105,106,107,108,109,110].
Figure 3. Comparison of cycling performance of different TM-based catalysts. Data for TM single/dual atom are from [40,55,56,57,59,60,62,63,64,65,66,67,68,69,70,71,72]; data for TM nanoparticles are from [51,54,75,76,77,78,79,80,81,82]; data for TM compounds are from [83,84,85,86,87,88,89,91,92,93,94,95,96,97]; data for TM heterostructures are from [44,98,99,100,102,103,104,105,106,107,108,109,110].
Batteries 11 00333 g003
Figure 4. (a) Formation of embeddings by a natural language processing (NLP) model. (b) A scatter graph of various single-atom candidates in as-built descriptors (adapted with permission from Ref. [74], Springer Nature 2025). (c) High-resolution XPS Mn 2p spectra of Mn-N2/CNs@S. (d) Potentiostatic discharge curves of a Na2S6 solution on Mn-N2/CNs, Mn-N2O2/CNs, and CNs surfaces (adapted with permission from Ref. [70], John Wiley and Sons 2024). (e) Aberration-corrected high-angle annular dark-field scanning transmission electron microscopy (AC HAADF-STEM) image of Zn-N3O/HCs. (f) UV/vis spectra and digital photo of Na2S6 solution before and after being adsorbed by Zn-N3O/HCs@S, Zn-N4/NCs@S, and HCs@S. (g) The -pCOHP curve for Na2S adsorption on Zn-N3O and Zn-N4 (adapted with permission from Ref. [71], John Wiley and Sons 2024). (h) A schematic diagram on d-band center theory and the comparison of Fe-N4S2 and Fe-N4 electronic structure (adapted with permission from Ref. [72], John Wiley & Sons—Books 2023). (i) AC HAADF-STEM magnified image of Fe-Co/NC. (j) CV curves of Fe-Co/NC, Co/NC, and Fe/NC. (k) Gibbs free energy profiles of NaPS species on the surface of Fe-Co/NC, Co/NC, and Fe/NC. (l) Na2S decomposition energy barriers on Fe-Co/NC, Co/NC, and Fe/NC (adapted with permission from Ref. [61], American Chemical Society 2025).
Figure 4. (a) Formation of embeddings by a natural language processing (NLP) model. (b) A scatter graph of various single-atom candidates in as-built descriptors (adapted with permission from Ref. [74], Springer Nature 2025). (c) High-resolution XPS Mn 2p spectra of Mn-N2/CNs@S. (d) Potentiostatic discharge curves of a Na2S6 solution on Mn-N2/CNs, Mn-N2O2/CNs, and CNs surfaces (adapted with permission from Ref. [70], John Wiley and Sons 2024). (e) Aberration-corrected high-angle annular dark-field scanning transmission electron microscopy (AC HAADF-STEM) image of Zn-N3O/HCs. (f) UV/vis spectra and digital photo of Na2S6 solution before and after being adsorbed by Zn-N3O/HCs@S, Zn-N4/NCs@S, and HCs@S. (g) The -pCOHP curve for Na2S adsorption on Zn-N3O and Zn-N4 (adapted with permission from Ref. [71], John Wiley and Sons 2024). (h) A schematic diagram on d-band center theory and the comparison of Fe-N4S2 and Fe-N4 electronic structure (adapted with permission from Ref. [72], John Wiley & Sons—Books 2023). (i) AC HAADF-STEM magnified image of Fe-Co/NC. (j) CV curves of Fe-Co/NC, Co/NC, and Fe/NC. (k) Gibbs free energy profiles of NaPS species on the surface of Fe-Co/NC, Co/NC, and Fe/NC. (l) Na2S decomposition energy barriers on Fe-Co/NC, Co/NC, and Fe/NC (adapted with permission from Ref. [61], American Chemical Society 2025).
