Next Article in Journal
Biosynthesis of Silver Nanoparticles Using Phytochemicals Extracted from Aqueous Clerodendrum glabrum for Anti-Diabetes and Anti-Inflammatory Activity: An In Vitro Study
Previous Article in Journal
Electromagnetic Interference in the Modern Era: Concerns, Trends, and Nanomaterial-Based Solutions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Engineering 3D Heterostructured NiCo2S4/Co9S8-CNFs via Electrospinning and Hydrothermal Strategies for Efficient Bifunctional Energy Conversion

Department of Chemistry, College of Natural Sciences, Yeungnam University, 280 Daehak-Ro, Gyeongsan 38541, Gyeongbuk, Republic of Korea
*
Authors to whom correspondence should be addressed.
Nanomaterials 2025, 15(20), 1559; https://doi.org/10.3390/nano15201559
Submission received: 25 September 2025 / Revised: 8 October 2025 / Accepted: 11 October 2025 / Published: 13 October 2025
(This article belongs to the Special Issue Design and Application of Nanomaterials in Photoenergy Conversions)

Abstract

The rational design of multifunctional electrocatalysts requires synergistic integration of conductive scaffolds with redox-active components. Here, a hierarchical core–shell NiCo2S4 grown/anchored on Co9S8-loaded carbon nanofibers (NCS/CS/CNFs) was synthesized via an electrospinning and hydrothermal approach and systematically characterized. FESEM/TEM confirmed a core-shell nanofiber structure with a NiCo2S4 shell thickness of ~30–70 nm, increasing the fiber diameter to ~290 ± 30 nm, while BET analysis revealed a surface area of 24.84 m2 g−1 and pore volume of 0.042 cm3 g−1, surpassing CS/CNFs (6.12 m2 g−1) and NCS (4.85 m2 g−1). XRD confirmed crystalline NiCo2S4 and Co9S8 phases, while XPS identified mixed Ni2+/Ni3+ and Co2+/Co3+ states with strong Ni-S/Co-S bonding, indicating enhanced electron delocalization. Electrochemical measurements in 1 M KOH demonstrated outstanding OER activity, with NCS/CS/CNFs requiring only 324 mV overpotential at 10 mA cm−2, a Tafel slope of 125.7 mV dec−1, and low charge-transfer resistance (0.33 Ω cm2). They also achieved a high areal capacitance of 1412.5 μF cm−2 and maintained a stable current density for >5 h. For methanol oxidation, the composite delivered 150 mA cm−2 at 0.1 M methanol, ~1.6 times that of CS and 1.3 times that of NCS, while maintaining stability for 18,000 s. This bifunctional activity underscores the synergy between conductive CNFs and hierarchical sulfides, offering a scalable route to durable electrocatalysts for water splitting and direct methanol fuel cells.

Graphical Abstract

1. Introduction

The increasing demand for sustainable energy technologies is driven by the urgent need to mitigate climate change, environmental degradation, and the depletion of fossil resources. Renewable energy sources such as solar and wind offer promising alternatives; however, their intermittency necessitates the development of efficient and scalable energy conversion and storage systems to stabilize the power supply [1,2,3,4,5]. Electrochemical devices, including water electrolyzers, metal-air batteries, and fuel cells, are desirable because they enable the direct and efficient interconversion of chemical and electrical energy with high efficiency [6]. Among these processes, the oxygen evolution reaction (OER) plays a vital role in water electrolysis and rechargeable metal-air batteries. However, it is hampered by sluggish four-electron kinetics and substantial overpotentials. In addition, the methanol oxidation reaction (MOR), central to direct methanol fuel cells (DMFCs), suffers from low catalytic efficiency, poisoning by carbonaceous intermediates, and poor long-term stability [7,8,9,10,11,12,13,14,15,16,17,18]. Noble metal catalysts, such as IrO2 and RuO2 (for OER), and Pt-based materials (for MOR) remain benchmarks; however, their scarcity, high cost, and limited durability restrict their large-scale application [19,20,21]. Hence, the development of low-cost, durable, and bifunctional electrocatalysts for both OER and MOR is of critical importance. Transition metal sulfides (TMSs), particularly Ni-Co sulfides, have emerged as attractive alternatives owing to their rich redox chemistry, tunable electronic structure, and relatively low cost [22,23,24]. Despite these advantages, conventional TMS catalysts still face limitations, including moderate conductivity, insufficient structural stability, and restricted bifunctional activity in alkaline environments [24,25].
Recent studies have shown that the catalytic performance of Ni/Co sulfides is strongly dependent on their morphology and structural configuration. For example, Co3O4@Co9S8 heterostructures grown on Ni foam achieved an OER overpotential of ~80 mV at 10 mA cm−2 with a Tafel slope of 107.2 mV dec−1, attributed to efficient charge transfer across the oxide-sulfide interface [26]. Sulfur-vacancy-engineered Co9S8 core-shell hollow spheres displayed dual activity, with an OER overpotential of 294 mV at 10 mA cm−2 and a MOR current density of 164.9 mA cm−2 at 1.8 V vs. RHE, enabled by their high surface area and optimized electronic states [27]. Similarly, core-shell Au@NiCo2S4 nanoparticles reduced the OER overpotential to 290 mV at 10 mA cm−2, with a Tafel slope of 44.5 mV dec−1, benefiting from the conductive Au core that promoted the formation of high-valence Ni/Co species [28]. Three-dimensional mushroom-like NiCo/NiCo2S4@NiCo arrays on Ni foam enabled overall water splitting at a cell voltage of 1.55 V and a current density of 10 mA cm−2, owing to their hierarchical architecture with abundant exposed edges [29]. Carbon-supported hybrid systems also play a crucial role in enhancing activity and durability [30]. NiCo2S4@graphene hybrids showed bifunctional activity, achieving an OER overpotential of ~350 mV and improved oxygen reduction reaction (ORR) currents. The graphene scaffold enhanced the conductivity and dispersion of active sites [31]. Similarly, Co9S8 nanoparticles embedded in N/S dual-doped hollow carbon nanofibers showed excellent durability and improved performance in terms of both ORR and OER [32]. More recently, bi-phase NiCo2S4-NiS2 deposited on carbon fiber paper achieved an OER overpotential of 165 mV at 10 mA cm−2 with a Tafel slope of 81.54 mV. dec−1, demonstrating the synergistic role of spinel and pyrite phases [33]. These examples underscore the importance of rational morphology control, heterostructure engineering, and conductive supports in optimizing bifunctional activity [30].
One promising strategy involves the construction of amorphous crystalline heterostructures, which leverage interfacial interaction, create defect-rich active sites, and improve charge transport [34,35,36]. In particular, embedding amorphous Co9S8 within conductive carbon nanofibers (CNFs) and coating with crystalline NiCo2S4 spinels provides a highly effective design. The CNF framework offers high conductivity and porosity [29], the amorphous Co9S8 introduces abundant defect-rich active sites [31], and the crystalline NiCo2S4 shell contributes enhance redox activity through multiple oxidation states [29,32]. The resulting amorphous/crystalline interfaces promote charge redistribution and accelerate catalytic kinetics.
Herein, we report the synthesis of a 3D NiCo2S4/Co9S8-CNF (NCS/CS/CNF) heterostructure via electrospinning and hydrothermal methods. This architecture integrates a conductive CNF network with a defect-enriched amorphous Co9S8 core and a redox-active crystalline NiCo2S4 shell. The synergistic interactions at the amorphous/crystalline interface accelerate charge transfer, enhance catalytic kinetics, and ensure long-term durability. Electrochemical characterization reveals that this heterostructure exhibits outstanding bifunctional activity in relation to both OER and MOR, along with excellent stability. Beyond introducing novel heterostructured materials, this study also establishes a generalizable design principle for hierarchical electrocatalysts in next-generation energy conversion systems.

2. Materials and Methods

A combination of synthesis, structural, and electrochemical methods was employed to design and evaluate Ni-Co sulfide-based carbon nanofiber composites. These approaches were selected to systematically investigate the morphology, phase composition, surface chemistry, and catalytic activity of the materials, thereby providing a comprehensive understanding of their performance in OER and MOR reactions.

2.1. Chemicals

All reagents were of analytical grade and used without further purification. Polyacrylonitrile (PAN, Mw = 90,000) and N,N-dimethylformamide (DMF) were used as the polymer matrix and solvent, respectively. Metal precursors, including cobalt acetate tetrahydrate (C4H14CoO8), nickel nitrate hexahydrate (Ni(NO3)2.6H2O; ≥98.0%), and cobalt nitrate hexahydrate (Co(NO3)2.6H2O; ≥98.0%), were used. Additional chemicals, such as thiourea (NH2CSNH2; ≥99.0%), ammonium fluoride (NH4F; ≥98.0%), potassium hydroxide (KOH; ≥85%), methanol (CH3OH; ≥99.9%), hydrochloric acid (HCl), and Nafion solution, were also utilized. Nickel foam was employed as the current collector. All chemicals were obtained from Merck (Darmstadt, Germany), Sigma-Aldrich (St. Louis, MO, USA), and the MTI Corporation (Richmond, CA, USA). Deionized water was used as the solvent medium throughout the experiments. Materials such as 20 mL syringes and aluminum foil were used during the preparation process.

