Next Article in Journal
Latest Nanoparticles to Modulate Hypoxic Microenvironment in Photodynamic Therapy of Cervical Cancer: A Review of In Vivo Studies
Previous Article in Journal
The Identification of Gyrophoric Acid, a Phytochemical Derived from Lichen, as a Potent Inhibitor for Aggregation of Amyloid Beta Peptide: In Silico and Biochemical Evaluation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Boosted Hydrogen Evolution Catalysis Using Biomass-Derived Mesoporous Carbon Nanosponges

1
Division of System Semiconductor, Dongguk University, Seoul 04620, Republic of Korea
2
Quantum-Functional Semiconductor Research Center, Dongguk University, Seoul 04620, Republic of Korea
3
Department of Chemistry, CMS College of Engineering, Ernapuram, Namakkal 637003, Tamil Nadu, India
4
Department of Mechanical Engineering, K. Ramakrishnan College of Technology, Trichy 621112, Tamil Nadu, India
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2025, 26(17), 8502; https://doi.org/10.3390/ijms26178502
Submission received: 20 July 2025 / Revised: 26 August 2025 / Accepted: 28 August 2025 / Published: 1 September 2025
(This article belongs to the Special Issue Advances in Electrochemical Nanomaterials for Energy and Catalysis)

Abstract

Carbon-based metal-free catalysts, particularly those such as biomass-derived mesoporous activated carbon (AC) nanostructures, hold great promises for cost-effective and sustainable electrocatalysis for enhancing hydrogen evolution reaction (HER) performance in green energy technology. Neem and ginkgo leaves are rich in bioactive compounds and self-doping heteroatoms with naturally porous structures and act as a low-cost, sustainable biomass precursors for high-performance HER catalysts. In this study, mesoporous AC nanoflakes and nanosponges were synthesized using biomass precursors of neem and ginkgo leaves through a KOH activation process. Notably, AC nanosponges derived from ginkgo leaves exhibited outstanding physicochemical characteristics, including a sponge-like porous morphology with a large specific surface area of 1025 m2/g. For electrochemical evaluation in 0.5 M H2SO4, the G-AC sample revealed superior electrocatalytic HER performance, with a remarkably low overpotential of 26 mV at −10 mA/cm2, a small Tafel slope of 24 mV/dec, and long-term durability over 30 h. These results depict biomass-derived mesoporous AC nanosponges to hold substantial potential for highly efficient hydrogen production, contributing significantly to the advancement of eco-friendly energy solutions.

1. Introduction

The urgent need for renewable energy sources is critical in addressing the escalating energy crisis and associated environmental challenges [1,2,3,4,5,6,7,8]. Green hydrogen has attracted considerable attention as a promising alternative energy source due to its environmental sustainability, high energy density, and zero carbon emissions [9,10,11]. Among various hydrogen production methods, electrocatalytic water splitting is regarded as the most effective, environmentally benign, and scalable approach for generating green hydrogen via the hydrogen evolution reaction (HER) [12,13,14,15]. As a key half-reaction in the water electrolysis process, HER offers a practical and cost-effective route for large-scale hydrogen production, making it a viable solution to both energy demands and environmental concerns [16,17,18]. However, the sluggish reaction kinetics and substantial overpotential associated with HER substantially limit the catalytic efficiency of water electrolysis systems [19,20,21,22]. Although platinum (Pt)-based materials persist among the highly efficient HER catalysts, their high cost and limited availability severely constrain their large-scale application [23,24]. Consequently, the development of low-cost, metal-free, acid-stable, and highly active HER catalysts derived from abundant natural resources is essential for advancing sustainable hydrogen production technologies.
Recently, carbonaceous materials and their nanostructures have attracted significant attention as HER electrocatalysts owing to their excellent electronic conductivity, chemical stability, and capacity to host electrochemically active species [25,26,27]. Among these, biomass-derived activated carbon (AC) has emerged as a promising candidate owing to its natural abundance, low cost, high conductivity, large surface area, tunable porous structure, and excellent durability [28,29,30,31]. As a result, substantial efforts have been devoted to synthesizing carbon nanomaterials from abundant biomass sources and waste materials to develop efficient carbon-based catalysts for energy conversion applications [32,33,34]. For example, Han et al. [35] prepared N-doped porous carbon fibers from natural cattail fiber, attaining a low overpotential of 244 mV and a Tafel slope of 70 mV/dec. Prabu et al. [36] synthesized hierarchical porous carbon from palm plants, recording an overpotential of 330 mV and a Tafel slope of 63 mV/dec. Similarly, Cao et al. [37] developed nitrogen-doped porous carbon from bean sprouts using the SiO2 template technique, obtaining an overpotential of 413 mV and a Tafel slope of 98 mV/dec. Additionally, Prabu et al. [38] derived nanoporous activated carbon sheets from Ooty Varkey food waste, demonstrating excellent HER activity with a low overpotential of 380 mV at 10 mA/cm2. Arul Saravanan et al. [39] prepared AC nanosheets from peanut shells via KOH activation, demonstrating outstanding HER performance with a notably low overpotential of 80 mV at 10 mA/cm2 and a Tafel slope of 75 mV/dec. Hoang et al. [40] synthesized Ni- and P-doped carbon from carrots using a facile one-step pyrolysis method, achieving an HER overpotential of 939 mV with a Tafel slope of 273 mV/dec. Among various biomass resources, neem leaves and ginkgo leaves stand out as promising sustainable precursors for producing AC nanostructures due to their natural richness, low cost, high carbon content, and eco-friendliness [41,42,43,44,45,46]. Utilizing these biomass sources offers an effective route for fabricating high-performance AC nanostructures. Rich in bioactive compounds (i.e., cellulose, protein, carbohydrates, calcium, vitamin, carotene, and phosphorous), self-doping heteroatoms (i.e., O, N, S) characteristics, high-reaction-active sites, unique leaf structures, and huge availability, and the fact that they are characterized by high surface areas and large porous natures, means that neem and ginkgo leaves are particularly well-suited for the fabrication of cost-effective biomass catalysts with high HER performance, and thus provide an environmentally sustainable platform for efficient electrocatalyst development. Despite these advantages, no studies to date have reported the use of neem leaf- and ginkgo leaf-derived mesoporous AC (N-AC and G-AC) nanostructures specifically as HER electrocatalysts.
Motivated by the aforementioned background, we synthesized mesoporous N-AC nanoflakes and G-AC nanosponges via the KOH activation method and evaluated their HER performances. Notably, the G-AC catalyst exhibited an impressively lower overpotential of 26 mV and a small Tafel slope of 24 mV/dec at 10 mA/cm2 in 0.5 M H2SO4. This work presents a comprehensive examination of the catalyst synthesis, material characterization, and electrocatalytic HER activities of the fabricated N-AC nanoflakes and G-AC nanosponges.