Batteries 11 00333 g004
Figure 5. (a) XRD pattern of MG-Co. (b) Cycling performance of MG-Co@S at 0.5C. (c) CV curves and (d) Nyquist plots of symmetric cells with Na2S6 (adapted with permission from Ref. [51], Royal Society of Chemistry 2020). (e) Charge density difference analysis of Co@NC and Co@BNC. (f) Co 2p3/2 XPS spectra of Co@BNC before and after Na2S6 adsorption (adapted with permission from Ref. [81], John Wiley & Sons—Books 2024). (g) AC-HAADF-STEM images of Fe1.88C0.12@CNTs. (h) Ex situ Raman spectra of Fe1.88C0.12@CNTs electrode after discharge to 0.9 V and charge to 2.4 V (adapted with permission from Ref. [75], Elsevier 2025). (i) Ni L-edge X-ray absorption near-edge structure (XANES) spectra of S@FeNi3@HC and FeNi3@HC. (j) Potentiostatic nucleation profiles of Na2S on FeNi3@HC (adapted with permission from Ref. [79], American Chemical Society 2021).
Figure 5. (a) XRD pattern of MG-Co. (b) Cycling performance of MG-Co@S at 0.5C. (c) CV curves and (d) Nyquist plots of symmetric cells with Na2S6 (adapted with permission from Ref. [51], Royal Society of Chemistry 2020). (e) Charge density difference analysis of Co@NC and Co@BNC. (f) Co 2p3/2 XPS spectra of Co@BNC before and after Na2S6 adsorption (adapted with permission from Ref. [81], John Wiley & Sons—Books 2024). (g) AC-HAADF-STEM images of Fe1.88C0.12@CNTs. (h) Ex situ Raman spectra of Fe1.88C0.12@CNTs electrode after discharge to 0.9 V and charge to 2.4 V (adapted with permission from Ref. [75], Elsevier 2025). (i) Ni L-edge X-ray absorption near-edge structure (XANES) spectra of S@FeNi3@HC and FeNi3@HC. (j) Potentiostatic nucleation profiles of Na2S on FeNi3@HC (adapted with permission from Ref. [79], American Chemical Society 2021).
Batteries 11 00333 g005
Figure 6. (a) STEM images of Nb2O5-CNR. (b) Comparison of the binding energies of various Na2Sx compounds on different substrates. The orange and blue bars correspond to the binding energies on the T-Nb2O5 and graphene surfaces, respectively. (c) Bond length variation in the Na2S molecule in different states (adapted with permission from Ref. [92], John Wiley and Sons 2022). (d) Cycling performance of S/YS-Fe2N@NC (blue), S/YS-Fe2O3@C (pink), and S/NC (reddish-brown) at 1C. (e) XPS spectra of S 2p of YS-Fe2N@NC and NC after adsorption of Na2S6 solution. (f) CV profiles of symmetrical batteries. Colored lines represent different electrodes in Na2S6 electrolyte: yellow, YS-Fe2N@NC; blue, YS-Fe2O3@C; olive green, N-doped Carbon. The black line corresponds to the control sample (YS-Fe2N@NC) without Na2S6. (adapted with permission from Ref. [84], Elsevier 2025). (g) High-resolution XPS spectra of O 1s in TiO2@SPC-S and TiO2@SPC. The spectra are deconvoluted into three contributing peaks: Ti-O (blue), C-O-Ti (green), and O-C (purple). (h) Contour plots of in situ UV spectra in different discharge and charge states for TiO2@SPC-S cathode (adapted with permission from Ref. [95], Elsevier 2025). (i) Relative content of Fe valence states. The pie chart shows the corresponding ratio of Fe3+/Fe2+, with the apricot yellow and pink sectors representing the proportions of Fe2+ and Fe3+, respectively. (j) Raman spectra of the suspension solution of A-FeSnOx after Na2S6 adsorption and a pure Na2S6 solution. The spectrum of the adsorbed species (A-FeSnOx-Na2S6) is shown in dark red, while the reference spectrum of pure Na2S6 is shown in black. (adapted with permission from Ref. [122], John Wiley & Sons, Inc. 2024).
