Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

2D catalysts for assisted water electrolysis: mechanistic insights and theoretical perspectives for industrial hydrogen generation

Hyojung Lim a, Seonghyeon Park a, Jinuk Choi a, Junho Shima, Subramani Surendrana and Uk Sim *ab
aHydrogen Energy Technology Laboratory, Korea Institute of Energy Technology (KENTECH), Naju, Jeonnam 58330, Republic of Korea. E-mail: usim@kentech.ac.kr
bResearch Institute, NEEL Sciences, Inc., Naju, Jeonnam 58326, Republic of Korea

Received 24th June 2025 , Accepted 24th September 2025

First published on 16th October 2025


Abstract

The efficiency of conventional water electrolysis is fundamentally constrained by the sluggish kinetics and high overpotential of the oxygen evolution reaction (OER). Assisted water electrolysis has emerged as a promising strategy to overcome this limitation by replacing OER with the selective oxidation of small organic or nitrogen-containing molecules such as urea, ammonia, methanol, ethanol, glycerol, and formic acid. These alternative anodic reactions offer significantly lower thermodynamic oxidation potentials, thereby enabling hydrogen production at reduced cell voltages while simultaneously achieving pollutant remediation and value-added chemical synthesis. Two-dimensional (2D) materials have garnered increasing attention as efficient catalysts for oxidation reactions in assisted water electrolysis, owing to their unique structural and electronic properties. This review summarizes recent progress in 2D catalysts, including layered double hydroxides, transition metal dichalcogenides, MXenes, metallenes, and graphene-based materials, emphasizing their roles in facilitating various oxidation reactions. Key strategies, including doping, defect engineering, and interface modulation, are discussed in relation to enhancing catalytic activity, selectivity, and durability. Thermodynamic analyses and Pourbaix diagrams are introduced to provide insight into the reaction pathways and stability windows of both feedstocks and catalysts under various electrochemical conditions. By integrating rational catalyst design with a comprehensive understanding of various oxidation reactions, assisted water electrolysis using 2D catalysts offers a compelling pathway toward sustainable hydrogen production. The co-benefits of improved energy efficiency and environmental sustainability position this approach as a promising solution to current energy and environmental challenges. Developing 2D materials and understanding reactions are expected to accelerate the implementation of next-generation electrolysis systems aligned with global carbon neutrality goals.

Keywords: 2D catalysts; Assisted water electrolysis; Hydrogen generation; Pourbaix diagrams; Density functional theory (DFT).


image file: d5im00116a-p1.tif

Hyojung Lim

Hyojung Lim is a doctoral student in the Hydrogen Energy Technology Laboratory at the Korea Institute of Energy Technology (KENTECH), Republic of Korea. She earned her B.S. in Chemistry and Energy Engineering from Sookmyung Women’s University in 2022 and her M.S. from KENTECH in 2025. Her research focuses on electrochemical ammonia synthesis and water splitting for sustainable energy applications.

image file: d5im00116a-p2.tif

Seonghyeon Park

Seonghyeon Park is a master’s student in the Hydrogen Energy Technology Laboratory at the Korea Institute of Energy Technology (KENTECH), Republic of Korea. He received his B.S. in Chemistry from Pukyong National University. His research focuses on the design and development of electrocatalysts for green ammonia production.

image file: d5im00116a-p3.tif

Jinuk Choi

Jinuk Choi is a doctoral student in the Hydrogen Energy Technology Laboratory at the Korea Institute of Energy Technology (KENTECH), Republic of Korea. He earned his B.S. and M.S. degrees in Materials Science and Engineering from Chosun University (2019) and Hanyang University (2021), respectively. His research focuses on the electrochemical synthesis of ammonia and urea under ambient conditions.

image file: d5im00116a-p4.tif

Uk Sim

Prof. Uk Sim is an Associate Professor in the Hydrogen Energy Technology Laboratory at the Korea Institute of Energy Technology (KENTECH), Republic of Korea. He earned his B.S. (2007), M.S. (2009), and Ph.D. (2016) in Materials Science and Engineering from Seoul National University. Following his doctoral studies, he completed a postdoctoral fellowship at Stanford University in 2017. He later served as an Associate Professor in the Department of Materials Science and Engineering at Chonnam National University. In addition to his academic roles, he is the Founder and CEO of Neel Sciences Inc., a research institute. His research focuses on the design and development of nanomaterials for energy production, conversion, and storage, with particular emphasis on renewable energy systems based on photo- and electrochemical reactions aimed at building a sustainable energy future.