2. Results and Discussion

Figure 1 displays the surface morphology of the N-AC and G-AC nanostructures derived from biomass neem leaves and ginkgo leaves, respectively, via the KOH activation method. The N-AC sample exhibits an irregularly stacked and aggregated layered flake-like structure (Figure 1a,b), while the G-AC sample displays an irregular, three-dimensional porous network with a sponge-like morphology (Figure 1c,d). Compared to N-AC, the G-AC structure offers greater porosity, a higher surface area, and improved accessibility to active sites, all of which are advantageous for enhancing HER activity.
The structural properties of the N-AC and G-AC nanostructures were analyzed using XRD. Figure 2a shows the XRD patterns of both samples, which reveal two characteristic diffraction peaks at 23.6° and 44.3°, corresponding to the (002) and (100) planes of disordered carbon, respectively, indicating their amorphous nature [38,47,48]. Compared to N-AC, the G-AC sample exhibits a more intense (002) peak, suggesting a lower degree of structural disorder and a higher pore density [49,50]. This improved graphitization contributes to enhanced electrical conductivity in the G-AC nanostructure. No secondary phases were detected, confirming the high purity of both samples. The degree of graphitization and structural disorder in the carbonaceous materials was further examined via Raman spectroscopy. Figure 2b presents the Raman spectra of N-AC and G-AC, both of which show three prominent peaks at 1351 cm−1 (D band), 1603 cm−1 (G band), and 2853 cm−1 (2D band), indicating the graphitic characteristics of the synthesized carbon materials [13,31,35,51]. The D band arises from the A1g vibrational mode, associated with structural defects and disordered vibrations at the edges of graphitic domains [52]. The G band corresponds to the E2g mode of sp2-hybridized carbon atoms, characteristic of an ordered graphitic lattice [53]. The presence of the 2D band serves as a distinguishing feature of activated carbon [54]. The intensity ratios of the D and G bands (ID/IG) were calculated as 0.97 for N-AC and 0.94 for G-AC, indicating that G-AC has a lower degree of disorder and a higher level of graphitization compared to N-AC.
The specific surface area and porosity of the N-AC and G-AC nanostructures were analyzed using the BET and BJH techniques. Figure 2c shows the N2 adsorption–desorption isotherm curves for both samples, which exhibit Type IV physisorption isotherms with Type H4 hysteresis loops (as classified by IUPAC), confirming their mesoporous nature [36,48,55,56]. BET analysis exhibited surface areas of 433 m2/g for N-AC and 1025 m2/g for G-AC. As illustrated in Figure 2d, the pore surface areas were determined to be 164 m2/g for N-AC and 360 m2/g for G-AC. Furthermore, G-AC exhibited a higher total pore volume (0.5805 cm3/g) compared to N-AC (0.3122 cm3/g). The average pore sizes were determined to be 2.88 nm for N-AC and 2.26 nm for G-AC. The smaller average pore size, along with the significantly higher surface area and pore volume of G-AC, clearly indicates its superior mesoporous structure, which is critical for enhancing electrocatalytic HER efficiency.
The chemical bonding structures of the N-AC and G-AC were investigated by XPS measurements. The XPS full survey spectra of the N-AC and G-AC clearly reveal the presence of their constituent elements such as C and O (Figure 3a). The C1s core-level spectra of both samples (Figure 3b,c) exhibited three characteristic peaks at 284.8, 286.3, and 289.5 eV, corresponding to the C–C (sp2-hybridized domains), C–O (epoxy, hydroxyl), and C=O (carboxyl) functional groups, respectively [57,58]. Similarly, the O1s core-level spectra of the both samples (Figure 3d,e) also displayed three prominent peaks at 531.6, 533.8, and 536.2 eV, ascribed to the C–O–C, C=O, and O–C=O functional groups, respectively [59,60]. The abundance of these oxygen-containing functional groups can significantly alter the surface polarity of the carbon material, thereby enhancing the wettability at the electrode–electrolyte interface. This improved interfacial interaction is expected to facilitate more efficient ion transport, ultimately contributing to enhanced electrochemical performance of the catalyst.
After confirming the structural characteristics of the fabricated catalysts, the electrocatalytic HER performance of N-AC and G-AC was systematically investigated. Figure 4a,b present the CV curves of the N-AC and G-AC catalysts recorded at scan rates ranging from 10 to 100 mV/s. Both catalysts exhibited typical rectangular-shaped CV profiles, indicating efficient electrical double-layer capacitive behavior [13,37,61,62]. As the scan rate amplified, the current density also rose, suggesting small diffusion resistance of the active materials. Notably, G-AC displayed a larger CV profile area and a higher current response compared to N-AC, implying a larger number of accessible active sites. The electrocatalytic HER activity was closely associated with the Cdl and ECSA of the catalysts. The Cdl values for N-AC and G-AC were determined to be 3.66 and 5.29 mF/cm2, respectively, from the non-Faradaic CV region at 0.07 V (see Figure 4c–f). Using Equations (9) and (10), the corresponding ECSA values were calculated as 105 cm2 for N-AC and 151 cm2 for G-AC. The significantly higher Cdl and ECSA values of the G-AC catalyst indicate a larger number of exposed active sites and enhanced electrical conductivity compared to the N-AC catalyst.
The HER performance of the N-AC and G-AC catalysts was evaluated using LSV at a scan rate of 5 mV/s in 0.5 M H2SO4. Figure 5a presents the iR-corrected and without-iR-correction LSV curves (Figure S1) of the N-AC and G-AC catalysts. The η at a current density of −10 mA/cm2 was determined to be 40 mV for N-AC and 26 mV for G-AC, as calculated using Equations (11) and (12). The lower η value of G-AC is attributed to its enhanced electrical conductivity and greater number of catalytically active sites (i.e., higher ECSA) compared to N-AC. Figure 5b displays the Tafel curves of the N-AC and G-AC catalysts. The ST values were determined to be 46 mV/dec for N-AC and 24 mV/dec for G-AC using Equation (13). The lower ST value of G-AC indicates more favorable and efficient HER kinetics, consistent with the Volmer–Heyrovsky mechanism. Consequently, it is extensively accepted that the HER mechanism on the cathode surface proceeds through a sequence of electrochemical steps. In an acidic medium, this multistep process typically follows the reactions outlined below [63,64,65]:
H 3 O +   +   M   +   e M     H   +   H 2 O   ( V o l u m e r )
M     H   +   H 3 O +   +   e H 2   +   H 2 O   +   M   ( H e y r o v s k y )
2 M     H 2 M   +   H 2   ( T a f e l )
where M denotes a vacant active site on the catalyst surface, while M–H refers to an adsorbed hydrogen species. The strength of the M–H bond plays a critical role in determining the HER kinetics of catalyst materials. In acidic solutions, the Volmer reaction (Equation (1)) involves an initial discharge of the hydronium ion and the formation of hydrogen intermediates (i.e., M-H), and the subsequent formation of H2 involves the electrochemical Heyrovsky step (Equation (2)) and the chemical Tafel step (Equation (3)). Therefore, the Volmer–Heyrovsky pathway is generally more efficient and favorable for HER activity. Compared to other electrocatalysts, the G-AC catalyst demonstrated superior HER performance, evidenced by lower η and ST values, due to its surface oxygen groups, porous nature, larger ECSA, and improved electrical conductivity that enhanced the wettability, defect density, and charge transfer of the G-AC (Table 1).
The excellent HER performance of the N-AC and G-AC catalysts was further validated through CP measurements. As shown in Figure 5c, the G-AC catalyst consistently exhibited lower overpotentials at all applied current densities (−10, −20, −30, −40, −50, and −100 mA/cm2) than N-AC, confirming its higher catalytic efficiency. Figure 5d presents the long-term HER stability test of both catalysts at −10 mA/cm2 over 30 h. The G-AC catalyst exhibited excellent stability throughout the test duration, which can be attributed to its higher ECSA, greater porosity, and lower internal resistance. Furthermore, the LSV curves before and after the long-term stability test showed negligible deviation (Figure 5e,f), indicating microstructural and electrochemical robustness. These findings designate G-AC as a highly stable and efficient electrocatalyst for HER. Following the HER stability test, FE-SEM analysis was conducted to assess any microstructural changes in the catalysts. The N-AC catalyst displayed an aggregated flake-like structure (inset of Figure 5e), whereas the G-AC catalyst retained its original agglomerated sponge-like morphology (inset of Figure 5f). Furthermore, Raman measurements (Figure S2) showed that both catalysts retained their original features, indicating the high stability of the materials.
To further investigate the enhanced HER kinetics of the G-AC catalyst, EIS analysis was conducted. Figure 6 displays the Nyquist plots of the N-AC and G-AC catalysts, along with the corresponding equivalent circuit (inset). Both catalysts exhibited a linear region at low frequencies, indicating effective ion diffusion across the catalyst surface [13,39,71]. The absence of a semicircular region in both cases suggests high ionic diffusivity and excellent electronic conductivity [66,67,72]. The series resistance (Rs) represents the combined resistance of the electrolyte, electrode, and interfaces, and a lower Rs indicates easier charge transport and reduced energy loss, leading to improve HER performance. Using the equivalent circuit model, the Rs values were calculated to be 0.94 Ω for N-AC and 0.83 Ω for G-AC. The lower Rs value and steeper Nyquist slope of the G-AC catalyst are attributed to its higher density of active sites, greater porosity, enhanced ion transport, and improved electrical conductivity. However, the post-stability EIS results show lower charge transfer resistance, confirming the catalyst’s long-term stability due to surface restructuring and improved electrode–electrolyte interaction. These results highlight the strong potential of the biomass-derived G-AC nanostructures from ginkgo leaves as sustainable and efficient HER electrocatalysts for future green hydrogen production.