2.2. Synthesis of CS/CNFs, NCS, and NCS/CS/CNFs

NCS/CS/CNF heterostructures were synthesized through a combined electrospinning, carbonization, and hydrothermal growth process (Figure 1). For the preparation of Co-loaded nanofibers, 2.0 g of PAN was dissolved in 20 mL of DMF under stirring for 6 h at room temperature, followed by the addition of 0.747 g of cobalt acetate tetrahydrate and continued stirring for 10 h. The resulting pink solution was loaded into a 20 mL syringe and electrospun at a voltage of 15 kV, a flow rate of 0.4 mL/h, and a tip-to-collector distance of 15 cm, using aluminum foil as the collector. The collected nanofibers were stabilized in air at 220 °C for 5 h and then carbonized as well as sulfurized at 700 °C for 3 h in a nitrogen atmosphere, using sulfur powder as the sulfur source. This treatment yielded Co9S8 nanoparticles embedded in a conductive carbon nanofiber (CS/CNF) matrix.
To fabricate NCS/CS/CNFs, 0.3 g of CS/CNFs was dispersed in 50 mL of deionized water. A 1:2 molar ratio of nickel nitrate hexahydrate (1 mM) and cobalt nitrate hexahydrate (2 mM), along with 4 mM of thiourea and 4 mM of NH4F, was added subsequently. After uniform dispersion, the mixture was sealed in a Teflon-lined autoclave and heated at 150 °C for 8 h. The product was collected by centrifugation, washed six times with deionized water and ethanol, and dried at 80 °C overnight. For comparison, pure NiCo2S4 (NCS) was prepared under identical hydrothermal conditions without the addition of CS/CNFs. During carbonization/sulfurization, cobalt reacts with sulfur vapor to form Co9S8, while PAN is converted into conductive carbon nanofibers. Subsequent hydrothermal treatment facilitates the nucleation and growth of NiCo2S4 on the CS/CNF scaffold, resulting in the formation of the core-shell heterostructure. The synthetic transformations can be summarized as follows:
( C 3 H 3 N ) n   +   Co ( CH 3 COO ) 2 . 4 H 2 O D M F ,       s t i r r i n g   PAN-Co   solution
PAN-Co   sol . E l e c t r o s p i n n i n g   ( 15   k V ,       0.4   m L   h 1 )   Co-loaded   PAN   nanofibers
Co-PAN   nanofibers 220   ° C ,       a i r   Oxidized   Co   species   +   stabilized   PAN
Stabilized   PAN-Co + S   ( solid ) 700   ° C ,       N 2 ,       3 h   Co 9 S 8 / CNF
Co 9   S 8 / CNF + N i 2 + + 2   C o 2 + + 4   S 2 H y d r o t h e r m a l ,       150   ° C ,       8 h   NiCo 2 S 4 / Co 9 S 8 / CNF

2.3. Material Characterization

The morphology and microstructure of the samples were examined by field emission scanning electron microscopy (FE-SEM, Hitachi S-4800) and transmission electron microscopy (TEM, Tecnai G2 F20ST Win). Elemental mapping was performed by energy-dispersive X-ray spectroscopy (EDS) coupled with SEM and TEM. X-ray photoelectron spectroscopy (XPS; Thermo Scientific, Waltham, MA, USA, Al Kα radiation, 1486.6 eV) was performed under high-vacuum conditions (5 × 10−9 Torr) to analyze the surface chemical states. Crystallographic phases were identified by powder X-ray diffraction (XRD; MPD 3 kW, PANalytical) using Cu Kα radiation (λ = 1.5406 Å) over a 2θ range of 10 to 80° with a step size of 0.02°. Textural properties, including surface area and pore-size distribution, were determined from N2 adsorption–desorption isotherms using the Brunauer–Emmett–Teller (BET) method (3-Flex physisorption analyzer, Micromeritics, Norcross, GA, USA).
  • Electrochemical measurements
Electrochemical tests were conducted to evaluate the activity, kinetics, and durability of the catalysts in both OER and MOR reactions. All measurements were performed using a Corr Test CS electrochemical workstation in a conventional three-electrode setup. Catalyst-coated nickel foam (1.0 × 0.5 cm2) was used as the working electrode, a platinum mesh (1.0 × 1.0 cm2) as the counter electrode, and a Hg/HgO electrode as the reference electrode. All experiments were conducted in 1.0 M KOH electrolyte at room temperature. Catalyst inks were prepared by dispersing 10 mg of sample (CS/CNFs, NCS, or NCS/CS/CNFs) in a 1 mL ethanol/deionized water mixture (3:7 v/v) containing 30 μL of 5 wt% Nafion solution, followed by ultrasonication for 1 h. The suspension was drop-cast onto Ni-foam substrates, and no electrochemical pre-activation was applied.
Potentials were converted to the reversible hydrogen electrode (RHE) scale using the following [37]:
ERHE = EHg/HgO + 0.924 V
OER activity was measured by linear sweep voltammetry (LSV) at a scan rate of 2 mV/s. Ohmic drop correction was applied according to [37,38]:
Ecorr = Emeasured − iR
The overpotential (η) was calculated as follows [13]:
η = ERHE − 1.23 V
Tafel slopes were obtained from LSV data according to [39]:
η = a + b log (j)
where j is the current density and a and b are fitting constants.
Electrochemical impedance spectroscopy (EIS) was conducted over the frequency range of 0.01 Hz to 0.1 MHz to assess charge-transfer resistance. Long-term OER durability was evaluated by chronoamperometry (CA) at 0.65 V (vs. Hg/HgO) for 5 h. In addition, the MOR activity of the samples was studied in 1.0 M KOH with varying concentrations of methanol (0.1 to 1.5 M). A combination of CV, LSV, and CA techniques was employed to comprehensively measure the catalytic performance of the samples toward methanol oxidation.

3. Results and Discussion

3.1. Surface Morphology and Elemental Analysis

The surface morphology of the synthesized nanostructures was systematically examined by FESEM, and the results are shown in Figure 2. At low magnification (Figure 2a), the CS/CNF composite exhibits a uniform, three-dimensional nanofibrous network with highly interconnected fibers and abundant void spaces. Such porous fibrous architectures are known to promote rapid ion/electron transport and mechanical robustness in electrochemical systems [40]. At medium magnification (Figure 2b), the nanofiber surfaces are uniformly decorated with fine nanoparticles (~30 nm), indicating successful incorporation/embedding of cobalt sulfide species. The conformal and continuous coating of nanoparticles along the CNF backbone (Figure 2c) provides intimate interfacial contact. It creates a robust platform for further nucleation of active materials, in agreement with previous reports on CNF-supported sulfide systems [41,42]. The morphology of NCS was synthesized without the use of a carbon nanofiber scaffold, as displayed in Figure 2d–f. At a lower magnification (Figure 2d), the NCS sample exhibits aggregated microspheres with rough, flower-like surfaces. Each microsphere consists of densely packed nanosheets, forming a hierarchical, petal-like architecture (Figure 2e). Such ultrathin nanosheets expose a large number of electroactive sites, which are favorable for redox activity (Figure 2f). However, the random aggregation of microspheres may hinder ion diffusion and compromise mechanical stability during extended cycling [43,44]. Figure 2g–l presents the FESEM images of the hierarchical NCS/CS/CNF hybrid composite from two representative regions. At low magnification (Figure 2g), the fibrous morphology of the CNF scaffold remains intact following hydrothermal growth, confirming its role as a mechanically robust host. A closer inspection (Figure 2h) reveals that NiCo2S4 nanostructures are uniformly anchored onto the CS/CNFs, forming a distinct core-shell configuration. The nanosheet-based shell has an estimated thickness of ~70 ± 5 nm (Supplementary Figure S1a), which increases the overall fiber diameter to ~290 nm ± 30 nm (Figure S1b). Such core–shell architectures are known to enhance electron transport, improve active surface exposure, and accelerate electrolyte diffusion [45,46,47]. High-magnification imaging (Figure 2i) reveals rough and porous shell features, further improving ion accessibility.
The second region (Figure 2i,j) confirms a consistent morphology, where NiCo2S4 nanosheets retain their flower-like features but are now firmly anchored and evenly dispersed in the composite. This uniform anchoring prevents aggregation, enhances mechanical stability, and establishes conductive pathways along the CNF backbone. Such uniform anchoring of hierarchical sulfides onto conductive supports has been widely recognized as an effective strategy for enhancing electrochemical efficiency [48,49,50].
The elemental distribution of the NCS/CS/CNF hybrid was further examined using energy-dispersive X-ray spectroscopy (EDS) (Figure S2). The elemental maps reveal a homogeneous distribution of Ni, Co, S, and C throughout the nanofibrous framework, with no impurities detected. Quantitative EDS analysis (Figure S2, spectrum, and table) yielded atomic ratios of Ni (11.51 at.%), Co (19.92 at.%), S (30.48 at.%), and C (29.46 at.%), which are consistent with the intended stoichiometry of the NiCo2S4 phase and the carbon scaffold. The uniform elemental dispersion agrees well with prior studies and supports the robust structural and compositional integrity of the NCS/CS/CNF hybrid [51,52].
To clarify the core-shell architecture, high-resolution transmission electron microscopy (HRTEM) and high-angle annular dark-field (HAADF) imaging were conducted (Figure 3). The TEM images (Figure 3a–c) clearly show a nanofiber core with a diameter of 140 nm, enveloped by a conformal NiCo2S4 shell 30 nm thick, which aligns well with the FESEM measurement. This type of NiCo2S4 core-shell nanostructure has been demonstrated in similar energy storage and conversion systems [28,53,54,55]. Slight variations in shell thickness are attributed to localized differences in nucleation and growth during the hydrothermal process; these are also evident in Figure 3a,b. The HRTEM image (Figure 3c) lacks visible lattice fringes, suggesting that the NiCo2S4 shell is amorphous or poorly crystalline in this region. Such mixed-phase (amorphous/crystalline) structures are often advantageous in electrochemical contexts, providing abundant active sites, enabling improved ion transport, and offering outstanding structural stability [56,57,58]. HAADF imaging, along with EDS elemental mapping (Figure 3d), further confirms the core–shell morphology. The analysis reveals that the shell primarily concentrates Ni, Co, and S, while the fiber core mainly confines carbon. The distribution appears more even, extending across both regions and suggesting either partial interdiffusion or uniform surface coverage. This imaging pattern is reinforced by the line-scan elemental profile (Figure S3), where the intensities of Ni, Co, and S rise sharply at the fiber edges, contrasting with the uniform distribution of carbon throughout the cross-section. Collectively, these findings validate the successful formation of the NCS/CS/CNF core-shell structure. This design leverages the high electrical conductivity of the carbon core together with the redox-rich sulfide shell, offering a synergistic strategy for enhanced electrochemical performance [59,60].