Figure 6. (a) STEM images of Nb2O5-CNR. (b) Comparison of the binding energies of various Na2Sx compounds on different substrates. The orange and blue bars correspond to the binding energies on the T-Nb2O5 and graphene surfaces, respectively. (c) Bond length variation in the Na2S molecule in different states (adapted with permission from Ref. [92], John Wiley and Sons 2022). (d) Cycling performance of S/YS-Fe2N@NC (blue), S/YS-Fe2O3@C (pink), and S/NC (reddish-brown) at 1C. (e) XPS spectra of S 2p of YS-Fe2N@NC and NC after adsorption of Na2S6 solution. (f) CV profiles of symmetrical batteries. Colored lines represent different electrodes in Na2S6 electrolyte: yellow, YS-Fe2N@NC; blue, YS-Fe2O3@C; olive green, N-doped Carbon. The black line corresponds to the control sample (YS-Fe2N@NC) without Na2S6. (adapted with permission from Ref. [84], Elsevier 2025). (g) High-resolution XPS spectra of O 1s in TiO2@SPC-S and TiO2@SPC. The spectra are deconvoluted into three contributing peaks: Ti-O (blue), C-O-Ti (green), and O-C (purple). (h) Contour plots of in situ UV spectra in different discharge and charge states for TiO2@SPC-S cathode (adapted with permission from Ref. [95], Elsevier 2025). (i) Relative content of Fe valence states. The pie chart shows the corresponding ratio of Fe3+/Fe2+, with the apricot yellow and pink sectors representing the proportions of Fe2+ and Fe3+, respectively. (j) Raman spectra of the suspension solution of A-FeSnOx after Na2S6 adsorption and a pure Na2S6 solution. The spectrum of the adsorbed species (A-FeSnOx-Na2S6) is shown in dark red, while the reference spectrum of pure Na2S6 is shown in black. (adapted with permission from Ref. [122], John Wiley & Sons, Inc. 2024).
Batteries 11 00333 g006
Figure 7. (a) High-resolution TEM images of TiN, TiO2, and TiN-TiO2 structure. (b) Cycling performance of S/TiN-TiO2@MCCFs and S/MCCFs at 0.1 A g−1. The red and black symbols represent the data points for S/TiN-TiO2@MCCFs and S/MCCFs, respectively. (c) Energy profiles of Na+ migration on the TiN and TiO2 surfaces. (d) Binding energy values of Na2S and Na2S4 adsorbed on the MCCFs, TiN, and TiO2 surfaces (adapted with permission from Ref. [106], American Chemical Society 2021). (e) TEM of S/P-Fe2O3@Fe-PPy. (f) Tafel plots of different sulfur cathodes. The black, red, brown, and green lines correspond to the S/AC, S/P-Fe2O3, S/P-Fe2O3@PPy, and S/P-Fe2O3@Fe-PPy cathodes, respectively. (adapted with permission from Ref. [100], John Wiley & Sons—Books 2023). (g) HRTEM of Ni-B after battery cycle. (h) Na2S nucleation test and SEM for CP-Ni-B. (i) Gibbs free energy curves (adapted with permission from Ref. [107], John Wiley & Sons—Books 2024).
Figure 7. (a) High-resolution TEM images of TiN, TiO2, and TiN-TiO2 structure. (b) Cycling performance of S/TiN-TiO2@MCCFs and S/MCCFs at 0.1 A g−1. The red and black symbols represent the data points for S/TiN-TiO2@MCCFs and S/MCCFs, respectively. (c) Energy profiles of Na+ migration on the TiN and TiO2 surfaces. (d) Binding energy values of Na2S and Na2S4 adsorbed on the MCCFs, TiN, and TiO2 surfaces (adapted with permission from Ref. [106], American Chemical Society 2021). (e) TEM of S/P-Fe2O3@Fe-PPy. (f) Tafel plots of different sulfur cathodes. The black, red, brown, and green lines correspond to the S/AC, S/P-Fe2O3, S/P-Fe2O3@PPy, and S/P-Fe2O3@Fe-PPy cathodes, respectively. (adapted with permission from Ref. [100], John Wiley & Sons—Books 2023). (g) HRTEM of Ni-B after battery cycle. (h) Na2S nucleation test and SEM for CP-Ni-B. (i) Gibbs free energy curves (adapted with permission from Ref. [107], John Wiley & Sons—Books 2024).
Batteries 11 00333 g007
Table 1. Summary of mechanisms and performance of different TM-based catalysts.