3. Materials and Methods

3.1. Materials

All chemicals were purchased from Sigma-Aldrich (Seoul, Republic of Korea) and used without further purification. Neem leaves (Azadirachta indica) were collected from Tholudur, Tamil Nadu, India, and ginkgo leaves (Ginkgo biloba) were obtained from Seoul, Republic of Korea.

3.2. Preparation of Activated Carbon Nanostructures

Figure 7 presents a schematic illustration of the fabrication process for N-AC nanoflakes (Figure 7a) and G-AC (Figure 7b) nanosponges derived from biomass neem leaves and ginkgo leaves, respectively. Mesoporous AC nanoflakes and nanosponges were synthesized from these two biomass sources via the KOH activation method. Initially, the raw leaves were separated, thoroughly washed with deionized (DI) water to remove surface impurities, and sun-dried for seven days. The dried neem leaves (NL) and ginkgo leaves (GL) were then carbonized at 300 °C in an air atmosphere for 1 h to obtain carbonized ashes, referred to as NLAs and GLAs, respectively. Following carbonization, 3 g of each carbonized ash was mixed with 12 g of KOH using a mortar and subsequently annealed at 700 °C in an air atmosphere for 2 h. During KOH activation, oxygen- and carbon-containing functional groups on the carbon precursor reacted with KOH, resulting in the formation of various carbonaceous species such as potassium carbonate (K2CO3) and carbon monoxide (CO). This high-temperature activation process can be represented by the following chemical reactions [13,73,74]:
2 C N L A s   o r   G L A s   +   6 K O H 3 H 2   +   2 K + 2 K 2 C O 3
K 2 C O 3 K 2 O   +   C O 2
C O 2   +   C 2 C O
K 2 C O 3   +   2 C 2 K + 3 C O
K 2 O   +   C 2 K   +   C O
After activation, the resulting colloidal suspension was collected and stirred with DI water for 8 h to eliminate unreacted potassium complexes. The resulting mixture was then filtered, rinsed with DI water, and dried at 120 °C for 10 h. The AC nanopowders derived from neem leaves and ginkgo leaves are denoted as N-AC and G-AC, respectively.

3.3. Material Characterizations

The microstructures of the fabricated N-AC and G-AC catalysts were examined using field-emission scanning electron microscopy (FE-SEM). Their amorphous nature and vibrational properties were characterized via X-ray diffraction (XRD) and Raman spectroscopy, respectively. Pore size distribution and textural characteristics were analyzed using the Brunauer–Emmett–Teller (BET) and Barrett–Joyner–Halenda (BJH) methods.

3.4. Electrocatalytic HER Measurements

The electrocatalytic HER performance of the N-AC and G-AC catalysts was evaluated using an electrochemical workstation configured in a standard three-electrode setup. For electrode preparation, either N-AC or G-AC was dispersed in N-methyl-2-pyrrolidone to form a homogeneous slurry, which was subsequently coated onto stainless-steel substrates (1 cm2) and dried at 160 °C for 6 h. A coiled Pt wire and a saturated calomel electrode (SCE) were employed as the counter and reference electrodes, respectively. All HER measurements—including cyclic voltammetry (CV), linear sweep voltammetry (LSV), electrochemical impedance spectroscopy (EIS), and chronopotentiometry (CP)—were conducted in 0.5 M H2SO4. LSV was performed within the potential range of 0.1 to −1.2 V at a scan rate of 5 mV/s. All LSV measurements were calculated with 100% iR compensation to ensure accurate potential values. CV measurements were carried out at scan rates ranging from 10 to 100 mV/s within a 0–1.0 V potential window. HER rate performance was evaluated by CP at current densities of −10, −20, −30, −40, −50, and −100 mA/cm2 for 10 min each. EIS analysis was conducted over a frequency range of 1 Hz to 10 kHz. The electrochemical double-layer capacitance (Cdl) and electrochemically active surface area (ECSA) of the catalysts were calculated from the non-Faradaic CV region using the following equations [75,76,77]:
J D L   =   C d l   ×   v / A
E C S A   =   C d l / C e
where Cdl is the non-Faradaic capacitance, JDL is the non-Faradaic current density, A is the electrode area, v is the scan rate, and Ce is the specific capacitance of the electrolyte, taken as 0.035 mF/cm2 for 0.5 M H2SO4 [78]. The overpotential (η) and Tafel slope (ST) values for the N-AC and G-AC catalysts were determined using the subsequent equations [19,35,39,79]:
E R H E   =   E S C E   +   0.059 · p H   +   E S C E 0
E R H E   =   η
η   =   S T log ( J )   +   c
where J represents the current density, ERHE is the potential vs. the reversible hydrogen electrode (RHE), c is a fitting constant, and E S C E 0 signifies the typical potential of the SCE.