3.2. Structural and Surface Chemistry Analysis

The structural and surface chemistry characteristics of the CS/CNF, NCS, and NCS/CS/CNF composites were systematically analyzed using XRD and XPS, as presented in Figure 4. The XRD pattern of CS/CNFs (Figure 4a) shows a broad hump without any distinct diffraction peaks, confirming its amorphous nature. This feature is typical of carbon nanofiber-based frameworks that lack long-range crystallinity [61]. Importantly, no crystalline reflections corresponding to cobalt sulfide were detected, suggesting that cobalt species exist either in disordered domains or as highly dispersed nanoclusters embedded in the CNF framework. This assumption was further supported by elemental mapping and EDS analysis (Figure S4), which revealed the uniform distribution of C, S, and Co throughout the fibrous network. Quantitative EDS confirmed the incorporation of 46.73 wt% C, 16.48 wt% S, and 13.37 wt% Co, thereby validating the successful doping of cobalt and sulfur into the CNF structure despite the absence of crystalline phases in XRD.
In contrast, the NCS sample displays well-defined peaks at 2θ values of 26.8°, 31.5°, 38.1°, 50.2°, and 55.3°, which are indexed to the (220), (311), (222), (400), (422), (511), and (440) planes of the spinel NiCo2S4 phase (JCPDS No. 00-043-1477) [62,63]. Additional reflections corresponding to Co9S8 (JCPDS No. 01-073-1442) and Co1-xS (JCPDS No. 00-042-0826) can also be observed, suggesting the coexistence of secondary cobalt sulfide phases, possibly due to local stoichiometric inhomogeneity or kinetically limited nucleation during synthesis [64,65,66]. The NCS/CS/CNF composite exhibits a hybrid XRD profile, where the crystalline reflections of NiCo2S4 are clearly retained alongside the broad amorphous hump of CNFs. This coexistence of crystalline NiCo2S4 with an amorphous CNF framework demonstrates the successful integration of active sulfide nanostructures into a conductive carbon scaffold. Such hybridization is advantageous because it combines the fast electron transport pathways of CNFs with the multi-redox activity of sulfides, thereby promoting synergistic electrochemical behavior effects [67].
The XPS survey spectra (Figure 4b) further corroborate these findings, showing characteristic signals for C 1s, O 1s, N 1s, S 2p, Ni 2p, and Co 2p in both NCS and NCS/CS/CNF samples. In contrast, CS/CNFs lack Ni-related peaks, consistent with the absence of nickel precursors. The relative intensity of C 1s and O 1s peaks in CS/CNFs and NCS/CS/CNFs suggests the presence of abundant oxygenated functional groups on CNFs, which may facilitate strong interfacial bonding with transition metal sulfides. High-resolution C 1s spectra (Figure 4c) deconvolute into three main components: (i) a peak at 284.6–284.7 eV (C-C/C=C, sp2-hybridized graphitic carbon), (ii) a peak at 258.8–286.2 eV (C-O/C-S bonds; hydroxyl, ether, or thiol groups), and (iii) a peak at 288.4–288.8 eV (C=O or O-C=O, carbonyl or carboxyl functionalities) [68,69,70]. Among the samples, NCS/CS/CNFs exhibit the strongest C-O/C-S and C=O features, suggesting enhanced surface oxidation and strong interfacial coupling between sulfides and CNFs, which are essential for enhancing charge transfer efficiency.
The Ni 2p spectra (Figure 4d) provide detailed information on the oxidation states and chemical environment of nickel in the composite systems. For the NCS sample, the Ni 2p3/2 and Ni 2p1/2 peaks at 855.4 eV and 873.6 eV correspond to Ni2+ species in sulfide phases, such as the NiCo2S4 phase [71,72]. Additional peaks at 858.3 eV and 877.9 eV can be attributed to Ni3+ species, indicating a mixed-valence Ni2+/Ni3+ couple. The accompanying satellite peaks at 861.71 eV and 881.1 eV arise from shake-up transitions of high-spin Ni2+. Remarkably, in the NCS/CS/CNF composite, a chemical shift toward lower binding energies can be observed, with peaks at 853.1 eV (2p3/2) and 870.4 eV (2p1/2) assigned to Ni-S covalent bonds [73]. This shift indicates stronger Ni-S interactions and enhanced electronic coupling with CNFs. Additional contributions at 856.3 eV, 857.9 eV, 874.5 eV, and 877.5 eV, along with satellite peaks at 861.9 eV and 880.9 eV, suggest the continued coexistence of Ni2+/Ni3+ states. These results suggest that electron transfer between NiCo2S4 and the conductive CS/CNF matrix results in electronic modulation, thereby enhancing the delocalization of charge and stabilizing sulfide species within the CNFs.
The Co 2p spectra (Figure 4e) reveal distinct differences in the chemical states of the cobalt among all samples. In the CS/CNF sample, two prominent peaks can be observed at 781.4 eV (Co 2p3/2) and 797.7 eV (Co 2p1/2), along with satellite features at 784.6, 787.6, 802, and 804.9 eV [73]. These binding energies are characteristic of Co2+ in Co9S8, with strong shake-up satellites resulting from high-spin Co2+ in an octahedral Co-S coordination environment [74]. The absence of Co0 or Co3+ contributions indicates that cobalt exists almost exclusively in the +2 oxidation state in CS/CNFs, consistent with a stable Co9S8 phase embedded within the CNF matrix [75]. In contrast, the NCS sample exhibits a more complex Co 2p spectrum, with Co 2p3/2 peaks at 777.9, 780.9, and 782.9 and Co 2p1/2 peaks at 796.4, 798.2, and 802.4 eV [73,76]. The presence of multiple components and satellite features at ~786.5 eV confirms the existence of Co2+ and Co3+ oxidation states, a characteristic known to be present in spinel-structured NiCo2S4. These findings are in close agreement with previous studies, which have shown that Co2+ and Co3+ coexist due to the mixed-valence occupation of tetrahedral and octahedral Co sites in the spinel lattice, contributing to enhanced redox activity. For the NCS/CS/CNF composite, the Co 2p3/2 spectrum includes peaks at 778.6, 780.9, and 782.7 eV, while the Co 2p1/2 region displays peaks at 793.6, 797.7, 801.1, and 803.7 eV. The low-binding energy feature at 778.6 is attributed to metallic Co0, suggesting the presence of partially reduced cobalt species, likely due to interfacial electronic interactions with the conductive CNF framework. The simultaneous presence of Co0, Co2+, and Co3+ in NCS/CS/CNFs indicates a broader distribution of cobalt oxidation states than in either CS/CNFs or NCS alone. This expanded redox window can be ascribed to the synergistic effect of phase integration and the conductive CNF matrix, which may facilitate electron delocalization and structural flexibility. Moreover, the broadened and intensified satellite peaks suggest strong Co-S covalence and charge transfer interactions, contributing to the improved electrochemical performance observed in the composite system.
The S 2p high-resolution XPS spectra (Figure 4f) were deconvoluted into S 2p3/2 and S 2p1/2 spin-orbit doublets with a consistent energy separation of ~1.18 eV and a fixed 2:1 area ratio, as expected for sulfur species. In the CS/CNF sample, two peaks centered at 163.65 eV (S 2p3/2) and 164.83 eV (S 2p1/2) are attributed to disulfide (S22−) or C-S bonding, which are typically observed in carbon-supported cobalt sulfides [77,78,79]. Peaks at 168.35 (S 2p3/2) and 169.53 eV (S 2p1/2) correspond to oxidized sulfur species (e.g., SOx and SO42) formed upon exposure to air. In the NCS sample, the prominent peaks at 161.35 eV (S 2p3/2) and 162.53 eV (S 2p1/2) are attributed to S2− species in metal-sulfide bonds, specifically Ni-S and Co-S coordination within the NiCo2S4 lattice. Additionally, distinct oxidized sulfur peaks can be observed at 168.53 eV (S 2p3/2) and 169.71 eV (S 2p1/2), confirming the coexistence of reduced and oxidized sulfur states in NCS. For the NCS/CS/CNF composite, two well-defined doublets appear at 162.03/163.21 eV (S 2p3/2/S 2p1/2) and 168.66/169.84 eV, representing reduced and oxidized sulfur, respectively. The relative proportion of oxidized sulfur (~71%) indicates enhanced surface oxidation and charge redistribution at the NCS-carbon interface, consistent with strong interfacial coupling between NiCo2S4, Co9S8, and the CNF matrix [53,77,80,81,82]. Overall, the XPS results clearly demonstrate that all samples contain both reduced and oxidized sulfur species, with the oxidized fraction increasing systematically from NCS < NCS/CS/CNFs < CS/CNFs. These XPS findings, in conjunction with XRD analysis, confirm that the CS/CNF sample consists of amorphous carbon nanofiber with oxidized Co species and covalently bonded sulfur. Also, the NCS and NCS/CS/CNF samples exhibit crystalline NiCo2S4 and Co9S8 phases featuring mixed-valence sulfur environments and strong electronic interactions at the sulfide-carbon interface.
The high-resolution O 1s XPS spectra of the CS/CNF, NCS, and NCS/CS/CNF samples (Figure S5) were deconvoluted into three distinct components, denoted as OI, OII, and OIII, corresponding to different oxygen chemical states. The first peak (OI, ≈530.1–530.3 eV) originates from lattice oxygen (O2−) bound to metal cations such as Ni or Co in metal-oxygen or oxy-sulfide bonds [67,72]. The second peak (OII, ≈531.4–531.7 eV) is assigned to oxygen in surface hydroxyl groups (-OH) or oxygen-vacancy-related species, indicating structural defects and unsaturated coordination sites that enhance electronic conductivity and redox activity [72,75]. The third component (OIII, ≈532.6–533.1 eV) corresponds to adsorbed oxygen species or oxygenated surface groups (C-O, C=O, or H2O) arising from partial surface oxidation of carbon or metal sulfides [75,77]. Among the samples, CS/CNFs exhibit the strongest OIII contribution, reflecting the abundance of oxygen-containing functional groups on the CNF surfaces [75,77]. NCS exhibits a dominant OI peak, consistent with lattice oxygen within the NiCo2S4 framework [67,72]. In contrast, the NCS/CS/CNF composite exhibits enhanced OI and OII features, indicating the presence of lattice oxygen and surface hydroxyl/defect content at the NCS-carbon interface. These findings reveal that the hybrid structure possesses more chemically diverse and defect-rich oxygen species, which facilitate charge transfer and ion transport. This oxygen-rich and defect-engineered surface promotes enhanced OH adsorption and lattice oxygen participation during redox reactions, thereby improving the electrochemical performance of the NCS/CS/CNF composite [75,83].
Figure S6 presents the nitrogen (N2) adsorption-desorption measurements and pore-size distributions of the CS/CNF, NCS, and NCS/CS/CNF samples. Figure S6a–c display the adsorption-desorption isotherms recorded at 77 K. The progressive increase in adsorption and the more pronounced hysteresis observed for NCS/CS/CNFs indicate the formation of a more open mesoporous framework with a larger accessible pore volume compared with the CS/CNF and NCS samples. The BET surface areas of the samples were measured as 6.12 m2/g for CS/CNFs, 4.85 m2/g for NCS, and 24.84 m2/g for NCS/CS/CNFs. Figure S6d–f illustrate the BJH pore-size distributions derived from the desorption branches. CS/CNFs exhibit a narrow mesopore distribution centered at 6.50 nm, with a pore volume of 0.0080 cm3/g. NCS displays a broader distribution centered at 14.5 nm, with a pore volume of 0.0173 cm3/g. In contrast, NCS/CS/CNFs demonstrate a dominant mesopore centered at 8.29 nm, along with the highest pore volume (0.0424 cm3/g). Overall, the NCS/CS/CNF composite provides a hierarchical mesoporous architecture that enhances the electrolyte-accessible surface area and pore volume, which is expected to improve ion transport and electrochemical accessibility.

