Open Access Article
Kumasser Kusse Kuchayita
a,
Yohannes Asmare Fessehaa,
Chih-Wei Chiu
c,
Jem-Kun Chen
c,
Ai-Wei Lee*d and
Chih-Chia Cheng
*ab
aGraduate Institute of Applied Science and Technology, National Taiwan University of Science and Technology, Taipei 10607, Taiwan. E-mail: cccheng@mail.ntust.edu.tw
bAdvanced Membrane Materials Research Center, National Taiwan University of Science and Technology, Taipei 10607, Taiwan
cDepartment of Materials Science and Engineering, National Taiwan University of Science and Technology, Taipei 10607, Taiwan
dDepartment of Anatomy and Cell Biology, School of Medicine, College of Medicine, Taipei Medical University, Taipei, 11031, Taiwan. E-mail: ammielee@tmu.edu.tw
First published on 16th December 2025
Tungsten disulfide (WS2) exhibits decent electrocatalytic performance toward the hydrogen evolution reaction (HER) in acidic conditions; however, poor structural stability under alkaline conditions limits its widespread application in electrocatalysis. We developed a robust exfoliation strategy using water-soluble sodium-ion-functionalized chitosan to obtain stable, well-dispersed WS2 nanosheets with excellent resistance to alkaline environments. Through subsequent electropolymerization and electroactivation, N-type WS2 nanosheets were combined with P-type polyaniline on nickel foam, creating a stable organic–inorganic P–N heterojunction electrocatalytic electrode. The electrode demonstrates high HER electrocatalytic performance in 1.0 M potassium hydroxide solution, achieving a low overpotential of 24.5 mV at 10 mA cm−2, a Tafel slope of 48.2 mV dec−1 and a low resistance of around 0.5 Ω, equivalent to the conventional noble-metal Pt/C electrocatalyst. More importantly, the electrode maintained excellent long-term electrocatalytic performance and structural integrity after 1000 cycles of cyclic voltammetry and 24 h of continuous operation at 100 mA cm−2. In contrast, the catalytic activity of commercial Pt/C declined substantially, indicating poor stability under alkaline conditions. Therefore, these findings overcome the limitations of WS2 under alkaline conditions and provide a cost-effective strategy for producing highly active, pH-universal electrocatalysts suitable for water electrolysis and sustainable hydrogen production.
Transition metal-based catalysts, such as sulfides, selenides, phosphides, and carbides, have garnered significant research interest as viable substitutes for noble metal-based electrocatalysts to enhance the kinetics of the HER.1,8 Although these materials typically display great activity and stability in acidic environments, their performance in alkaline media remains limited due to lower catalytic activity or poor stability.9–12 For example, tungsten disulfide (WS2), one of the two-dimensional transition metal dichalcogenides (TMDs), holds great potential for applications due to its favorable hydrogen adsorption energy at edge sites, inherent stability, and high activity for the HER in acidic media.13,14 However, the performance of WS2 in alkaline media is hindered by poor conductivity, strong hydroxide adsorption, and sluggish water dissociation kinetics; these issues stem from structural and electronic properties.15,16 Despite extensive efforts, strong hydrogen evolution activity remains largely confined to acidic environments, while the reaction kinetics in the more practically viable alkaline media are relatively sluggish.17,18 Similarly, due to the intrinsic properties of molybdenum disulfide (MoS2), including inefficient water adsorption and dissociation caused by an unfavorable alignment of orbitals, MoS2, a member of the TMDs, also exhibits limited activity in alkaline hydrogen evolution catalysis.19,20 Thus, due to these limitations, development of hybrid materials or improvements to existing catalysts are necessary to enhance the performance of the HER across a broader pH range.