Table 1. Summary of mechanisms and performance of different TM-based catalysts.
CategoryCatalytic MaterialsCatalyst MechanismCatalyst ContentElectrolyteCycling Stability
(Sulfur Loading)
Ref.
TM single/dual atomsMn/NCStrong Lewis acid–base interaction promotes the direct conversion of Na2S4 to Na2S\1 M NaFSI in TEGDME with 1wt% NaNO3720 mAh g−1@500th cycle@500 mA g−1 (1.8–2.5 mg cm−2)[55]
Con-HCPolarity–polarity interaction28%1 M NaClO4 in PC/EC (1:1 Vol%) with 5 wt% FEC508 mAh g−1@600th cycle@100 mA g−1[56]
Fe1-NMCElectron transfer17.5%1 M NaClO4 in PC/EC (1:1 Vol%) with 5 wt% FEC540 mAh g−1@500th cycle@0.1 A g−1[57]
Fe/NC/700Metallic bonding enhances electron transfer28%1 M NaClO4 in PC/EC (1:1 Vol%)325 mAh g−1@5000th cycle@10 A g−1 (1.5–2 mg cm−2)[58]
Cu SA/NOC-2Coordination activates S8 cleavage into short-chain sulfur23%1 M NaSO3CF3 in DEGDME776 mAh g−1@100th cycle@0.1 A g−1 (2.0–2.5 mg cm−2)[59]
VCo DACs/N-MXeneElectron spillover effect23%1 M NaSO3CF3 in DEGDME1199.3 mAh g−1@100th cycle@0.2C (0.8–1.2 mg cm−2)[60]
Fe-Co/NCIron atoms induce the delocalization of Co electrons, altering the electron spin state32%1.0 M NaClO4 in EC/DMC (1:1 Vol%) with 5.0% FEC379 mAh g−1@2000th cycle@1C (1 mg cm−2)[61]
Co-N2O2/MOFcStrong d-p hybridization for Co single-atom d-electron density regulation23%1 M NaSO3CF3 in DEGDME425 mAh g−1@1000th cycle@1 A g−1 (1.0–1.2 mg cm−2)[62]
Mn-N-CStrong p-d hybridization32%1 M NaClO4 in PC/EC (1:1 Vol%) with 5 wt% FEC888 mAh g−1@200th cycle@0.2 A g−1 (1.6 mg cm−2)[63]
Fe-N1Unsaturated coordination35%2 M NaTFSI in PC/FEC (1:1 Vol%)647.8 mAh g−1@200th cycle@167.5 mA g−1[64]
Zn-N2/CFAsymmetric electron distribution\1 M NaClO4 in EC/DEC (1:1 Vol%) with 5% FEC317.4 mAh g−1@100th cycle@0.1 A g−1 (1.0 mg cm−2)[40]
Zn-N2@NGUnsaturated coordination14%1 M NaClO4 in EC/DEC (1:1 Vol%) with 5% FEC414 mAh g−1@400th cycle@0.6 A g−1 (1.0 mg cm−2)[65]
Ca-O4N-CAxial-N-ligand-induced Ca site charge localization promotes p-p orbital hybridization23%2 M NaTFSI in PC/FEC (1:1 Vol%)887 mAh g−1@800th cycle@3C (1 mg cm−2)[66]
NHC-InN5 SACsHigh s-p orbital overlap23%1 M NaClO4 in PC/EC (1:1 Vol%) with 5 wt% FEC384.9 mAh g−1@800th cycle@1 A g−1 (1.5 mg cm−2)[67]
Co, N-MPC-10%Polar interaction32%1 M NaClO4 in PC/EC (1:1 Vol%) with 3 wt% FEC1275.01 mAh g−1@100th cycle@0.1C (1.3 mg cm−2)[68]
3D-PNCVPromote electron transfer35%1 M NaTFSI in PC/FEC (1:1 Vol%)991 mAh g−1@100th cycle@0.2 A g−1[69]
Mn-N2/CNsRegulating the coordination number of Mn single atoms enhances d-p orbital hybridization23%2 M NaTFSI in PC/FEC (1:1 Vol%)926 mAh g−1@100th cycle@0.1C (0.9 mg cm−2)[70]