3.3. Electrocatalytic Activity for OER

The electrochemical behavior of CS/CNF, NCS, and NCS/CS/CNF electrodes was systematically examined in a N2-saturated 1 M KOH electrolyte using CV, LSV, Tafel analysis, EIS, and CA. As shown in the CV curves at 20 mV/s (Figure 5a), the CS/CNF electrode delivered only weak current responses due to the low conductivity of Co9S8 and the electrochemical inertness of CNFs. In contrast, NCS exhibited well-defined redox peaks with significantly higher current density, confirming enhanced Faradaic activity from the Ni/Co redox centers and the conductive sulfide framework. Remarkably, the NCS/CS/CNF hybrid electrode achieved the largest CV integral area with pronounced redox features, reflecting high reversibility and efficient utilization of active sites. This enhancement arises from the synergistic contributions of the conductive CNFs, electroactive NCS, and the structural stability imparted by the CS/CNF framework [84,85].
The LSV profiles (Figure 5b) further highlight the catalytic differences. CS/CNFs required a high overpotential of 382 mV to achieve a current density of 10 mA cm−2, while NCS reduced the overpotential to 355 mV due to its abundant redox-active sites. Notably, the NCS/CS/CNF hybrid achieved the lowest overpotential (324 mV) with the highest current density, indicating accelerated charge-transfer kinetics and enhanced catalytic efficiency. Consistent with these results, Tafel slope (Figure 5c) analysis revealed sluggish kinetics for CS/CNFs (105.2 mV dec−1), moderate improvement for NCS (149.52 mV dec−1), and the most favorable slope for NCS/CS/CNFs (125.8 mV dec−1), confirming efficient electron transport in the hybrid structure [86,87].
The electrochemically active surface area (ECSA) was estimated from cyclic voltammetry (CV) performed in the non-Faradaic region (0.10–0.20 vs. Hg/HgO) at scan rates of 20 to 100 mV/s (Figure S7). The slope of the current density difference versus the scan rate graph was used to determine the double-layer capacitance (Cdl), which is proportional to the ECSA [88]. Capacitance (Cdl) measurements (Figure 5d) derived from CV curves demonstrated that NCS/CS/CNFs delivered the highest areal capacitance (1412.5 μF cm−2), outperforming both NCS (842.87 μF cm−2) and CS/CNFs (808.95 μF cm−2). The larger ECSA of the hybrid electrode (35.31 cm2) compared to NCS (21.07 cm2) and CS/CNFs (20.22 cm2) underscores its superior accessibility of electrochemically active sites and more efficient ion transport. EIS analysis (Figure 5e) further corroborates these results. The Nyquist plots exhibit a depressed semicircle in the high-to-medium frequency region, indicative of non-ideal capacitive behavior and surface heterogeneities. The intersection of the plots on the real impedance axis at higher frequencies represents the solution resistance (Rs). The measured Rs values for CS/CNFs, NCS, and NCS/CS/CNFs were found to be 0.33, 0.43, and 0.45 Ohm cm2, respectively. The smaller Rs and reduced charge-transfer resistance (Rct) of the NCS/CS/CNF electrode reveal faster electron and ion transport at the electrode–electrolyte interface [89,90]. These results demonstrate that the hybrid architecture provides abundant active sites and highly conductive interfaces, promoting rapid charge transfer and improved reaction kinetics, consistent with previous findings for NiOOH-CuO nano-heterostructures, where synergistic electronic interactions enhanced both OER and ethanol oxidation performance [88].
The stability (Figure 5f) of all the prepared samples was evaluated using CA at 0.65 V for 5 h. All electrodes maintained relatively stable currents with minimal decay, but the NCS/CS/CNF electrode exhibited the most robust performance. This durability can be attributed to the conductive CNF backbone, the structural reinforcement from CS, and the chemical stability of NiCo2S4 under alkaline conditions. Overall, the NCS/CS/CNF electrode demonstrated excellent OER activity, stability, and durability, making it a promising candidate for long-term electrochemical energy conversion.

3.4. Methanol Oxidation Reaction (MOR) Performance as a Bifunctional Extension

To further explore the multifunctional catalytic behavior of the designed materials, the methanol oxidation activities of CS/CNF, NCS, and NCS/CS/CNF electrodes were evaluated in N2-saturated 1.0 M KOH using CV and CA. As shown in Figure 6a, all catalysts exhibit characteristic Ni/Co redox peaks in 1.0 M KOH at a scan rate of 20 mV s−1, confirming their intrinsic electrochemical activity. Upon the introduction of methanol (0.1–2 M), the anodic current density increases significantly for all electrodes, reflecting active methanol electro-oxidation at the Ni/Co active sites. Among the samples, the NCS/CS/CNF electrode delivers the highest current response (Figure 6b), achieving 150 mA/cm2 at just 0.1 M methanol, while CS/CNF and NCS reach 91.6 and 114 mA cm−2 at 0.25 M and 0.5 M methanol, respectively. These results indicate that NCS/CS/CNFs can efficiently catalyze methanol oxidation even at low methanol concentrations, highlighting their superior bifunctional catalytic activity.
The methanol oxidation on the NCS/CS/CNF catalyst proceeds through a multi-step reaction mechanism involving surface adsorption, intermediate oxidation, and electron transfer, as described by the following reactions:
N / Co   +   2 O H   Ni ( OH ) 2 / Co ( OH ) 2 + 2 e
(CH3OH)bulk → (CH3OH)surface (diffusion)
(CH3OH)surface → (CH3OH)ads (adsorption)
Ni ( OH ) 2 / Co ( OH ) 2 + 2 O H   NiO ( OH ) / Co ( OH ) + H 2 O + 2 e
The NCS/CS/CNF catalyst exhibits a prominent anodic peak at 1.4 V vs. RHE, suggesting improved oxidation kinetics. Furthermore, its onset potential shows a noticeable negative shift compared to CS/CNFs and NCS, implying enhanced methanol adsorption and activation facilitated by synergistic Ni/Co interactions. Quantitatively, the anodic peak current density of NCS/CS/CNFs is approximately 1.63-times and 1.31-times higher than those of CS/CNFs and NCS, respectively, confirming its superior MOR activity.
Long-term stability tests were performed under optimized methanol concentrations corresponding to the highest observed current densities, as shown in Figure 6c. The CA curves recorded at 0.65 V for 18,000 s (5 h) reveal that NCS/CS/CNFs maintain a more stable current response than CS/CNFs and NCS in both KOH and KOH–methanol solutions. Additionally, Figure 6d presents the durability of NCS/CS/CNFs under varying methanol concentrations. The catalyst exhibits the most stable performance in 1 M KOH + 0.1 M CH3OH, indicating excellent tolerance to intermediate poisoning and long-term operational stability. Overall, this bifunctional electrocatalyst demonstrates high activity and durability toward both OER and MOR, making it a promising candidate for integrated electrochemical energy conversion systems.

4. Conclusions

In conclusion, transition metal sulfides (TMSs) are attractive electrocatalysts because of their tunable electronic structures, natural abundance, and strong interactions with reaction intermediates. In this work, we demonstrated a 3D heterostructure consisting of crystalline NiCo2S4 grown on amorphous Co9S8-embedded CNFs, synthesized via a combined electrospinning and hydrothermal route. XRD confirmed the coexistence of spinel NiCo2S4 and Co9S8 phases, while FESEM revealed a porous, interconnected network that facilitates efficient ion and mass transport. XPS further verified the presence of multivalent Ni2+/Ni3+ and Co2+/Co3+ states, indicating strong electronic interactions and charge redistribution across the heterointerface. The hierarchical architecture integrates three functional components: (i) a conductive CNF backbone for rapid electron transport and mechanical robustness, (ii) amorphous Co9S8 domains that introduce defect-rich active sites that accelerate reaction kinetics, and (iii) a crystalline NiCo2S4 shell that enhances multivalent redox activity and provides an abundant electroactive surface area.
As a result of these synergistic features, the NiCo2S4/Co9S8-CNF heterostructure exhibits outstanding bifunctional electrocatalytic activity. For the OER, the catalyst delivers a low overpotential of 324 mV at 10 mA cm−2, with a Tafel slope of 125.7 mV dec−1. In contrast, for the MOR, it achieves a high current density of 150 mA cm−2 at 0.1 M methanol and exhibits excellent durability at 0.65 V over 18,000 s. It demonstrates that engineering amorphous/crystalline heterointerfaces provides a practical and scalable strategy for developing next-generation electrocatalysts for alkaline water splitting and direct methanol fuel cells.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/nano15201559/s1. Figure S1: (a–b). FESEM images of NCS/CS/CNFs at different magnifications. Figure S2: (Top) SEM image and EDS elemental maps of Ni, Co, C, O, and S. (Bottom) EDS spectrum and quantified elemental composition of the NCS/CS/CNF composite. Figure S3: EDS line-scan profile across a single NCS/CS/CNF. The corresponding HAADF image shows the line-scan region. Figure S4: (a–f) SEM image and EDS elemental maps of C, Co, S, N, and O. (g and h) EDS spectrum and quantified elemental composition of the CS/CNF composite. Figure S5: Deconvoluted O 1s XPS spectra of CS/CNF, NCS, and NCS/CS/CNF samples showing the relative contributions of lattice oxygen, oxygen vacancies, and surface-adsorbed species. Figure S6: N2 adsorption–desorption isotherms (a–c) and BJH pore-size distributions (d–f) for (a,d) CS/CNFs, (b,e) NCS, and (c,f) NCS/CS/CNFs. Figure S7: CV curves of (a) CS/CNFs, (b) NCS, and (c) NCS/CS/CNFs at different scan rates (20, 40, 60, 80, and 100 mV/s) in the potential range of 0.10–0.20 V vs. Hg/HgO, which were used to calculate the Cdl.