Researchers have employed various strategies with the aim of accelerating the overall kinetics of the HER under alkaline conditions, such as creating a nanoscale morphology and size modulation of WS2, heteroatom doping, defect engineering, and construction of heterostructures. Multiple functional components have been integrated into hybrid heterostructure systems in attempts to optimize various steps of the HER process.16 For example, combining WS2 with water dissociation promoters or conductive polymers enhanced electron transfer, improved mass transport, and increased the density of exposed active sites. Feng et al. prepared a trinickel disulfide (Ni3S2)/MoS2 heterostructure to demonstrate that modulation of the interfacial and electronic structure of MoS2 could enhance its performance in the alkaline HER. Specifically, the integration of Ni3S2 promotes H2O adsorption and dissociation and thereby improves the overall alkaline HER kinetics of the Ni3S2/MoS2 composite.16 Moreover, Hussain et al. developed a tungsten carbide (W2C)/WS2 hybrid that exhibited remarkable electrocatalytic performance in both alkaline and acidic conditions. The heterostructure exhibited enhanced performance, as evidenced by overpotentials (η10) of 133 mV and 105 mV at 10 mA cm−2, along with Tafel slopes of 70 mV dec−1 and 84 mV dec−1, respectively.21 Similarly, Hussain et al. also fabricated a molybdenum carbide (Mo2C)@WS2 hybrid via a simple chemical method. The resulting heterostructure electrocatalyst exhibited low Tafel slopes of 59 and 95 mV dec−1 and low overpotentials of 93 mV and 98 mV at 10 mA cm−2 in acidic and alkaline solutions, respectively, during the HER.22 The formation of heterostructures between two active components induces a non-equilibrium state in the catalyst. This generates charged interfacial regions and strong built-in electric fields, which significantly alter the electronic density of interfacial atoms. The built-in electric field at the interface can enhance catalytic activity by accelerating charge transfer, improving conductivity, and optimizing the electronic structure.23 Significant progress has been made through techniques such as heteroatom doping, defect engineering, and heterostructure formation.24 However, the challenging issues related to interfacial compatibility and stability between components—which arise from differences in material properties—continue to limit the electrocatalytic performance of electrodes for the HER.25,26 Even with optimized structural and compositional tuning, the improvements are still quite limited.27
To overcome the challenges of interfacial compatibility and stability, many research groups have focused on exploring various approaches to enhance the electrocatalytic activity and stability of WS2 under different environmental conditions. For example, the creation of a heterojunction interface between WS2 and the zirconium-based metal–organic framework UiO-66 induces synergistic interactions that facilitate efficient charge transfer and thereby significantly enhanced HER/OER activity under alkaline conditions.28 Similarly, Zhang et al. developed a sea-anemone-like core–shell heterojunction composed of oxygen-vacancy-rich CoP@CoOOH on a carbon paper substrate, which exhibited remarkable HER catalytic performance in both alkaline and neutral media. This enhancement was ascribed to a synergistic interaction between the core and shell constituents.29 Recently, our team employed an integrated strategy combining sodium-carboxymethyl chitosan (Na-CMC)-assisted exfoliation, electropolymerization (EP), and electrochemical activation (EA) to couple P-type PANI with N-type WS2 on nickel foam, thereby creating an intrinsic P–N heterojunction electrode assigned as Na-CMC/WS2/PANI/NF. This design improved interfacial compatibility between PANI and WS2, facilitated efficient charge transport, increased the density of accessible catalytic active sites, and accelerated the reaction kinetics. Consequently, the electrodes demonstrated excellent HER electrocatalytic performance and long-term structural stability in acidic media.26 These findings also demonstrate that the presence of a P–N heterojunction within the material matrix not only represents an effective strategy to enhance interfacial compatibility and structural stability but also facilitates stable charge transport across the interface to thereby achieve the desired electrocatalytic performance. Therefore, we confidently propose that incorporating a P–N heterojunction into the WS2 matrix can markedly improve its long-term structural stability and overall electrocatalytic performance for the HER in alkaline environments. This strategy not only mitigates the intrinsic instability of WS2 under alkaline conditions but also unlocks its capabilities as a noble-metal-free electrocatalyst for diverse energy conversion applications.