Zn-N3O/HCsN/O coordination to generate localized electrons at the single-atom center of Zn to enhance d-p hybridization23%2 M NaTFSI in PC/FEC (1:1 Vol%)1155 mAh g−1@100th cycle@0.1C (0.8–1.0 mg cm−2)[71]
SA Fe-N/S@CNFRegulation of the coordination structure of Fe-N4S2\1 M NaClO4 in EC/DEC (1:1 Vol%) with 5 wt% FEC595 mAh g−1@500th cycle@1 A g−1 (4 mg cm−2)[72]
Mn1-PNCElectron transfer28%1 M NaClO4 in EC/PC (1:1 Vol%) with 5 wt% FEC344.1 mAh g−1@3000th cycle@2 A g−1 (2.6 mg cm−2)[73]
SA Co-N/SStrong interaction between Co site and sulfur\2 M NaTFSI in PC/FEC (1:1 Vol%)2.54 mAh cm−2@70th@0.5 A g−1 (4.2 mg cm−2)[74]
TM nanoparticlesFe1.88C0.12@CNTsStrong d-p orbital hybridization based on carbon vacancies and iron defects forms Fe-S bonds\1 M NaClO4 in TEGDME with 2% NaNO3486 mAh g−1@1000th cycle@1 A g−1 (0.8–1.4 mg cm−2)[75]
Co-NMCNThe generation of Co-S bonds35%1 M NaClO4 in PC/EC (1:1 Vol%) with 5 wt% FEC529.6 mAh g−1@100th cycle@0.5C (3 mg cm−2)[76]
CoNCThe conversion of Co nanoparticles and CoN4 sites into CoS35%a molar ratio of NaFSI/DME/BTFE = 1:1:0.8530 mAh g−1@150th cycle@80 mA g−1 (0.6–0.7 mg cm−2)[77]
CNTs/Co@NC-0.25Catalysis based on the Co nanoparticles\1 M NaClO4 in EC/DEC (1:1 Vol%) with 2 wt% FEC634.6 mAh g−1@100th cycle@0.1C (0.7–0.9 mg cm−2)[54]
Co@NPCNFsCo nanoparticles enable rapid sodium intercalation and reduction of NaPSs\1 M NaClO4 in EC/DEC (1:1 Vol%)411 mAh g−1@800th cycle@1C (1 mg cm−2)[78]
FeNi3@HCAlloying regulates the electronic structure35%2 M NaClO4 in PC/FEC (1:1 Vol%)862 mAh g−1@100th cycle@0.2 A g−1[79]
Ni/Co-C-12Adjusting the Ni/Co ratio delays the conversion of Na2S to NaPSs and accelerates the conversion of NaPSs to sulfur40%1 M NaClO4 in TEGDME813.5 mAh g−1@200th cycle@0.5C[80]
Co@BNCThe formation of electron defect centers in Co nanoparticles based on B doping31.5%1 M NaClO4 in PC/EC (1:1 Vol%) with 3 wt% FEC416 mAh g−1@600th cycle@0.5C (0.8–1 mg cm−2)[81]
MG-CoHigh-conductivity MXene in synergy with Co nanoparticles for catalysis40%1.0 M NaPF6 in DOL/DIGLYME (1:1 Vol%)360 mAh g−1@200th cycle@0.5C[51]
W@N-GThe reduction of Na2S4 is promoted by embedding W nanoparticles7.28%1 M NaClO4 in EC/PC (1:1 Vol%) with 3 wt% FEC398 mAh g−1@1000th cycle@1C (1 mg cm−2)[82]
TM compoundsNOC@MoS2Strong Lewis acid–base interactions35%1 M NaCF3SO3 in DEGDME576 mAh g−1@100th cycle@0.1 A g−1 (0.9–1.5 mg cm−2)[83]
YS-Fe2N@NCN-doped carbon encapsulating Fe2N to promote the transfer of electrons from NaPSs to N in Fe2N56%1 M NaClO4 in DEGDME467 mAh g−1@350th cycle@2C (3.1 mg cm−2)[84]
MnO@NACMThe conductive network of carbon nanorods works in synergy with MnO for adsorption and catalysis44%1 M NaClO4 in EC/PC (1:1 Vol%) with 5 wt% FEC234 mAh g−1@1000th cycle@2 A g−1 (0.7–0.9 mg cm−2)[85]