4. Conclusions

The mesoporous N-AC and G-AC nanostructures were successfully synthesized from two different biomass sources—neem and ginkgo leaves—via the KOH activation process. The N-AC sample exhibited a stacked and aggregated layered flake morphology, whereas the G-AC sample displayed a three-dimensional porous sponge-like morphology with an enhanced specific surface area. Owing to its high specific surface area, substantial mesoporosity, and superior graphitization, the G-AC catalyst achieved a low overpotential of 26 mV at −10 mA/cm2 in 0.5 M H2SO4. Furthermore, the G-AC catalyst exhibited a small Tafel slope of 24 mV/dec and demonstrated excellent long-term durability. These results specify that porous G-AC nanosponges are auspicious candidates for high-performance HER electrocatalysts in future green hydrogen production.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms26178502/s1.

Author Contributions

S.S. (Sankar Sekar): Methodology, Formal analysis, Investigation, Writing—original draft. S.S. (Sutha Sadhasivam): Formal analysis, Investigation. A.S.: Methodology, Formal analysis. S.S. (Saravanan Sekar): Formal analysis, Investigation. Y.L.: Data curation, Validation, Supervision, Writing—review and editing. S.L.: Conceptualization, Supervision, Validation, Writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Research Foundation (NRF) of Korea through the basic science research program (RS-2023-NR076644) funded by the Korean Government.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Materials. Further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Liu, S.; Wang, J.; Yao, Z.; Xia, Q.; Xu, H.; Guo, W.; Wu, X. Enhanced oxygen activation in antibiotic degradation using N-doped carbon skeleton encapsulated high-entropy alloy catalysts. Chem. Eng. J. 2025, 518, 164832. [Google Scholar] [CrossRef]
  2. Li, H.; Zhang, G.; Jia, X.; Pan, Y.; Yang, Z.; Li, G.; Guan, T.; Wang, K. Chemical vapor deposition-assisted activation for tailoring mesoporous carbon from polar components of coal tar pitch for zinc-ion hybrid capacitors. Chem. Eng. J. 2025, 519, 164974. [Google Scholar] [CrossRef]
  3. Li, W.; Liu, Y.; Wu, M.; Feng, X.; Redfern, S.A.T.; Shang, Y.; Yong, X.; Feng, T.; Wu, K.; Liu, Z.; et al. Carbon-Quantum-Dots-Loaded Ruthenium Nanoparticles as an Efficient Electrocatalyst for Hydrogen Production in Alkaline Media. Adv. Mater. 2018, 30, 1800676. [Google Scholar] [CrossRef] [PubMed]
  4. Zhang, M.; Guan, J.; Tu, Y.; Chen, S.; Wang, Y.; Wang, S.; Yu, L.; Ma, C.; Deng, D.; Bao, X. Highly efficient H2 production from H2S via a robust graphene-encapsulated metal catalyst. Energy Environ. Sci. 2020, 13, 119–126. [Google Scholar] [CrossRef]
  5. Jamesh, M.-I.; Hu, D.; Wang, J.; Naz, F.; Feng, J.; Yu, L.; Cai, Z.; Colmenares, J.C.; Lee, D.-J.; Chu, P.K.; et al. Recent advances in noble metal-free electrocatalysts to achieve efficient alkaline water splitting. J. Mater. Chem. A 2024, 12, 11771–11820. [Google Scholar] [CrossRef]
  6. Sekar, S.; Lee, E.; Yun, J.; Arumugasamy, S.K.; Choi, M.-J.; Lee, Y.; Lee, S. High-performance electrocatalyst of activated carbon-decorated molybdenum trioxide nanocomposites for effective production of H2 and H2O2. Sep. Purif. Technol. 2025, 361, 131614. [Google Scholar] [CrossRef]
  7. Sekar, S.; Sadhasivam, S.; Lee, Y.; Lee, S. Enhanced bifunctional water electrolysis activities of activated carbon-decorated trimetallic Ni-Al-La nanocomposites. Appl. Surf. Sci. 2025, 698, 163098. [Google Scholar] [CrossRef]
  8. Sekar, S.; Ilanchezhiyan, P.; Ahmed, A.T.A.; Lee, Y.; Lee, S. Substantial Electrocatalytic Oxygen Evolution Performances of Activated Carbon-Decorated Vanadium Pentoxide Nanocomposites. Int. J. Energy Res. 2024, 2024, 9953038. [Google Scholar] [CrossRef]
  9. Sadeq, A.M.; Homod, R.Z.; Hussein, A.K.; Togun, H.; Mahmoodi, A.; Isleem, H.F.; Patil, A.R.; Moghaddam, A.H. Hydrogen energy systems: Technologies, trends, and future prospects. Sci. Total Environ. 2024, 939, 173622. [Google Scholar] [CrossRef]
  10. Hossain Bhuiyan, M.M.; Siddique, Z. Hydrogen as an alternative fuel: A comprehensive review of challenges and opportunities in production, storage, and transportation. Int. J. Hydrogen Energy 2025, 102, 1026–1044. [Google Scholar] [CrossRef]
  11. Segovia-Hernández, J.G.; Hernández, S.; Cossío-Vargas, E.; Juarez-García, M.; Sánchez-Ramírez, E. Green hydrogen production for sustainable development: A critical examination of barriers and strategic opportunities. RSC Sustain. 2025, 3, 134–157. [Google Scholar] [CrossRef]
  12. Abdullah, A.; Tariq, F.; Kulkarni, M.A.; Thaalbi, H.; Din, H.U.; Kang, S.H.; Ryu, S.-W. High-efficiency electrocatalytic hydrogen generation under harsh acidic condition by commercially viable Pt nanocluster-decorated non-polar faceted GaN nanowires. Int. J. Hydrogen Energy 2024, 94, 1257–1265. [Google Scholar] [CrossRef]
  13. Sekar, S.; Shanmugam, A.; Senthilkumar, G.; Thangasami, K.; Jung, H.; Lee, Y.; Lee, S. Enhanced Hydrogen Evolution Reaction Using Biomass-Activated Carbon Nanosheets Derived from Eucalyptus Leaves. Materials 2025, 18, 670. [Google Scholar] [CrossRef] [PubMed]
  14. Gao, X.; Chen, Y.; Wang, Y.; Zhao, L.; Zhao, X.; Du, J.; Wu, H.; Chen, A. Next-Generation Green Hydrogen: Progress and Perspective from Electricity, Catalyst to Electrolyte in Electrocatalytic Water Splitting. Nano-Micro Lett. 2024, 16, 237. [Google Scholar] [CrossRef]
  15. Sekar, S.; Shanmugam, A.; Lee, Y.; Lee, S. Highly Efficient Electrocatalyst of 2D–2D gC3N4–MoS2 Composites for Enhanced Overall Water Electrolysis. Materials 2025, 18, 3775. [Google Scholar] [CrossRef]
  16. Sekar, S.; Sim, D.H.; Lee, S. Excellent Electrocatalytic Hydrogen Evolution Reaction Performances of Partially Graphitized Activated-Carbon Nanobundles Derived from Biomass Human Hair Wastes. Nanomaterials 2022, 12, 531. [Google Scholar] [CrossRef]
  17. Li, X.; Zhao, L.; Yu, J.; Liu, X.; Zhang, X.; Liu, H.; Zhou, W. Water Splitting: From Electrode to Green Energy System. Nano-Micro Lett. 2020, 12, 131. [Google Scholar] [CrossRef]
  18. Chatenet, M.; Pollet, B.G.; Dekel, D.R.; Dionigi, F.; Deseure, J.; Millet, P.; Braatz, R.D.; Bazant, M.Z.; Eikerling, M.; Staffell, I.; et al. Water electrolysis: From textbook knowledge to the latest scientific strategies and industrial developments. Chem. Soc. Rev. 2022, 51, 4583–4762. [Google Scholar] [CrossRef]
  19. Sekar, S.; Aqueel Ahmed, A.T.; Sim, D.H.; Lee, S. Extraordinarily high hydrogen-evolution-reaction activity of corrugated graphene nanosheets derived from biomass rice husks. Int. J. Hydrogen Energy 2022, 47, 40317–40326. [Google Scholar] [CrossRef]