In this work, we fabricated a Na-CMC/WS2/PANI/NF electrode featuring P–N heterojunction interfaces through a process involving Na-CMC-assisted exfoliation, EP, and EA. This electrode not only effectively overcomes the inherent structural instability of WS2 under alkaline conditions, but also exhibits excellent HER catalytic activity in alkaline media. Electrocatalytic HER evaluation demonstrated that the fabricated electrode exhibits outstanding electrocatalytic activity in a 1.0 M potassium hydroxide (KOH) solution after electroactivation processing treatment. This performance was defined by an η10 of 24.5 mV with a Tafel slope of 48.2 mV dec−1 and a resistance of approximately 0.5 Ω, demonstrating results on par with those of commercial Pt/C electrocatalysts. Importantly, unlike Pt/C, which suffers notable performance loss after 1000 cycles of cyclic voltammetry (CV) and 24 h of continuous operation at 100 mA cm−2, the activated electrode (Na-CMC/WS2/PANI/NF) maintains high catalytic activity and structural integrity. Therefore, this newly developed strategy for constructing P–N heterojunctions offers an effective approach to overcome the intrinsic limitations of WS2 for the HER under alkaline conditions and may help to enable the realization of cost-effective and high-efficiency WS2-based HER electrocatalysts with promising potential to replace Pt/C catalysts in various energy-related applications.
000 g mol−1), and KOH (85%) were purchased from Thermo Scientific Chemicals. Sulfuric acid (H2SO4, 97%) and hydrochloric acid (HCl, 37%) were obtained from Honeywell Research Chemicals and used as received without further purification. NF, supplied by MTI Corporation, had a high purity (>99.96%), a density of 346 g m−2, a thickness of 1.6 mm, and a porosity greater than 95%. The synthesis of Na-CMC and Na-CMC-functionalized exfoliated WS2 nanosheets followed the procedures outlined in our previously published work.26,30 Additional materials used include 20% Pt/C as a reference catalyst, Nafion solution as a binder, Pt wire and graphite rod as counter electrodes, and a silver/silver chloride (Ag/AgCl) electrode as the reference electrode.
:
3 (Na-CMC
:
WS2) and subjected to ultrasonication for 30 minutes at 50% amplitude utilizing a probe sonicator (Q700, Osonica, Newtown, CT, USA), while maintaining the temperature below 5 °C using an ice bath. The resulting dark green suspension was filtered and centrifuged at 4000 rpm for 10 min to remove unexfoliated materials. The upper ∼85% of the supernatant was carefully harvested for subsequent analysis and electrode fabrication.
:
2 volume ratio. This polymerization process was carried out at a constant current (0.10 mA) for a duration of 40 minutes. After deposition, the resulting electrodes were cleaned carefully with DI water to eliminate any loosely bound species and residual aniline salts and then dried under ambient conditions. For comparison, PANI was electrodeposited on NF under identical conditions from a saturated anilinium chloride solution without WS2, resulting in PANI/NF control electrodes. The successful electropolymerization and nanosheets incorporation were evidenced by a greenish coloration of the NF substrate and a concurrent decrease in potentiometric response.
:
5 v/v) and 100 µL of 5 wt% Nafion solution. This mixture was sonicated for 30 min to ensure uniform dispersion. Subsequently, 250 µL of the ink was drop-cast onto a 1 × 1 cm2 NF substrate and dried naturally to obtain the Pt/C/NF electrode.
| η(HER) = (0 − Eobs) V vs. RHE. |
The overpotential required to reach a current density of −10 mA cm−2, denoted as η10, was used as a benchmark to evaluate the electrocatalytic HER performance.