NiSe-CRegulating the electronic state of the d/p orbital at the NiSe catalytic site35%2 M NaTFSI in PC/FEC (1:1 Vol%)401.4 mAh g−1@1000th cycle@2 A g−1[86]
HPC/Mo2CStrong adsorption and high catalysis of highly conductive Mo2C38.5%2 M NaTFSI in PC/FEC (1:1 Vol%)1098 mAh g−1@120th cycle@0.2 A g−1[87]
Mo5N6Mo5N6 based on high d bands5%1 M NaClO4 in EC/PC (1:1 Vol%) with 5 wt% FEC186 mAh g−1@10000th cycle@1675 mA g−1 (1.2 mg cm−2)[88]
2D/3D Co4N-NC@CCThe abundant adsorption and catalytic active sites of 2D/3D CoN4 accelerate charge transfer\1 M NaClO4 in EC/PC (1:1 Vol%) with 2M NaNO3592 mAh g−1@1000th cycle@1.0C (1.0 mg cm−2)[89]
Fe3N-NMCNThe Na-N and Fe-S bonds of Fe3N and NaPSs\1 M NaPF6 in DOL/DIGLYME (1:1 Vol%)696 mAh g−1@2800th cycle@8375 mA g−1[90]
NCCSEnhance conductivity50%1 M NaClO4 in EC/DEC (1:1 Vol%)470.3 mAh g−1@500th cycle@1C[91]
Nb2O5-CNRExcellent sodium affinity and sulfur affinity40%1 M NaClO4 in EC/PC (1:1 Vol%) with 5 wt% FEC617 mAh g−1@600th cycle@0.5 A g−1[92]
GeOx/NCThe O vacancy provides highly active sites\1 M NaSO3CF3 in DEGDME385 mAh g−1@1200th cycle@1 A g−1 (1 mg cm−2)[93]
PC/CeO2-xThe O vacancy leads to a reduction in the CeO2-x band gap, increasing the electrical conductivity35%2 M NaTFSI in PC/FEC (1:1 Vol%)906 mAh g−1@200th cycle@0.5 A g−1[94]
TiO2@SPCValence regulation forms defect sites35%2 M NaTFSI in EC/PEC400 mAh g−1@400th cycle@2A g−1[95]
MoTe2Multiple active sites8%\498 mAh g−1@500th cycle@1C[96]
Ti3C2Tx/Ni(OH)2/CSynergistic effect of physical constraints and MXene adsorption as well as Ni(OH)2 catalysis28%1 M NaSO3CF3 in DIGLYME1144.7 mAh g−1@100th cycle@0.22C[97]
TM heterostructuresCo3O4-NC@C3N4Unpaired Co 3d electrons promote electron transfer24%1 M NaClO4 in EC/DEC (1:1 Vol%) with 5 wt% FEC737.2 mAh g−1@1000th cycle@1C (1.5 mg cm−2)[98]
Co1-CoS2/NCCoS2/Co1 dual-terminal electron donor–catalytic synergistic effect28%1 M NaClO4 in EC/PC (1:1 Vol%) with 5 wt% FEC642 mAh g−1@150th cycle@0.2 A g−1 (1.6 mg cm−2)[99]
P-Fe2O3@Fe-PPySeries-segmented catalysis\1 M NaFSI in TEGDME with 1wt% NaNO3674 mAh g−1@500th cycle@500 mA g−1 (1.8–2.5 mg cm−1)[100]
ZnS-NC@Ni-N4Series-segmented catalysis35%1 M NaClO4 in
EC/DEC (1:1 Vol%) with 5 wt% FEC
725 mAh g−1@2000th cycle@1 A g−1 (1.0 mg cm−1)[101]
HCS@Ni-MnO2Heterostructure interface rich in micro/mesoporous hollow structures35%1 M NaCF3SO3 in DEGDME586.8 mAh g−1@1000th cycle@5 A g−1 (1.2–1.4 mg cm−1)[102]
MoC-W2C-MCNFThe MoC-W2C heterointerface accelerates the quasi-solid-state conversion of electron/ion transport\1 M NaClO4 in EC/PC (1:1 Vol%) with 5 wt% FEC640 mAh g−1@500th cycle@0.2 A g−1 (1.5 mg cm−2)[103]