  20. Veeramani, K.; Janani, G.; Kim, J.; Surendran, S.; Lim, J.; Jesudass, S.C.; Mahadik, S.; Lee, H.; Kim, T.-H.; Kim, J.K.; et al. Hydrogen and value-added products yield from hybrid water electrolysis: A critical review on recent developments. Renew. Sustain. Energy Rev. 2023, 177, 113227. [Google Scholar] [CrossRef]
  21. Wan, Z.; Yang, K.; Li, P.; Yang, S.; Wang, X.; Gao, R.; Xie, X.; Deng, G.; Yang, M.; Wang, Z. Atomically dispersed Au single sites and nanoengineered structural defects enable a high electrocatalytic activity and durability for hydrogen evolution reaction and overall urea electrolysis. Electrochim. Acta 2024, 499, 144685. [Google Scholar] [CrossRef]
  22. Li, P.; Luo, S.; Xiong, Z.; Xiao, H.; Wang, X.; Peng, K.; Xie, X.; Zhang, Z.; Deng, G.; Yang, M.; et al. RuxMoS2 interfacial heterojunctions achieve efficient overall water splitting and stability in both alkaline and acidic media under large current density exceeding 100 mA cm−2. Mol. Catal. 2025, 570, 114710. [Google Scholar] [CrossRef]
  23. Hussain, M.B.; Ahmad, M.; Cheng, X.; Mehmood, R.; Ajmal, Z.; Hussain, S.; Tayyab, M.; Gitis, V. ZrO2 supported Pt nanoparticles for robust electrocatalytic hydrogen evolution reactions. Int. J. Hydrogen Energy 2025, 106, 825–833. [Google Scholar] [CrossRef]
  24. Sekar, S.; Sadhasivam, S.; Nangai, E.K.; Saravanan, S.; Kim, D.Y.; Lee, S. Enhanced Hydrogen Evolution Reaction Performances of Ultrathin CuBi2O4 Nanoflakes. Int. J. Energy Res. 2023, 2023, 5038466. [Google Scholar] [CrossRef]
  25. Ahmed, K.; Hameed, S.; Patchigolla, K.; Dawood, N.; Ghouri, Z.K. Carbon-based electrocatalysts for hydrogen evolution reaction. Energy Convers. Manage. X 2025, 26, 100892. [Google Scholar] [CrossRef]
  26. Li, W.; Xu, Y.; Wang, G.; Xu, T.; Wang, K.; Zhai, S.; Si, C. Sustainable Carbon-Based Catalyst Materials Derived From Lignocellulosic Biomass for Energy Storage and Conversion: Atomic Modulation and Properties Improvement. Carbon Energy 2025, 7, e708. [Google Scholar] [CrossRef]
  27. Nemiwal, M.; Zhang, T.C.; Kumar, D. Graphene-based electrocatalysts: Hydrogen evolution reactions and overall water splitting. Int. J. Hydrogen Energy 2021, 46, 21401–21418. [Google Scholar] [CrossRef]
  28. Cui, W.; Liu, Q.; Cheng, N.; Asiri, A.M.; Sun, X. Activated carbon nanotubes: A highly-active metal-free electrocatalyst for hydrogen evolution reaction. Chem. Commun. 2014, 50, 9340–9342. [Google Scholar] [CrossRef]
  29. Liu, T.; Yabu, H. Biomass-Derived Electrocatalysts: Low-Cost, Robust Materials for Sustainable Electrochemical Energy Conversion. Adv. Energy Sustain. Res. 2024, 5, 2300168. [Google Scholar] [CrossRef]
  30. Feng, Y.; Jiang, J.; Xu, Y.; Wang, S.; An, W.; Chai, Q.; Prova, U.H.; Wang, C.; Huang, G. Biomass derived diverse carbon nanostructure for electrocatalysis, energy conversion and storage. Carbon 2023, 211, 118105. [Google Scholar] [CrossRef]
  31. Gao, J.; Chu, X.; Lu, H.; Wang, H.; Li, X.; Yin, Z.; Tan, X. Efficient carbon-based electrocatalyst derived from biomass for hydrogen peroxide generation. Mater. Today Commun. 2021, 26, 102051. [Google Scholar] [CrossRef]
  32. Cao, Y.; Sun, Y.; Zheng, R.; Wang, Q.; Li, X.; Wei, H.; Wang, L.; Li, Z.; Wang, F.; Han, N. Biomass-derived carbon material as efficient electrocatalysts for the oxygen reduction reaction. Biomass Bioenergy 2023, 168, 106676. [Google Scholar] [CrossRef]
  33. Tiwari, S.K.; Bystrzejewski, M.; De Adhikari, A.; Huczko, A.; Wang, N. Methods for the conversion of biomass waste into value-added carbon nanomaterials: Recent progress and applications. Prog. Energy Combust. Sci. 2022, 92, 101023. [Google Scholar] [CrossRef]
  34. Fahad Halim, A.F.M.; Poinern, G.E.J.; Fawcett, D.; Sharma, R.; Surendran, S.; Rajeshkannan, R. Biomass-Derived Carbon Nanomaterials: Synthesis and Applications in Textile Wastewater Treatment, Sensors, Energy Storage, and Conversion Technologies. CleanMat 2025, 2, 4–58. [Google Scholar] [CrossRef]
  35. Han, G.; Hu, M.; Liu, Y.; Gao, J.; Han, L.; Lu, S.; Cao, H.; Wu, X.; Li, B. Efficient carbon-based catalyst derived from natural cattail fiber for hydrogen evolution reaction. J. Solid State Chem. 2019, 274, 207–214. [Google Scholar] [CrossRef]
  36. Prabu, N.; Saravanan, R.S.A.; Kesavan, T.; Maduraiveeran, G.; Sasidharan, M. An efficient palm waste derived hierarchical porous carbon for electrocatalytic hydrogen evolution reaction. Carbon 2019, 152, 188–197. [Google Scholar] [CrossRef]
  37. Cao, X.; Li, Z.; Chen, H.; Zhang, C.; Zhang, Y.; Gu, C.; Xu, X.; Li, Q. Synthesis of biomass porous carbon materials from bean sprouts for hydrogen evolution reaction electrocatalysis and supercapacitor electrode. Int. J. Hydrogen Energy 2021, 46, 18887–18897. [Google Scholar] [CrossRef]
  38. Prabu, N.; Kesavan, T.; Maduraiveeran, G.; Sasidharan, M. Bio-derived nanoporous activated carbon sheets as electrocatalyst for enhanced electrochemical water splitting. Int. J. Hydrogen Energy 2019, 44, 19995–20006. [Google Scholar] [CrossRef]
  39. Arul Saravanan, K.R.; Prabu, N.; Sasidharan, M.; Maduraiveeran, G. Nitrogen-self doped activated carbon nanosheets derived from peanut shells for enhanced hydrogen evolution reaction. Appl. Surf. Sci. 2019, 489, 725–733. [Google Scholar] [CrossRef]
  40. Hoang, V.C.; Gomes, V.G.; Dinh, K.N. Ni- and P-doped carbon from waste biomass: A sustainable multifunctional electrode for oxygen reduction, oxygen evolution and hydrogen evolution reactions. Electrochim. Acta 2019, 314, 49–60. [Google Scholar] [CrossRef]
  41. Wang, Q.; Fei, Z.; Shen, D.; Cheng, C.; Dyson, P.J. Ginkgo Leaf-Derived Carbon Supports for the Immobilization of Iron/Iron Phosphide Nanospheres for Electrocatalytic Hydrogen Evolution. Small 2024, 20, 2309830. [Google Scholar] [CrossRef]
  42. Sundriyal, S.; Shrivastav, V.; Kaur, A.; Mansi; Deep, A.; Dhakate, S.R. Surface and diffusion charge contribution study of neem leaves derived porous carbon electrode for supercapacitor applications using acidic, basic, and neutral electrolytes. J. Energy Storage 2021, 41, 103000. [Google Scholar] [CrossRef]
  43. Zhang, L.; Cheng, X.; Li, L.; Li, X.; Wu, H.; Zheng, J.; Yao, J.; Li, G. Porous and high specific surface area carbon material derived from ginkgo leaves for high-performance symmetric supercapacitors. Biomass Bioenergy 2024, 191, 107481. [Google Scholar] [CrossRef]