Author Contributions

Conceptualization, D.M.; methodology, D.M.; software, D.M. and R.K.C.; validation, D.M. and R.K.C.; formal analysis, R.K.C.; investigation, D.M.; resources, M.K.; data curation, R.K.C.; writing—original draft preparation, D.M.; writing—review and editing, R.K.C. and M.K.; visualization, M.K.; supervision, M.K.; project administration, M.K.; funding acquisition, M.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by a National Research Foundation of Korea (NRF) grant, which was funded by the Korea government (MSIT) (No. RS-2019-NR040065).

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors on request.

Conflicts of Interest

The authors declare that they have no competing financial interests.

Abbreviations

The following abbreviations are used in this manuscript:
OEROxygen evolution reaction
MORMethanol oxidation reaction
DMFCsDirect methanol fuel cells
ORROxygen reduction reaction
NCS/CS/CNFsNiCo2S4 grown/anchored on Co9S8-loaded carbon nanofibers
CS/CNFCarbon nanofiber
NCSNiCo2S4
DIDeionized
KOHPotassium hydroxide
PANPolyacrylonitrile
DMFN, N-dimethylformamide
NMPN-Methyl-2-pyrrolidone
PVDFpolyvinylidene fluoride
TMSTransition metal sulfide
FESEMField emission scanning electron microscopy
TEMTransmission electron microscopy
XPSX-ray photoelectron spectroscopy
XRDPowder X-ray diffraction
BETBrunauer–Emmett–Teller
RHEReversible hydrogen electrode
LSVLinear sweep voltammetry
ECSAElectrochemically active surface area
CV Cyclic voltammetry
EISElectrochemical impedance spectroscopy
CAChronoamperometry
HAADFHigh-angle annular dark-field