The Tafel slope provides crucial evidence about the rate-determining step and reaction mechanisms based on measuring the voltage change required to increase the electrochemical current by a factor of ten.31
The Tafel slopes were determined by fitting η as a function of the logarithm of the current density (log
j) using the Tafel equation, assuming uniform concentrations in the bulk solution and at the electrode–electrolyte interface:
17,18,32 and commercial Pt/C catalysts33,34 during HER processes in alkaline environments. Prior to testing HER activity, the overall dispersion stability of exfoliated WS2 nanosheets in alkaline solution was examined by varying the mixing ratios of Na-CMC and WS2. The transmittance of the obtained 1/1, 1/3, and 1/5 Na-CMC/WS2 dispersions at 500 nm was monitored over time using UV-vis spectroscopy at pH 13.5 and 25 °C. As shown in Fig. S1, the 1/3 and 1/5 dispersions exhibited only an 8% increase in transmittance after 63 days of monitoring, whereas the 1/1 dispersion increased by 48%, indicating that excess Na-CMC promotes WS2 nanosheets self-aggregation.25,26 Therefore, the 1/3 Na-CMC/WS2 dispersion was selected as the reference solution for subsequent electrode preparation and evaluation of its electrocatalytic performance in the HER under alkaline conditions.
Upon integration with PANI, a stable P–N heterojunction forms between the N-type WS2 nanosheets and the P-type polymer. This interfacial architecture is expected to improve electron mobility, facilitate transfer of charge between PANI and WS2, and improve WS2 structural stability in alkaline environments, ultimately boosting the overall HER performance. Therefore, we fabricated Na-CMC/WS2/PANI/NF electrodes for the HER and investigated their electrochemical performance. Specifically, the designed composite electrode was constructed on a NF (1.0 cm2 area) substrate via EP for 40 minutes.26 Subsequently, the Tafel slopes and resistance of the resulting electrodes were evaluated using LSV and EIS in 1.0 M KOH. The HER performance of these electrodes was evaluated by LSV and EIS in 1.0 M KOH. As depicted in Fig. 1a–c, the developed electrode displayed reduced overpotential (η10, 141.6 mV) and a Tafel slope (81.5 mV dec−1) with a resistance of 17.6 Ω compared to blank nickel foam and PANI@NF electrodes. These findings confirm that combining N-type WS2 nanosheets with P-type PANI effectively enhances the performance in the HER. Nevertheless, the HER performance of these electrodes in alkaline conditions is still significantly lower than that of the benchmark electrocatalyst (Pt/C), which exhibits an η10 of 21.8 mV with a Tafel slope of 45.3 mV dec−1 and a resistance of 0.8 Ω. Similar to the results obtained under acidic conditions, the EP-treated electrode showed an analogous trend in the alkaline environment.26 This can be attributed to the interfacial incompatibility between the PANI and the exfoliated WS2, which adversely affects the overall efficiency of HER catalysis.
To enhance the interfacial compatibility between the components, the fabricated electrode was treated via electrochemical activation at 500 mA cm−2.26 As depicted in Fig. S2a, the negative potential steadily reduced throughout the 12-h monitoring time and stabilized at approximately −0.4 V after around 2 h, suggesting that the high current density effectively promoted interfacial interactions between the components and led to the construction of high-density P–N heterojunctions. This interface is expected to enhance charge transfer and thereby improve the overall electrocatalytic performance. Furthermore, the electrocatalytic performance of the developed electrodes after varying periods of EA treatment was further evaluated using LSV and EIS. As depicted in Fig. S2b–d, electrodes subjected to varying periods of EA treatment exhibited similar electrochemical activity. For example, following a 4-h EA treatment, the electrode exhibited an η10 of 24.5 mV with a Tafel slope of 48.2 mV dec−1, and a resistance of 0.5 Ω. Moreover, compared to the results achieved following EP treatment, the EA technique considerably boosted the catalytic activity of the exfoliated WS2 nanosheets. Based on these findings and considering the optimized processing conditions, the fabricated electrode treated by EA for 4 h was selected for further investigation of its HER performance.