  44. Zhu, X.; Yu, S.; Xu, K.; Zhang, Y.; Zhang, L.; Lou, G.; Wu, Y.; Zhu, E.; Chen, H.; Shen, Z.; et al. Sustainable activated carbons from dead ginkgo leaves for supercapacitor electrode active materials. Chem. Eng. Sci. 2018, 181, 36–45. [Google Scholar] [CrossRef]
  45. Sekar, S.; Aqueel Ahmed, A.T.; Pawar, S.M.; Lee, Y.; Im, H.; Kim, D.Y.; Lee, S. Enhanced water splitting performance of biomass activated carbon-anchored WO3 nanoflakes. Appl. Surf. Sci. 2020, 508, 145127. [Google Scholar] [CrossRef]
  46. Sekar, S.; Lee, S.; Vijayarengan, P.; Kalirajan, K.M.; Santhakumar, T.; Sekar, S.; Sadhasivam, S. Upcycling of Wastewater via Effective Photocatalytic Hydrogen Production Using MnO2 Nanoparticles—Decorated Activated Carbon Nanoflakes. Nanomaterials 2020, 10, 1610. [Google Scholar] [CrossRef]
  47. Manasa, P.; Lei, Z.J.; Ran, F. Biomass Waste Derived Low Cost Activated Carbon from Carchorus Olitorius (Jute Fiber) as Sustainable and Novel Electrode Material. J. Energy Storage 2020, 30, 101494. [Google Scholar] [CrossRef]
  48. Sankar, S.; Saravanan, S.; Ahmed, A.T.A.; Inamdar, A.I.; Im, H.; Lee, S.; Kim, D.Y. Spherical activated-carbon nanoparticles derived from biomass green tea wastes for anode material of lithium-ion battery. Mater. Lett. 2019, 240, 189–192. [Google Scholar] [CrossRef]
  49. Xiao, P.-W.; Meng, Q.; Zhao, L.; Li, J.-J.; Wei, Z.; Han, B.-H. Biomass-derived flexible porous carbon materials and their applications in supercapacitor and gas adsorption. Mater. Des. 2017, 129, 164–172. [Google Scholar] [CrossRef]
  50. Qian, W.; Sun, F.; Xu, Y.; Qiu, L.; Liu, C.; Wang, S.; Yan, F. Human hair-derived carbon flakes for electrochemical supercapacitors. Energy Environ. Sci. 2014, 7, 379–386. [Google Scholar] [CrossRef]
  51. Li, Z.; Deng, L.; Kinloch, I.A.; Young, R.J. Raman spectroscopy of carbon materials and their composites: Graphene, nanotubes and fibres. Prog. Mater. Sci. 2023, 135, 101089. [Google Scholar] [CrossRef]
  52. Fu, P.; Zhou, L.; Sun, L.; Huang, B.; Yuan, Y. Nitrogen-doped porous activated carbon derived from cocoon silk as a highly efficient metal-free electrocatalyst for the oxygen reduction reaction. RSC Adv. 2017, 7, 13383–13389. [Google Scholar] [CrossRef]
  53. Watcharamaisakul, S.; Janphuang, N.; Chueangam, W.; Srisom, K.; Rueangwittayanon, A.; Rittihong, U.; Tunmee, S.; Chanlek, N.; Pornsetmetakul, P.; Wirojsirasak, W.; et al. Synthesis of Turbostratic Graphene Derived from Biomass Waste Using Long Pulse Joule Heating Technique. Nanomaterials 2025, 15, 468. [Google Scholar] [CrossRef]
  54. Ferrari, A.C.; Basko, D.M. Raman spectroscopy as a versatile tool for studying the properties of graphene. Nat. Nano. 2013, 8, 235–246. [Google Scholar] [CrossRef] [PubMed]
  55. Homayounfard, A.M.; Maleki, M.; Ghanbari, H.; Kahnamouei, M.H.; Safaei, B. Growth of few-layer flower-like MoS2 on heteroatom-doped activated carbon as a hydrogen evolution reaction electrode. Int. J. Hydrogen Energy 2024, 55, 1360–1370. [Google Scholar] [CrossRef]
  56. Rajasekaran, S.J.; Raghavan, V. Facile synthesis of activated carbon derived from Eucalyptus globulus seed as efficient electrode material for supercapacitors. Diam. Relat. Mater 2020, 109, 108038. [Google Scholar] [CrossRef]
  57. Vijayakumar, M.; Bharathi Sankar, A.; Sri Rohita, D.; Nanaji, K.; Narasinga Rao, T.; Karthik, M. Achieving High Voltage and Excellent Rate Capability Supercapacitor Electrodes Derived From Bio-renewable and Sustainable Resource. ChemistrySelect 2020, 5, 8759–8772. [Google Scholar] [CrossRef]
  58. Lin, P.; Xia, Y.; Liu, Z. Influence of Different Activators on the Structure and Properties of Activated Carbon Based on Bamboo Fiber. Polymers 2022, 14, 5500. [Google Scholar] [CrossRef]
  59. Zhang, L.; Tu, L.-y.; Liang, Y.; Chen, Q.; Li, Z.-s.; Li, C.-h.; Wang, Z.-h.; Li, W. Coconut-based activated carbon fibers for efficient adsorption of various organic dyes. RSC Adv. 2018, 8, 42280–42291. [Google Scholar] [CrossRef]
  60. Shrestha, D.; Maensiri, S.; Wongpratat, U.; Lee, S.W.; Nyachhyon, A.R. Shorea robusta derived activated carbon decorated with manganese dioxide hybrid composite for improved capacitive behaviors. J. Environ. Chem. Eng. 2019, 7, 103227. [Google Scholar] [CrossRef]
  61. Bejjanki, D.; Banothu, P.; Kumar, V.B.; Kumar, P.S. Biomass-Derived N-Doped Activated Carbon from Eucalyptus Leaves as an Efficient Supercapacitor Electrode Material. C 2023, 9, 24. [Google Scholar] [CrossRef]
  62. Sukhbaatar, B.; Qing, W.; Seo, J.; Yoon, S.; Yoo, B. Uniformly dispersed ruthenium nanoparticles on porous carbon from coffee waste outperform platinum for hydrogen evolution reaction in alkaline media. Sci. Rep. 2024, 14, 5850. [Google Scholar] [CrossRef] [PubMed]
  63. Mahmood, N.; Yao, Y.; Zhang, J.-W.; Pan, L.; Zhang, X.; Zou, J.-J. Electrocatalysts for Hydrogen Evolution in Alkaline Electrolytes: Mechanisms, Challenges, and Prospective Solutions. Adv. Sci. 2018, 5, 1700464. [Google Scholar] [CrossRef]
  64. Sukhbaatar, B.; Yoon, S.; Yoo, B. Simple synthesis of a CoO nanoparticle-decorated nitrogen-doped carbon catalyst from spent coffee grounds for alkaline hydrogen evolution. J. Mater. Sci. 2022, 57, 18075–18088. [Google Scholar] [CrossRef]
  65. Shinagawa, T.; Garcia-Esparza, A.T.; Takanabe, K. Insight on Tafel slopes from a microkinetic analysis of aqueous electrocatalysis for energy conversion. Sci. Rep. 2015, 5, 13801, Correction in Sci. Rep. 2020, 10, 1–1. [Google Scholar] [CrossRef]
  66. Yan, X.; Jia, Y.; Odedairo, T.; Zhao, X.; Jin, Z.; Zhu, Z.; Yao, X. Activated carbon becomes active for oxygen reduction and hydrogen evolution reactions. Chem. Commun. 2016, 52, 8156–8159. [Google Scholar] [CrossRef]
  67. Chinnadurai, D.; Karuppiah, P.; Chen, S.-M.; Kim, H.-j.; Prabakar, K. Metal-free multiporous carbon for electrochemical energy storage and electrocatalysis applications. New J. Chem. 2019, 43, 11653–11659. [Google Scholar] [CrossRef]
  68. Sathiskumar, C.; Ramakrishnan, S.; Vinothkannan, M.; Rhan Kim, A.; Karthikeyan, S.; Yoo, D.J. Nitrogen-Doped Porous Carbon Derived from Biomass Used as Trifunctional Electrocatalyst toward Oxygen Reduction, Oxygen Evolution and Hydrogen Evolution Reactions. Nanomaterials 2020, 10, 76. [Google Scholar] [CrossRef]
  69. Xia, C.; Surendran, S.; Ji, S.; Kim, D.; Chae, Y.; Kim, J.; Je, M.; Han, M.-K.; Choe, W.-S.; Choi, C.H.; et al. A sulfur self-doped multifunctional biochar catalyst for overall water splitting and a supercapacitor from Camellia japonica flowers. Carbon Energy 2022, 4, 491–505. [Google Scholar] [CrossRef]