References

  1. Zhang, J. Energy Access Challenge and the Role of Fossil Fuels in Meeting Electricity Demand: Promoting Renewable Energy Capacity for Sustainable Development. Geosci. Front. 2024, 15, 101873. [Google Scholar] [CrossRef]
  2. Miao, Y.; Ather Bukhari, A.A.; Bukhari, W.A.A.; Ahmad, S.; Hayat, N. Why Fossil Fuels Stifle Green Economic Growth? An Environmental Management Perspective in Assessing the Spatial Spillover Impact of Energy Consumption in South Asia. J. Environ. Manag. 2025, 373, 123471. [Google Scholar] [CrossRef]
  3. Perera, F. Pollution from Fossil-Fuel Combustion Is the Leading Environmental Threat to Global Pediatric Health and Equity: Solutions Exist. Int. J. Environ. Res. Public Health 2017, 15, 16. [Google Scholar] [CrossRef]
  4. Kabeyi, M.J.B.; Olanrewaju, O.A. Sustainable Energy Transition for Renewable and Low Carbon Grid Electricity Generation and Supply. Front. Energy Res. 2022, 9, 743114. [Google Scholar] [CrossRef]
  5. Ramgopal, N.C.; Roy, N.; El-marghany, A.; Alhammadi, S.; Sreedevi, G.; Arla, S.K.; Merum, D.; Joo, S.W. Enhancing Photo Electrocatalytic Water Splitting Efficiency Using Bi2O2CO3@Ni(OH)2 Composite with Flower-like Morphology. Ceram. Int. 2025, 51, 4388–4399. [Google Scholar] [CrossRef]
  6. Sun, H.; Xu, X.; Fei, L.; Zhou, W.; Shao, Z. Electrochemical Oxidation of Small Molecules for Energy-Saving Hydrogen Production. Adv. Energy Mater. 2024, 14, 2401242. [Google Scholar] [CrossRef]
  7. Liu, T.; Lu, J.; Chen, Z.; Luo, Z.; Ren, Y.; Zhuge, X.; Luo, K.; Ren, G.; Lei, W.; Liu, D. Advances, Mechanisms and Applications in Oxygen Evolution Electrocatalysis of Gold-Driven. Chem. Eng. J. 2024, 496, 153719. [Google Scholar] [CrossRef]
  8. Rong, C.; Huang, X.; Arandiyan, H.; Shao, Z.; Wang, Y.; Chen, Y. Advances in Oxygen Evolution Reaction Electrocatalysts via Direct Oxygen–Oxygen Radical Coupling Pathway. Adv. Mater. 2025, 37, 2416362. [Google Scholar] [CrossRef]
  9. Wu, Z.; Lu, X.F.; Zang, S.; Lou, X.W. (David) Non-Noble-Metal-Based Electrocatalysts toward the Oxygen Evolution Reaction. Adv. Funct. Mater. 2020, 30, 1910274. [Google Scholar] [CrossRef]
  10. Liu, Y.; Zhang, Y.; Tian, X.; Wang, H.; Huo, L. Ultra-High Activity Methanol Oxidation Electrocatalyzed by a Flexible Integrated Pt–Zn Array Electrode. J. Mater. Chem. C 2025, 13, 3054–3061. [Google Scholar] [CrossRef]
  11. Ding, J.; Jing, S.; Yin, C.; Ban, C.; Wang, K.; Liu, X.; Duan, Y.; Zhang, Y.; Han, G.; Gan, L.; et al. A New Insight into the Promoting Effects of Transition Metal Phosphides in Methanol Electrooxidation. Chin. Chem. Lett. 2023, 34, 107899. [Google Scholar] [CrossRef]
  12. Magama, N.; Ojemaye, M.O.; Manene, N.C.; Okoh, O.O.; Okoh, A.I. Global Research Progression on Electro-Catalysts for Direct Methanol Fuel Cells between 1992 and 2023 Using Bibliometric Indicators. Ionics 2024, 30, 5951–5968. [Google Scholar] [CrossRef]
  13. Raveendran, A.; Chandran, M.; Dhanusuraman, R. A Comprehensive Review on the Electrochemical Parameters and Recent Material Development of Electrochemical Water Splitting Electrocatalysts. RSC Adv. 2023, 13, 3843–3876. [Google Scholar] [CrossRef] [PubMed]
  14. Yi, S.; Song, X.; Shen, Y.; Xu, R.; Zhao, Y.; Chen, P. Research Progress in Alloy Catalysts for Oxygen Reduction Reaction. J. Alloys Compd. 2024, 1002, 175258. [Google Scholar] [CrossRef]
  15. Rong, C.; Dastafkan, K.; Wang, Y.; Zhao, C. Breaking the Activity and Stability Bottlenecks of Electrocatalysts for Oxygen Evolution Reactions in Acids. Adv. Mater. 2023, 35, 2211884. [Google Scholar] [CrossRef]
  16. Patil, O.U.; Park, S. Recent Progress of the Electrocatalytic CO 2 Reduction Reaction Using Porous Materials. Chem. Commun. 2025, 61, 9531–9542. [Google Scholar] [CrossRef] [PubMed]
  17. Fu, H.; Chen, Z.; Chen, X.; Jing, F.; Yu, H.; Chen, D.; Yu, B.; Hu, Y.H.; Jin, Y. Modification Strategies for Development of 2D Material-Based Electrocatalysts for Alcohol Oxidation Reaction. Adv. Sci. 2024, 11, 2306132. [Google Scholar] [CrossRef]
  18. Liu, Y.; Liu, C.; Chen, Z.; Zheng, X.; Jiang, R.; Tong, X.; Deng, Y.; Hu, W. Fabrication of Amorphous PdNiCuP Nanoparticles as Efficient Bifunctional and Highly Durable Electrocatalyst for Methanol and Formic Acid Oxidation. J. Mater. Sci. Technol. 2022, 122, 148–155. [Google Scholar] [CrossRef]
  19. Zhan, F.; Huang, L.; Luo, Y.; Chen, M.; Tan, R.; Liu, X.; Liu, G.; Feng, Z. Recent Advances on Support Materials for Enhanced Pt-Based Catalysts: Applications in Oxygen Reduction Reactions for Electrochemical Energy Storage. J. Mater. Sci. 2025, 60, 2199–2223. [Google Scholar] [CrossRef]
  20. Wang, L.; Luo, L.; Guo, Z.; Wang, Y.; Liu, X. Challenges and Strategic Advancements in Platinum-Based Catalysts for Tailored Methanol Oxidation Reaction. J. Electroanal. Chem. 2025, 978, 118875. [Google Scholar] [CrossRef]
  21. Sun, Q.; Zhang, J.; Pang, W.K.; Johannessen, B.; Li, P.; Zhao, G.; Yang, H. Advanced RuO2-Based Electrocatalysts for Oxygen Evolution Reaction: A Perspective from Coordination Structures. Mater. Today Catal. 2025, 10, 100110. [Google Scholar] [CrossRef]
  22. Martuza, M.A.; Mädler, L.; Pokhrel, S. Metal Sulfides as Potential Materials for Next Generation Lithium Ion Batteries: A Review. Adv. Energy Sustain. Res. 2025, 6, 2400448. [Google Scholar] [CrossRef]
  23. Jia, H.; Meng, L.; Lu, Y.; Liang, T.; Yuan, Y.; Hu, Y.; Dong, Z.; Zhou, Y.; Guan, P.; Zhou, L.; et al. Boosting the Efficiency of Electrocatalytic Water Splitting via in Situ Grown Transition Metal Sulfides: A Review. J. Mater. Chem. A 2024, 12, 28595–28617. [Google Scholar] [CrossRef]
  24. Zhu, M.; Wang, D.; Ge, Z.; Pan, L.; Chen, Y.; Wang, W.; Mitsuzaki, N.; Jia, S.; Chen, Z. Recent Advances in Transition Metal Sulfide-Based Electrode Materials for Supercapacitors. Chem. Commun. 2025, 61, 8314–8326. [Google Scholar] [CrossRef]
  25. Öztürk, O.; Gür, E. Layered Transition Metal Sulfides for Supercapacitor Applications. ChemElectroChem 2024, 11, e202300575. [Google Scholar] [CrossRef]
  26. Tong, Y.-L.; Xing, L.; Dai, M.-Z.; Wu, X. Hybrid Co3O4@Co9S8 Electrocatalysts for Oxygen Evolution Reaction. Front. Mater. 2019, 6, 233. [Google Scholar] [CrossRef]
  27. Li, L.; Sheng, Z.; Xiao, Q.; Hu, Q. Co9S8 Core–Shell Hollow Spheres for Enhanced Oxygen Evolution and Methanol Oxidation Reactions by Sulfur Vacancy Engineering. Dalton Trans. 2024, 53, 180–185. [Google Scholar] [CrossRef]
  28. Lv, Y.; Duan, S.; Zhu, Y.; Yin, P.; Wang, R. Enhanced OER Performances of Au@NiCo2S4 Core-Shell Heterostructure. Nanomaterials 2020, 10, 611. [Google Scholar] [CrossRef]
  29. Ning, Y.; Ma, D.; Shen, Y.; Wang, F.; Zhang, X. Constructing Hierarchical Mushroom-like Bifunctional NiCo/NiCo2S4@NiCo/Ni Foam Electrocatalysts for Efficient Overall Water Splitting in Alkaline Media. Electrochim. Acta 2018, 265, 19–31. [Google Scholar] [CrossRef]
  30. Abdelghafar, F.; Xu, X.; Guan, D.; Lin, Z.; Hu, Z.; Ni, M.; Huang, H.; Bhatelia, T.; Jiang, S.P.; Shao, Z. New Nanocomposites Derived from Cation-Nonstoichiometric Bax(Co, Fe, Zr, Y)O3−δ as Efficient Electrocatalysts for Water Oxidation in Alkaline Solution. ACS Mater. Lett. 2024, 6, 2985–2994. [Google Scholar] [CrossRef]
  31. Liu, Q.; Jin, J.; Zhang, J. NiCo2S4@graphene as a Bifunctional Electrocatalyst for Oxygen Reduction and Evolution Reactions. ACS Appl. Mater. Interfaces 2013, 5, 5002–5008. [Google Scholar] [CrossRef]
  32. Peng, W.; Wang, Y.; Yang, X.; Mao, L.; Jin, J.; Yang, S.; Fu, K.; Li, G. Co9S8 Nanoparticles Embedded in Multiple Doped and Electrospun Hollow Carbon Nanofibers as Bifunctional Oxygen Electrocatalysts for Rechargeable Zinc-Air Battery. Appl. Catal. B Environ. 2020, 268, 118437. [Google Scholar] [CrossRef]
  33. Li, J.; Xia, Y.; Luo, X.; Mao, T.; Wang, Z.; Hong, Z.; Yue, G. Bi-Phase NiCo2S4-NiS2/CFP Nanocomposites as a Highly Active Catalyst for Oxygen Evolution Reaction. Coatings 2023, 13, 313. [Google Scholar] [CrossRef]
  34. Zeng, S.; Qu, D.; Sun, H.; Chen, Y.; Wang, J.; Zheng, Y.; Pan, J.; Cao, J.; Li, C. Crystalline/Amorphous Interface Engineering and d–Sp Orbital Hybridization Synergistically Boosting the Electrocatalytic Performance of PdCu Bimetallene toward Formic Acid-Assisted Overall Water Splitting. ACS Appl. Mater. Interfaces 2024, 16, 64797–64806. [Google Scholar] [CrossRef]
  35. Zhou, Y.; Liang, Y.; Wu, Z.; Wang, X.; Guan, R.; Li, C.; Qiao, F.; Wang, J.; Fu, Y.; Baek, J. Amorphous/Crystalline Heterostructured Nanomaterials: An Emerging Platform for Electrochemical Energy Storage. Small 2025, 21, 2411941. [Google Scholar] [CrossRef] [PubMed]
  36. Zhang, Y.; Gao, F.; Wang, D.; Li, Z.; Wang, X.; Wang, C.; Zhang, K.; Du, Y. Amorphous/Crystalline Heterostructure Transition-Metal-Based Catalysts for High-Performance Water Splitting. Coord. Chem. Rev. 2023, 475, 214916. [Google Scholar] [CrossRef]
  37. Li, S.; Sirisomboonchai, S.; An, X.; Ma, X.; Li, P.; Ling, L.; Hao, X.; Abudula, A.; Guan, G. Engineering Interfacial Structures to Accelerate Hydrogen Evolution Efficiency of MoS 2 over a Wide PH Range. Nanoscale 2020, 12, 6810–6820. [Google Scholar] [CrossRef] [PubMed]
  38. García-Osorio, D.A.; Jaimes, R.; Vazquez-Arenas, J.; Lara, R.H.; Alvarez-Ramirez, J. The Kinetic Parameters of the Oxygen Evolution Reaction (OER) Calculated on Inactive Anodes via EIS Transfer Functions: •OH Formation. J. Electrochem. Soc. 2017, 164, E3321–E3328. [Google Scholar] [CrossRef]
  39. van der Heijden, O.; Park, S.; Vos, R.E.; Eggebeen, J.J.J.; Koper, M.T.M. Tafel Slope Plot as a Tool to Analyze Electrocatalytic Reactions. ACS Energy Lett. 2024, 9, 1871–1879. [Google Scholar] [CrossRef]
  40. Rezić Meštrović, I.; Somogyi Škoc, M. Nanofiber-Based Innovations in Energy Storage Systems. Polymers 2025, 17, 1456. [Google Scholar] [CrossRef]
  41. Muralee Gopi, C.V.V.; Ravi, S.; Rao, S.S.; Eswar Reddy, A.; Kim, H.-J. Carbon Nanotube/Metal-Sulfide Composite Flexible Electrodes for High-Performance Quantum Dot-Sensitized Solar Cells and Supercapacitors. Sci. Rep. 2017, 7, 46519. [Google Scholar] [CrossRef]
  42. Cai, W.; Lai, T.; Lai, J.; Xie, H.; Ouyang, L.; Ye, J.; Yu, C. Transition Metal Sulfides Grown on Graphene Fibers for Wearable Asymmetric Supercapacitors with High Volumetric Capacitance and High Energy Density. Sci. Rep. 2016, 6, 26890. [Google Scholar] [CrossRef]
  43. He, W.; Xu, W.; Li, Z.; Hu, Z.; Yang, J.; Qin, G.; Teng, W.; Zhang, T.; Zhang, W.; Sun, Z.; et al. Structural Design and Challenges of Micron-Scale Silicon-Based Lithium-ion Batteries. Adv. Sci. 2025, 12, 2407540. [Google Scholar] [CrossRef] [PubMed]
  44. Sun, Y.-G.; Hu, Y.; Dong, L.; Zhou, T.-T.; Qian, X.-Y.; Zhang, F.-J.; Shen, J.-Q.; Shan, Z.-Y.; Yang, L.-P.; Lin, X.-J. Unlocking the Potential of Na2Ti3O7-C Hollow Microspheres in Sodium-Ion Batteries via Template-Free Synthesis. Nanomaterials 2025, 15, 423. [Google Scholar] [CrossRef]
  45. Fan, P.; Ye, C.; Xu, L. Core-shell Nanofiber-based Electrodes for High-performance Asymmetric Supercapacitors. ChemistrySelect 2023, 8, e202204669. [Google Scholar] [CrossRef]
  46. Fan, P.; Wang, J.; Ding, W.; Xu, L. Core–Shell Structured Carbon Nanofiber-Based Electrodes for High-Performance Supercapacitors. Molecules 2023, 28, 4571. [Google Scholar] [CrossRef] [PubMed]
  47. Wang, D.; Zhao, X.; Lan, J.; Zhang, P.; Shao, G.; Miao, F. Rational Design of Core–Shell Nanofibers Structure Integrated with Cobalt Sulfide Catalyst for High-Performance Lithium-Sulfur Batteries. Appl. Surf. Sci. 2025, 710, 163944. [Google Scholar] [CrossRef]
  48. Govindasamy, M.; Shanthi, S.; Elaiyappillai, E.; Wang, S.-F.; Johnson, P.M.; Ikeda, H.; Hayakawa, Y.; Ponnusamy, S.; Muthamizhchelvan, C. Fabrication of Hierarchical NiCo2S4@CoS2 Nanostructures on Highly Conductive Flexible Carbon Cloth Substrate as a Hybrid Electrode Material for Supercapacitors with Enhanced Electrochemical Performance. Electrochim. Acta 2019, 293, 328–337. [Google Scholar] [CrossRef]
  49. Li, M.; Zhou, S.; Sun, R.; Han, S.; Jiang, J. Hierarchical Electronic Coupling Engineering of Bimetallic Sulfide Driven by CoFe Bimetal MOFs Anchored on MXene Promotes Efficient Overall Water Splitting. Fuel 2024, 358, 130256. [Google Scholar] [CrossRef]
  50. Duan, Y.; Guo, Z.; Wang, T.; Zhang, J. Uniform Anchoring of MoS2 Nanosheets on MOFs-Derived CoFe2O4 Porous Nanolayers to Construct Heterogeneous Structural Configurations for Efficient and Stable Overall Water Splitting. J. Colloid Interface Sci. 2025, 680, 541–551. [Google Scholar] [CrossRef]
  51. Wang, C.; Kosari, M.; Xi, S.; Zeng, H.C. Uniform Si-Infused UiO-66 as a Robust Catalyst Host for Efficient CO2 Hydrogenation to Methanol. Adv. Funct. Mater. 2023, 33, 2212478. [Google Scholar] [CrossRef]
  52. Ding, F.; Ji, P.; Han, Z.; Hou, X.; Yang, Y.; Hu, Z.; Niu, Y.; Liu, Y.; Zhang, J.; Rong, X.; et al. Tailoring Planar Strain for Robust Structural Stability in High-Entropy Layered Sodium Oxide Cathode Materials. Nat. Energy 2024, 9, 1529–1539. [Google Scholar] [CrossRef]