Further comparison with the Pt/C catalyst revealed that the EA-treated electrode (treated for 4 h) displayed comparable overpotential, Tafel slope, and resistance (Fig. 1a–c). This finding demonstrates that EA treatment boosts both HER activity and WS2 stability in alkaline media through the generation of robust P–N heterojunctions. To further elucidate how EA treatment influences the electrocatalytic properties of the fabricated electrode, we investigated the changes in its electrochemically active surface area and surface morphology before and after EA treatment of the designed electrode. The Na-CMC/WS2/PANI/NF electrodes showed Cdl values of 1.76 mF cm−2 after EP treatment and 5.10 mF cm−2 after EA treatment, as indicated in Fig. 1d and S3. The Cdl of the EA-treated electrode is approximately threefold higher than that of the EP-treated electrode. This enhancement confirms that EA treatment effectively increases the number of active catalytic sites and surface area, thereby significantly improving HER electrocatalytic performance.35 In terms of surface morphology, SEM images revealed that the EA-treated electrode exhibited a uniform and textured surface morphology (Fig. 1e–g), in contrast to the rough and cracked surface of the electrode after EP treatment (Fig. S4). This indicates that the EA strategy facilitates improved interaction between PANI and WS2 nanosheets, leading to the creation of stable P–N heterojunctions at their interface and thereby enhancing interfacial compatibility.
Furthermore, EDX spectroscopic examination, as illustrated in Fig. 1h, verified the presence of all anticipated elemental components on the surface of the activated Na-CMC/WS2/PANI/NF electrode. The final electrode surface was found to contain 1.1 wt% of WS2, as summarized in the table adjacent to Fig. 1i. Overall, these findings indicate that the EA treatment and the presence of P–N heterojunctions offer a strategy to bolster the structural integrity of WS2 in alkaline conditions, substantially elevate its electrocatalytic activity, and consequently deliver exceptional alkaline HER performance.
Next, we surveyed recent WS2-based electrocatalytic systems and summarized their reported HER performance. Table S1 clearly indicates that the EA-treated fabricated electrode exhibits a notably reduced overpotential and Tafel slope for HER electrocatalysis compared to various WS2-based systems,28,36–41 even with differences in substrate materials. This confirms that the presence of the P–N heterojunction interfaces in the electrode matrix not only greatly enhances the electrocatalytic activity of WS2 but also effectively mitigates its intrinsic structural instability in alkaline environments17,18,32 and may ultimately enable its potential extension to practical alkaline water electrolysis applications. Therefore, the strategy of constructing the heterojunction interfaces presents a viable approach for developing highly effective and durable HER electrocatalysts in acidic as well as alkaline environments.26
Subsequent to confirming the electrocatalytic efficacy of the electrochemically activated electrode under alkaline conditions, we further evaluated its long-term stability by performing chronopotentiometry in 1 M KOH. The electroactivated electrode was operated at a fixed current density of 100 mA cm−2 for 24 h to evaluate its electrocatalytic durability. As depicted in Fig. 2a–d, the LSV curves, Tafel slopes, and resistance values remained virtually unchanged after continuous operation at a current density of 100 mA cm−2 for 24 h, indicating excellent durability. This result demonstrates the excellent structural stability of the EA-treated electrode enables it to maintain stable HER catalytic performance over time. Similar results were observed in a long-term CV stability evaluation. As shown in Fig. 2e–h, the activated electrode exhibited negligible variations in electrocatalytic activity after 1000 consecutive CV cycles. This clearly indicates that the P–N heterojunction interface within the composite electrode matrix is critical for maintaining the structural stability of WS2 nanosheets, which in turn ensures consistently efficient HER performance in alkaline environments.42 In contrast, the electrocatalytic activity of Pt/C for the HER is substantially diminished in alkaline environments. After either a continuous 24-h test at 100 mA cm−2 (Fig. 3a–d) or following 1000 CV cycles (Fig. 3e–h), Pt/C exhibited significantly deteriorated performance in 1 M KOH compared to its initial state. Specifically, the overpotential of Pt/C at −10 mA cm−2 increased to 30.2 mV after the 24-h test (Fig. 3b) and rose further to 35.8 mV after 1000 CV cycles (Fig. 3f). In addition, the Tafel slope of Pt/C increased to 71.8 mV dec−1 after the 24-h test (Fig. 3c) and reached 93.0 mV dec−1 following the CV cycling (Fig. 3g). The decline in the electrocatalytic efficiency of Pt/C under alkaline conditions can be attributed to several key factors. Firstly, in alkaline media, platinum tends to undergo oxidation, which leads to the formation of surface species such as platinum oxide (PtO) or platinum hydroxide [Pt(OH)2]. This process reduces the number of active catalytic sites and alters the electronic structure of platinum and ultimately decreases its activity in reactions such as hydrogen evolution and oxygen reduction.43 Secondly, under alkaline conditions and at high operating potentials, the carbon support gradually undergoes degradation. This degradation leads to the detachment and loss of the Pt nanoparticles dispersed on the carbon matrix and thereby reduces the overall active surface area and compromises catalytic performance.44 Finally, in alkaline environments, due to their high concentration, hydroxide ions (OH−) tend to adsorb onto the surface of Pt and block the active sites.3,45 This not only reduces catalytic activity but also promotes the formation of large Pt clusters and aggregates.46,47 Based on the discussion above, our findings readily confirm that the electroactivated electrode demonstrates excellent electrocatalytic stability under alkaline conditions and holds significant promise as a viable alternative to noble metal Pt/C catalysts, achieving the desired performance for hydrogen production via water electrolysis. The above results inspired our curiosity to further investigate the structural attributes, surface morphology, and elemental distribution of the electrodes after the long-term stability tests.
Raman spectroscopy analysis (Fig. S5) revealed that the exfoliated WS2 nanosheets in the EA-treated electrode retained their structural integrity after testing at a current density of 100 mA cm−2 and 1000 cycles of CV, compared to the pristine EA-treated electrode, indicating good structural stability. This result suggests that the exfoliated WS2 nanosheets exhibit excellent enduring catalytic stability and resilience to alkaline environments, which can be attributed to the presence and protective effect of the P–N heterojunctions within the electrode. The same outcomes were also noted in SEM measurements. The surface and microstructural morphologies of the EA-treated composite electrode remained largely unchanged after 24 h of continuous process at 100 mA cm−2 (as depicted in Fig. 4a–e) and after 1000 CV cycles (Fig. 4m–q), when compared to their pristine states (Fig. 1e–i). Furthermore, all characteristic elemental components remained present in the structure (Fig. 4e and q), further indicating the excellent structural stability of the EA-treated electrode. It is worth noting that, compared to the pre-test results (right table in Fig. 1i), the WS2 content (wt%) of the EA-treated electrode did not significantly change after both stability tests (top-right tables in Fig. 4e and q). This finding reveals that the WS2 nanosheets remained stably anchored to the electrode surface, even after prolonged exposure to high current density or continuous CV cycling. Moreover, elemental mapping images, acquired via EDX integrated with SEM, further corroborated the presence and uniform distribution of all distinctive elements across the surface of the EA-treated electrodes following both stability tests (Fig. 4f–l, and r–x). This further supports our observation that the presence of P–N heterojunctions within the electrode matrix not only contributes to the structural integrity of the electrode but also facilitates charge flow from PANI to the WS2 structure throughout the HER process. Consequently, WS2 can stably and efficiently catalyze hydrogen evolution, ultimately achieving the desired performance for water electrolysis in an alkaline environment.
Supplementary information (SI): includes supplementary characterization and electrochemical data. See DOI: https://doi.org/10.1039/d5ra08680a.
| This journal is © The Royal Society of Chemistry 2025 |