  70. Thomas, A.; Padmanabhan, N.T.; Jelmy, E.J.; Gopinath, P.; John, H. Multifunctional carbonaceous materials derived from tea waste: Towards sustainable solutions for wastewater treatment and hydrogen evolution. Clean Waste Syst. 2025, 10, 100233. [Google Scholar] [CrossRef]
  71. Hoang, V.C.; Dinh, K.N.; Gomes, V.G. Hybrid Ni/NiO composite with N-doped activated carbon from waste cauliflower leaves: A sustainable bifunctional electrocatalyst for efficient water splitting. Carbon 2020, 157, 515–524. [Google Scholar] [CrossRef]
  72. Sekar, S.; Kim, D.Y.; Lee, S. Excellent Oxygen Evolution Reaction of Activated Carbon-Anchored NiO Nanotablets Prepared by Green Routes. Nanomaterials 2020, 10, 1382. [Google Scholar] [CrossRef] [PubMed]
  73. Sankar, S.; Ahmed, A.T.A.; Inamdar, A.I.; Im, H.; Im, Y.B.; Lee, Y.; Kim, D.Y.; Lee, S. Biomass-derived ultrathin mesoporous graphitic carbon nanoflakes as stable electrode material for high-performance supercapacitors. Mater. Des. 2019, 169, 107688. [Google Scholar] [CrossRef]
  74. Wang, J.; Kaskel, S. KOH activation of carbon-based materials for energy storage. J. Mater. Chem. 2012, 22, 23710–23725. [Google Scholar] [CrossRef]
  75. Sekar, S.; Sadhasivam, S.; Shanmugam, A.; Saravanan, S.; Pugazhendi, I.; Lee, Y.; Kim, D.Y.; Manikandan, R.; Chang, S.-C.; Lee, S. Enhanced bifunctional water electrolysis performance of spherical ZnMn2O4 nanoparticles. Int. J. Hydrogen Energy 2025, 141, 721–728. [Google Scholar] [CrossRef]
  76. Ahmed, A.T.A.; Ansari, A.S.; Sree, V.G.; Jana, A.; Meena, A.; Sekar, S.; Cho, S.; Kim, H.; Im, H. Nitrogen-Doped CuO@CuS Core–Shell Structure for Highly Efficient Catalytic OER Application. Nanomaterials 2023, 13, 3160. [Google Scholar] [CrossRef]
  77. Singh, D.K.; Jenjeti, R.N.; Sampath, S.; Eswaramoorthy, M. Two in one: N-doped tubular carbon nanostructure as an efficient metal-free dual electrocatalyst for hydrogen evolution and oxygen reduction reactions. J. Mater. Chem. A 2017, 5, 6025–6031. [Google Scholar] [CrossRef]
  78. Kumar, M.; Nagaiah, T.C. A NiCu–MoS2 electrocatalyst for pH-universal hydrogen evolution reaction and Zn–air batteries driven self-power water splitting. J. Mater. Chem. A 2023, 11, 18336–18348. [Google Scholar] [CrossRef]
  79. Chen, H.; Luo, X.; Huang, S.; Yu, F.; Li, D.; Chen, Y. Phosphorus-doped activated carbon as a platinum-based catalyst support for electrocatalytic hydrogen evolution reaction. J. Electroanal. Chem. 2023, 948, 117820. [Google Scholar] [CrossRef]
Figure 1. Low- and high-magnification FE-SEM images of (a,b) N-AC and (c,d) G-AC nanostructures.
Figure 1. Low- and high-magnification FE-SEM images of (a,b) N-AC and (c,d) G-AC nanostructures.
Ijms 26 08502 g001
Figure 2. (a) XRD patterns, (b) Raman spectra, (c) N2-absorption and desorption isotherms, and (d) pore size distributions of the N-AC and G-AC nanostructures.
Figure 2. (a) XRD patterns, (b) Raman spectra, (c) N2-absorption and desorption isotherms, and (d) pore size distributions of the N-AC and G-AC nanostructures.
Ijms 26 08502 g002
Figure 3. (a) Full survey XPS spectra of N-AC and G-AC nanostructures. (b) C 1s and (c) O 1s core-level spectra of N-AC nanoflakes. (d) C 1s and (e) O 1s core-level spectra of G-AC nanosponges.
Figure 3. (a) Full survey XPS spectra of N-AC and G-AC nanostructures. (b) C 1s and (c) O 1s core-level spectra of N-AC nanoflakes. (d) C 1s and (e) O 1s core-level spectra of G-AC nanosponges.
Ijms 26 08502 g003
Figure 4. CV curves of (a) N-AC and (b) G-AC catalysts. Non-Faradaic CV curves of the (c) N-AC and (d) G-AC catalysts. Non-Faradaic JDL at 0.07 V as a function of scan rate for (e) N-AC and (f) G-AC catalysts.
Figure 4. CV curves of (a) N-AC and (b) G-AC catalysts. Non-Faradaic CV curves of the (c) N-AC and (d) G-AC catalysts. Non-Faradaic JDL at 0.07 V as a function of scan rate for (e) N-AC and (f) G-AC catalysts.
Ijms 26 08502 g004
Figure 5. Electrocatalytic HER performance of N-AC and G-AC catalysts: (a) iR-corrected LSV curves, (b) Tafel plots, (c) chronopotentiometric profiles at various current densities (−10 to −100 mA/cm2), and (d) long-term durability analysis. LSV curves of the (e) N-AC and (f) G-AC catalysts before and after the HER durability test. FE-SEM images of the N-AC (inset of (e)) and the G-AC (inset of (f)) catalysts after the stability test.
Figure 5. Electrocatalytic HER performance of N-AC and G-AC catalysts: (a) iR-corrected LSV curves, (b) Tafel plots, (c) chronopotentiometric profiles at various current densities (−10 to −100 mA/cm2), and (d) long-term durability analysis. LSV curves of the (e) N-AC and (f) G-AC catalysts before and after the HER durability test. FE-SEM images of the N-AC (inset of (e)) and the G-AC (inset of (f)) catalysts after the stability test.
Ijms 26 08502 g005
Figure 6. Nyquist plots of (a) N-AC and (b) G-AC catalysts before and after the stability test (inset: equivalent circuit model).
Figure 6. Nyquist plots of (a) N-AC and (b) G-AC catalysts before and after the stability test (inset: equivalent circuit model).
Ijms 26 08502 g006
Figure 7. Schematic representation of the synthesis process of (a) N-AC and (b) G-AC nanostructures from neem and ginkgo leaves, respectively, via KOH activation.
Figure 7. Schematic representation of the synthesis process of (a) N-AC and (b) G-AC nanostructures from neem and ginkgo leaves, respectively, via KOH activation.
Ijms 26 08502 g007
Table 1. Comparison of HER performance between biomass-derived N-AC and G-AC nanostructures and previously reported carbon-based electrocatalysts.
Table 1. Comparison of HER performance between biomass-derived N-AC and G-AC nanostructures and previously reported carbon-based electrocatalysts.
Biomass ResourcesCatalystActivation Sourceη (mV)ST (mV/dec)ElectrolyteRef.
Ginkgo LeavesG-ACKOH26240.5 M H2SO4This Work
Neem LeavesN-ACKOH40460.5 M H2SO4This Work
Bean SproutsBS-800HF Etching413980.5 M H2SO4[37]
CarrotsPCThermal Annealing9392730.1 M KOH[40]
Ooty VarkeyNACSKOH380850.5 M H2SO4[38]
Cattail FiberNPCFKOH2441350.5 M H2SO4[35]
Eucalyptus LeavesELC-700KOH39360.5 M H2SO4[13]
Palm WasteHPNSKOH330630.5 M H2SO4[36]
Human HairHH-AC-700KOH16510.5 M H2SO4[16]
Commercial ACD-ACNH3334660.5 M H2SO4[66]
Broccoli StemsNA9KOH1841641 M H2SO4[67]
Peanut ShellsPSACKOH80750.5 M H2SO4[39]
Golden Shower PodsN-PCUrea179981 M KOH[68]
Rice HuskRH-CG-600KOH33670.5 M H2SO4[19]
Camellia Japonica FlowerSA-CameKOH154891 M KOH[69]
Tea WasteG-ACOHKOH3491280.5 M H2SO4[70]
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