  53. Gong, J.; Luo, W.; Zhao, Y.; Xie, M.; Wang, J.; Yang, J.; Dai, Y. Co9S8/NiCo2S4 Core-Shell Array Structure Cathode Hybridized with PPy/MnO2 Core-Shell Structure Anode for High-Performance Flexible Quasi-Solid-State Alkaline Aqueous Batteries. Chem. Eng. J. 2022, 434, 134640. [Google Scholar] [CrossRef]
  54. Jin, R.; Liu, J.; Yu, H.; Xia, J.; Wu, X.; Yuan, B.; Li, R.; Xu, H. Core-Shell NiCo2S4@C with Three Dimensional Carbon Frameworks for Enhanced Asymmetrical Supercapacitor and Lithium Storage Performance. Colloids Surf. A Physicochem. Eng. Asp. 2024, 697, 134445. [Google Scholar] [CrossRef]
  55. Kong, W.; Lu, C.; Zhang, W.; Pu, J.; Wang, Z. Homogeneous Core–Shell NiCo2S4 Nanostructures Supported on Nickel Foam for Supercapacitors. J. Mater. Chem. A 2015, 3, 12452–12460. [Google Scholar] [CrossRef]
  56. Meng, Y.; Ding, J.; Liu, Y.; Hu, G.; Feng, Y.; Wu, Y.; Liu, X. Advancements in Amorphous Oxides For Electrocatalytic Carbon Dioxide Reduction. Mater. Today Catal. 2024, 7, 100065. [Google Scholar] [CrossRef]
  57. Jin, Y.; Zhang, M.; Song, L.; Zhang, M. Research Advances in Amorphous-Crystalline Heterostructures Toward Efficient Electrochemical Applications. Small 2023, 19, 2206081. [Google Scholar] [CrossRef] [PubMed]
  58. Lan, H.; Wang, J.; Cheng, L.; Yu, D.; Wang, H.; Guo, L. The Synthesis and Application of Crystalline–Amorphous Hybrid Materials. Chem. Soc. Rev. 2024, 53, 684–713. [Google Scholar] [CrossRef] [PubMed]
  59. Abedi, M.; Rezaee, S.; Shahrokhian, S. Designing Core-Shell Heterostructure Arrays Based on Snowflake NiCoFe-LTH Shelled over W2N-WC Nanowires as an Advanced Bi-Functional Electrocatalyst for Boosting Alkaline Water/Seawater Electrolysis. J. Colloid Interface Sci. 2024, 666, 307–321. [Google Scholar] [CrossRef]
  60. Mouloua, D.; Vicart, T.; Rajput, N.S.; Asbani, B.; El Marssi, M.; El Khakani, M.A.; Jouiad, M. Core/Shell 1T/2H-MoS2 Nanoparticle Induced Synergistic Effects for Enhanced Hydrogen Evolution Reaction. J. Colloid Interface Sci. 2025, 687, 851–859. [Google Scholar] [CrossRef]
  61. Zhu, Y.; Fang, Z.; Zhang, Z.; Wu, H. Discontinuous Phase Diagram of Amorphous Carbons. Natl. Sci. Rev. 2024, 11, nwae051. [Google Scholar] [CrossRef]
  62. Ahmed, S.; Ahmad, M.; Yousaf, M.H.; Haider, S.; Imran, Z.; Batool, S.S.; Ahmad, I.; Shahzad, M.I.; Azeem, M. Solvent-Free Synthesis of NiCo2S4 Having the Metallic Nature. Front. Chem. 2022, 10, 1027024. [Google Scholar] [CrossRef] [PubMed]
  63. Aftab, U.; Tahira, A.; Mazzaro, R.; Morandi, V.; Ishaq Abro, M.; Baloch, M.M.; Yu, C.; Ibupoto, Z.H. Nickel–Cobalt Bimetallic Sulfide NiCo 2 S 4 Nanostructures for a Robust Hydrogen Evolution Reaction in Acidic Media. RSC Adv. 2020, 10, 22196–22203. [Google Scholar] [CrossRef]
  64. Wang, H.; Liang, Y.; Li, Y.; Dai, H. Co1-xS–Graphene Hybrid: A High-Performance Metal Chalcogenide Electrocatalyst for Oxygen Reduction. Angew. Chem. Int. Ed. 2011, 50, 10969–10972. [Google Scholar] [CrossRef]
  65. Serhan, M.; Jackemeyer, D.; Long, M.; Sprowls, M.; Perez, I.D.; Maret, W.; Chen, F.; Tao, N.; Forzani, E. Total Iron Measurement in Human Serum with a novel Smartphone-Based Assay. IEEE J Transl Eng Health Med. 2020, 8, 2800309. [Google Scholar] [CrossRef]
  66. Sibokoza, S.B.; Moloto, M.J.; Moloto, N.; Sibiya, P.N. The Effect of Temperature and Precursor Concentration on the Synthesis of Cobalt Sulphide Nanoparticles Using Cobalt Diethyldithiocarbamate Complex. Chalcogenide Lett. 2017, 14, 69–78. [Google Scholar]
  67. Liao, M.; Zhang, K.; Yan, W.; Yue, H.; Luo, C.; Wu, G.; Zeng, H. Constructing Amorphous/Crystalline Heterointerface in Cobalt Sulfide for High-Performance Supercapacitors. J. Power Sources 2025, 625, 235663. [Google Scholar] [CrossRef]
  68. Ghosh, A.; Ghosh, S.; Seshadhri, G.M.; Ramaprabhu, S. Green Synthesis of Nitrogen-Doped Self-Assembled Porous Carbon-Metal Oxide Composite towards Energy and Environmental Applications. Sci. Rep. 2019, 9, 5187. [Google Scholar] [CrossRef] [PubMed]
  69. Greczynski, G.; Hultman, L. The Same Chemical State of Carbon Gives Rise to Two Peaks in X-Ray Photoelectron Spectroscopy. Sci. Rep. 2021, 11, 11195. [Google Scholar] [CrossRef]
  70. Fujimoto, A.; Yamada, Y.; Koinuma, M.; Sato, S. Origins of Sp3C Peaks in C1s X-Ray Photoelectron Spectra of Carbon Materials. Anal. Chem. 2016, 88, 6110–6114. [Google Scholar] [CrossRef] [PubMed]
  71. Mi, L.; Wei, W.; Huang, S.; Cui, S.; Zhang, W.; Hou, H.; Chen, W. A Nest-like Ni@Ni1.4Co1.6S2 Electrode for Flexible High-Performance Rolling Supercapacitor Device Design. J. Mater. Chem. A 2015, 3, 20973–20982. [Google Scholar] [CrossRef]
  72. Zhao, Y.; Luo, Y.; Sun, B.; Li, T.; Han, S.; Dong, Z.; Lin, H. Rational Construction of Reduced NiCo2S4@CuCo2S4 Composites with Sulfur Vacancies as High-Performance Supercapacitor Electrode for Enhancing Electrochemical Energy Storage. Compos. Part B Eng. 2022, 243, 110088. [Google Scholar] [CrossRef]
  73. Zhao, J.; Wang, Y.; Qian, Y.; Jin, H.; Tang, X.; Huang, Z.; Lou, J.; Zhang, Q.; Lei, Y.; Wang, S. Hierarchical Design of Cross-Linked NiCo2S4 Nanowires Bridged NiCo-Hydrocarbonate Polyhedrons for High-Performance Asymmetric Supercapacitor. Adv. Funct. Mater. 2023, 33, 2210238. [Google Scholar] [CrossRef]
  74. He, W.; Wu, S.; Zhang, Z.; Yang, Q. Vacancy-Rich Graphene Supported Electrocatalysts Synthesized by Radio-Frequency Plasma for Oxygen Evolution Reaction. Inorg. Chem. Front. 2022, 9, 3854–3864. [Google Scholar] [CrossRef]
  75. Lu, Z.; Wang, Y.; Cheng, R.; Yang, L.; Wang, N. Highly Dispersed Co/Co9S8 Nanoparticles Encapsulated in S, N Co-Doped Lo Ngan Shell-Derived Hierarchical Porous Carbon for Corrosion-Resistant, Waterproof High-Performance Microwave Absorption. SSRN Electron. J. 2022, 637, 147–158. [Google Scholar] [CrossRef]
  76. Lu, A.; Chen, Y.; Li, H.; Dowd, A.; Cortie, M.B.; Xie, Q.; Guo, H.; Qi, Q.; Peng, D.-L. Magnetic Metal Phosphide Nanorods as Effective Hydrogen-Evolution Electrocatalysts. Int. J. Hydrogen Energy 2014, 39, 18919–18928. [Google Scholar] [CrossRef]
  77. Dong, Y.; Ran, J.; Liu, Q.; Zhang, G.; Jiang, X.; Gao, D. Hydrogen-Etched CoS2 to Produce a Co9S8@CoS2 Heterostructure Electrocatalyst for Highly Efficient Oxygen Evolution Reaction. RSC Adv. 2021, 11, 30448–30454. [Google Scholar] [CrossRef] [PubMed]
  78. Dong, Z.; Li, M.; Zhang, W.; Liu, Y.; Wang, Y.; Qin, C.; Yu, L.; Yang, J.; Zhang, X.; Dai, X. Cobalt Nanoparticles Embedded in N, S Co-Doped Carbon towards Oxygen Reduction Reaction Derived by in Situ Reducing Cobalt Sulfide. ChemCatChem 2019, 11, 6039–6050. [Google Scholar] [CrossRef]
  79. Zhang, S.; Zhai, D.; Sun, T.; Han, A.; Zhai, Y.; Cheong, W.-C.; Liu, Y.; Su, C.; Wang, D.; Li, Y. In Situ Embedding Co9S8 into Nitrogen and Sulfur Codoped Hollow Porous Carbon as a Bifunctional Electrocatalyst for Oxygen Reduction and Hydrogen Evolution Reactions. Appl. Catal. B Environ. 2019, 254, 186–193. [Google Scholar] [CrossRef]
  80. Kim, S.; van Ommen, J.R.; La Zara, D.; Courtois, N.; Davin, J.; Recker, C.; Schoeffel, J.; Blume, A.; Talma, A.; Dierkes, W.K. Molecular Layer Deposition (MLD) of a Blocked Mercapto Silane on Precipitated Silica. Org. Mater. 2023, 5, 139–147. [Google Scholar] [CrossRef]
  81. Yang, Y.; Qian, D.; Zhu, H.; Zhou, Q.; Zhang, Z.; Li, Z.; Hu, Z. Construction of Tremella-like Co9S8@NiCo2S4 Heterostructure Nanosheets Integrated Electrode for High-Performance Hybrid SupercapacitorsConceptualization, Methodology, Formal Analysis. J. Alloys Compd. 2022, 898, 162850. [Google Scholar] [CrossRef]
  82. Song, Y.; Li, C.; Zhou, Y.; Tang, T.; Wang, J.; Shang, Y.; Deng, C. Co/Co9S8 Nanofibers Decorated with Carbon Nanotubes as Oxygen Electrocatalysts for Zn–Air Batteries. ACS Appl. Nano Mater. 2024, 7, 25152–25161. [Google Scholar] [CrossRef]
  83. Zhang, H.; Guan, D.; Gu, Y.; Xu, H.; Wang, C.; Shao, Z.; Guo, Y. Tuning Synergy between Nickel and Iron in Ruddlesden–Popper Perovskites through Controllable Crystal Dimensionalities towards Enhanced Oxygen-evolving Activity and Stability. Carbon Energy 2024, 6, e465. [Google Scholar] [CrossRef]
  84. Sun, C.; Zhang, W.; Gu, Z.; Li, X.; Kang, H.; Li, Z.; Li, Z. Improved Energy Storage Performance of NiS2/CoNi2S4 Heterostructure with Reduced Graphene Oxide and Carbon Nanofiber Synergistic Optimization for Hybrid Supercapacitor. Diam. Relat. Mater. 2025, 159, 112782. [Google Scholar] [CrossRef]
  85. Quan, G.; Wu, Y.; Li, W.; Li, D.; Liu, X.; Wang, K.; Dai, S.; Xiao, L.; Ao, Y. Construction of Cellulose Nanofiber/Carbon Nanotube Synergistic Network on Carbon Fiber Surface to Enhance Mechanical Properties and Thermal Conductivity of Composites. Compos. Sci. Technol. 2024, 248, 110454. [Google Scholar] [CrossRef]
  86. Sabir, A.S.; Pervaiz, E.; Khosa, R.; Sohail, U. An Inclusive Review and Perspective on Cu-Based Materials for Electrochemical Water Splitting. RSC Adv. 2023, 13, 4963–4993. [Google Scholar] [CrossRef] [PubMed]
  87. Majhi, K.C.; Yadav, M. Transition Metal Chalcogenides Based Nanocomposites as Efficient Electrocatalyst for Hydrogen Evolution Reaction over the Entire PH Range. Int. J. Hydrogen Energy 2020, 45, 24219–24231. [Google Scholar] [CrossRef]
  88. Sun, H.; Li, L.; Chen, Y.; Kim, H.; Xu, X.; Guan, D.; Hu, Z.; Zhang, L.; Shao, Z.; Jung, W. Boosting Ethanol Oxidation by NiOOH-CuO Nano-Heterostructure for Energy-Saving Hydrogen Production and Biomass Upgrading. Appl. Catal. B Environ. 2023, 325, 122388. [Google Scholar] [CrossRef]
  89. Zafar, S.; Thomas, A.; Mahapatra, S.N.; Karmodak, N.; Arora, H.S.; Lochab, B. Morphology-Dependent Enhancement of the Electrochemical Performance of CNF-Guided Tunable VS 4 Heterostructures for Symmetric Supercapacitors. J. Mater. Chem. A 2023, 11, 21263–21271. [Google Scholar] [CrossRef]
  90. Jagadale, A.; Zhou, X.; Blaisdell, D.; Yang, S. Carbon Nanofibers (CNFs) Supported Cobalt- Nickel Sulfide (CoNi2S4) Nanoparticles Hybrid Anode for High Performance Lithium Ion Capacitor. Sci. Rep. 2018, 8, 1602. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of the synthesis of NCS/CS/CNF composites via electrospinning of PAN–cobalt acetate precursor, followed by carbonization/sulfurization and subsequent hydrothermal growth of Ni-Co sulfide nanostructures.
Figure 1. Schematic illustration of the synthesis of NCS/CS/CNF composites via electrospinning of PAN–cobalt acetate precursor, followed by carbonization/sulfurization and subsequent hydrothermal growth of Ni-Co sulfide nanostructures.
Nanomaterials 15 01559 g001
Figure 2. FESEM images of (ac) CS/CNFs, (df) NCS, and (gl) NCS/CS/CNFs from two distinct regions. Scale bars: 5 μm (a,d,g,j), 1 μm (b,e,h,k), and 500 nm (c,f,i,l).
Figure 2. FESEM images of (ac) CS/CNFs, (df) NCS, and (gl) NCS/CS/CNFs from two distinct regions. Scale bars: 5 μm (a,d,g,j), 1 μm (b,e,h,k), and 500 nm (c,f,i,l).
Nanomaterials 15 01559 g002
Figure 3. (ac) TEM and HRTEM images of NCS/CS/CNFs; (d) HAADF image and corresponding elemental maps for C, Ni, Co, and S.
Figure 3. (ac) TEM and HRTEM images of NCS/CS/CNFs; (d) HAADF image and corresponding elemental maps for C, Ni, Co, and S.
Nanomaterials 15 01559 g003
Figure 4. (a) XRD spectra; (b) XPS survey spectra; and (c) C 1s, (d) Ni 2p, (e) Co 2p, and (f) S 2p XPS spectra of CS/CNF, NCS, and NCS/CS/CNF samples.
Figure 4. (a) XRD spectra; (b) XPS survey spectra; and (c) C 1s, (d) Ni 2p, (e) Co 2p, and (f) S 2p XPS spectra of CS/CNF, NCS, and NCS/CS/CNF samples.
Nanomaterials 15 01559 g004
Figure 5. OER performance of CS/CNF, NCS, and NCS/CS/CNF electrodes in 1 M KOH: (a) CV curves at 20 mV s−1, (b) LSV profiles, (c) Tafel slopes, (d) areal capacitance derived from CV, (e) Nyquist plots with equivalent circuit model, and (f) chronoamperometry stability at 0.65 V for 5 h.
Figure 5. OER performance of CS/CNF, NCS, and NCS/CS/CNF electrodes in 1 M KOH: (a) CV curves at 20 mV s−1, (b) LSV profiles, (c) Tafel slopes, (d) areal capacitance derived from CV, (e) Nyquist plots with equivalent circuit model, and (f) chronoamperometry stability at 0.65 V for 5 h.
Nanomaterials 15 01559 g005
Figure 6. MOR performance of CS/CNF, NCS, and NCS/CS/CNF electrodes: (a) CV curves in 1 M KOH with and without methanol at 20 mV s−1; (b) comparison of anodic current densities at different methanol concentrations; (c) chronoamperometry stability curves of CS/CNFs (1 M KOH with/with 0.25 M CH3OH), NCS (1 M KOH with/with 0.5 M CH3OH), and NCS/CS/CNFs (1 M KOH with/with 0.1 M CH3OH) at 0.65 V for 18,000 s; and (d) long-term durability tests of NCS/CS/CNFs at various methanol concentrations in 1 M KOH.
Figure 6. MOR performance of CS/CNF, NCS, and NCS/CS/CNF electrodes: (a) CV curves in 1 M KOH with and without methanol at 20 mV s−1; (b) comparison of anodic current densities at different methanol concentrations; (c) chronoamperometry stability curves of CS/CNFs (1 M KOH with/with 0.25 M CH3OH), NCS (1 M KOH with/with 0.5 M CH3OH), and NCS/CS/CNFs (1 M KOH with/with 0.1 M CH3OH) at 0.65 V for 18,000 s; and (d) long-term durability tests of NCS/CS/CNFs at various methanol concentrations in 1 M KOH.
Nanomaterials 15 01559 g006
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