FCNT@Co3C-CoStrong adsorption and catalysis based on Lewis acid sites of Co3C and Co\1 M NaClO4 in EC/DMC (6:4 Vol%)448 mAh g−1@500th cycle@2C (3.2 mg cm−2)[104]
MoN-W2N@PCThe Mo2N-W2N heterojunction35%2 M NaTFSI in PC/FEC (1:1 Vol%)517 mAh g−1@400th cycle@1 A g−1 (0.9 mg cm−2)[105]
TiN-TiO2@MCCFsThe synergistic effect of TiN catalysis and TiO2 adsorption\1 M NaClO4 in EC/PC (1:1 Vol%) with 5 wt% FEC640 mAh g−1@100th cycle@0.1 A g−1 (1.08 mg cm−2)[106]
Ni-B@GOCatalysis of the interface between amorphous Ni-B and crystalline Ni-Sx16%1 M NaTFSI in PC/FEC470 mAh g−1@1000th cycle@2 A g−1[107]
MoS2-Mo1/SGFAtomic-level dual active site delocalization of electrons optimizes the Mo electronic structure7%1 M NaClO4 in EC/PC (1:1 Vol%) with 5 wt% FEC505 mAh g−1@1000th cycle@0.1 A g−1 (0.96 mg cm−2)[108]
MoS2/MoSAC/CFThe regulation of the electronic structure at the center of the Mod band promotes the hybridization of the d-p orbitals35%1 M NaClO4 in EC/PC (1:1 Vol%) with 5 wt% FEC441.38 mAh g−1@100th cycle@1 A g−1 (1.0 mg cm−2)[44]
Co-S-C@MCThe heterogeneous interface Co-S-C regulates the electronic structure and enhances the center of the Co d band32%1 M NaTFSI in TEGDME/FEC (1:1 Vol%)1215 mAh g−1@500th cycle@0.1C (1.0 mg cm−2)[109]
CoS2-CoSe2@CNFsStrong adsorption of CoS2 and high catalysis of CoSe2\2 M NaTFSI in PC/FEC (1:1 Vol%)749 mAh g−1@200th cycle@1 A g−1[110]
The “@” symbol denotes “at”. For example, “720 mAh g−1@500th cycle@500 mA g−1” means a capacity of 720 mAh g−1 at the 500th cycle under a current density of 500 mA 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, J.; Wang, Y.; Yang, Y.; Lei, P.; Cao, H.; Xiang, Y. Transition Metal-Based Catalysts Powering Practical Room-Temperature Na-S Batteries: From Advances to Further Perspectives. Batteries 2025, 11, 333. https://doi.org/10.3390/batteries11090333

AMA Style

Li J, Wang Y, Yang Y, Lei P, Cao H, Xiang Y. Transition Metal-Based Catalysts Powering Practical Room-Temperature Na-S Batteries: From Advances to Further Perspectives. Batteries. 2025; 11(9):333. https://doi.org/10.3390/batteries11090333

Chicago/Turabian Style

Li, Junsheng, Yongli Wang, Yuanyuan Yang, Peng Lei, Huatang Cao, and Yinyu Xiang. 2025. "Transition Metal-Based Catalysts Powering Practical Room-Temperature Na-S Batteries: From Advances to Further Perspectives" Batteries 11, no. 9: 333. https://doi.org/10.3390/batteries11090333

APA Style

Li, J., Wang, Y., Yang, Y., Lei, P., Cao, H., & Xiang, Y. (2025). Transition Metal-Based Catalysts Powering Practical Room-Temperature Na-S Batteries: From Advances to Further Perspectives. Batteries, 11(9), 333. https://doi.org/10.3390/batteries11090333

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