Sekar, S.; Sadhasivam, S.; Shanmugam, A.; Sekar, S.; Lee, Y.; Lee, S. Boosted Hydrogen Evolution Catalysis Using Biomass-Derived Mesoporous Carbon Nanosponges. Int. J. Mol. Sci. 2025, 26, 8502. https://doi.org/10.3390/ijms26178502

AMA Style

Sekar S, Sadhasivam S, Shanmugam A, Sekar S, Lee Y, Lee S. Boosted Hydrogen Evolution Catalysis Using Biomass-Derived Mesoporous Carbon Nanosponges. International Journal of Molecular Sciences. 2025; 26(17):8502. https://doi.org/10.3390/ijms26178502

Chicago/Turabian Style

Sekar, Sankar, Sutha Sadhasivam, Atsaya Shanmugam, Saravanan Sekar, Youngmin Lee, and Sejoon Lee. 2025. "Boosted Hydrogen Evolution Catalysis Using Biomass-Derived Mesoporous Carbon Nanosponges" International Journal of Molecular Sciences 26, no. 17: 8502. https://doi.org/10.3390/ijms26178502

APA Style

Sekar, S., Sadhasivam, S., Shanmugam, A., Sekar, S., Lee, Y., & Lee, S. (2025). Boosted Hydrogen Evolution Catalysis Using Biomass-Derived Mesoporous Carbon Nanosponges. International Journal of Molecular Sciences, 26(17), 8502. https://doi.org/10.3390/ijms26178502

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