Merum, D.; Chava, R.K.; Kang, M. Engineering 3D Heterostructured NiCo2S4/Co9S8-CNFs via Electrospinning and Hydrothermal Strategies for Efficient Bifunctional Energy Conversion. Nanomaterials 2025, 15, 1559. https://doi.org/10.3390/nano15201559

AMA Style

Merum D, Chava RK, Kang M. Engineering 3D Heterostructured NiCo2S4/Co9S8-CNFs via Electrospinning and Hydrothermal Strategies for Efficient Bifunctional Energy Conversion. Nanomaterials. 2025; 15(20):1559. https://doi.org/10.3390/nano15201559

Chicago/Turabian Style

Merum, Dhananjaya, Rama Krishna Chava, and Misook Kang. 2025. "Engineering 3D Heterostructured NiCo2S4/Co9S8-CNFs via Electrospinning and Hydrothermal Strategies for Efficient Bifunctional Energy Conversion" Nanomaterials 15, no. 20: 1559. https://doi.org/10.3390/nano15201559

APA Style

Merum, D., Chava, R. K., & Kang, M. (2025). Engineering 3D Heterostructured NiCo2S4/Co9S8-CNFs via Electrospinning and Hydrothermal Strategies for Efficient Bifunctional Energy Conversion. Nanomaterials, 15(20), 1559. https://doi.org/10.3390/nano15201559

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