1 Introduction

1.1 Importance of assisted water electrolysis

With the urgency of phasing out fossil fuels, electrochemical water electrolysis has gained prominence as a sustainable pathway for hydrogen production. Unlike thermochemical methods that rely on hydrocarbons, water electrolysis offers a carbon-free approach that can be directly powered by renewable electricity.1–3 This eliminates greenhouse gas emissions at the source and enables decentralized hydrogen generation. However, the overall water electrolysis process is energetically demanding due to the sluggish oxygen evolution reaction (OER), which not only requires a high overpotential but also suffers from poor selectivity and stability in practical conditions.4–7 To address these limitations, assisted water electrolysis strategies have garnered significant attention.8–12 By replacing the OER with alternative oxidation reactions involving small organic or nitrogen-containing molecules, the overall energy input can be substantially reduced. These assisted oxidation reactions typically exhibit lower theoretical oxidation potentials and faster kinetics compared to OER, enabling more energy-efficient hydrogen production and coupled chemical synthesis.

Moreover, assisted water oxidation not only enables the conversion of pollutants such as ammonia (NH3), hydrazine (N2H4), and urea (CO(NH2)2) into non-toxic products, but also facilitates the synthesis of value-added products like formic acid (HCOOH) from methanol (CH3OH) and glycerol (C3H8O3).13–17 Therefore, selecting appropriate oxidation substrates and understanding their electrochemical behavior are crucial for designing effective assisted water electrolysis systems.

1.2 Types of assisted water electrolysis

Various small molecules, including ammonia, hydrazine, urea, methanol, ethanol, and formic acid, have been explored as anodic substrates to replace the oxygen evolution reaction (OER) in water electrolysis (Fig. 1). Each of these species exhibits distinct electrochemical behavior depending on the pH, applied potentials, and thermodynamic properties. The ammonia oxidation reaction (AOR) produces N2, which can exist in either the gas or aqueous phase.18 The standard electrode potential (E0) of the reaction varies slightly depending on the phase of the N2 product. In addition, the speciation of the ammonia reactant depends on the solution pH, governed by the acid–base equilibrium of NH4+ ⇄ NH3 + H+, with a pKα value of 9.251 under standard conditions (25 °C, 1 atm). Consequently, the AOR pathway differs depending on the phase states of both NH3 and N2, as illustrated in the following half-reactions (eqn (1)–(4)):19,20
 
2NH4+ ⇌ N2(g) + 8H+ + 6e,  E0 = 0.275 V vs. SHE (1)
 
2NH4+ ⇌ N2(aq) + 8H+ + 6e,  E0 = 0.306 V vs. SHE (2)
 
2NH3(aq) + 6OH ⇌ N2(g) + 6H2O(l) + 6e, E0 = 0.092 V vs. SHE (3)
 
2NH3(aq) + 6OH ⇌ N2(aq) + 6H2O(l) + 6e, E0 = 0.123 V vs. SHE (4)

image file: d5im00116a-f1.tif
Fig. 1 Schematic illustration of (a) conventional water electrolysis (OER-HER) and (b) assisted water electrolysis, where various anodic reactions (AOR, UOR, MOR, EOR, FAOR, HzOR) replace the OER.

Urea is a major pollutant commonly found in wastewater and has been utilized as a feedstock for assisted water electrolysis.21 The theoretical potentials of the urea oxidation reaction (UOR) have been extensively investigated due to the relatively complex nature of the reaction, which involves multi-electron transfer and consideration of the product phases. The UOR proceeds according to the following half-reaction (eqn (9)):22,23

 
CO(NH2)2(aq) + 6OH ⇌ N2(g) + CO2(g) + 5H2O(l) + 6e, E0 = 0.072 V vs. SHE (5)

Hydrazine, commonly found in industrial wastewater, has also been widely investigated as an anodic feedstock for assisted water electrolysis.24 Similar to the AOR, the standard electrode potentials of the hydrazine oxidation reaction (HzOR) depend on the pH of the electrolyte and the physical phase of the N2 product. The acid–base equilibrium of hydrazine is governed by the reaction N2H5+ ⇄ N2H4 + H+, with a pKα value of 8.10 under standard conditions.25 Accordingly, the thermodynamic driving force for HzOR varies based on the protonation state of hydrazine and the phase of the N2 produced, as illustrated in the following half-reactions (eqn (5)–(8)):26

 
N2H5+ ⇌ N2(aq) + 5H+ + 4e,  E0 = −0.167 V vs. SHE (6)
 
N2H5+ ⇌ N2(g) + 5H+ + 4e,  E0 = −0.214 V vs. SHE (7)
 
N2H4(aq) + 4OH ⇌ N2(aq) + 4H2O(l) + 4e, E0 = −0.284 V vs. SHE (8)
 
N2H4(aq) + 4OH ⇌ N2(g) + 4H2O(l) + 4e, E0 = −0.332 V vs. SHE (9)

The theoretical potential of the methanol oxidation reaction (MOR) is influenced by both the pH of the electrolyte and the nature of carbon-containing intermediates or products.27 Typically, a standard potential of 0.016 V vs. SHE is used for methanol oxidation in direct methanol fuel cells.28 However, under aqueous conditions relevant to assisted water electrolysis, the thermodynamic properties of methanol differ. The MOR proceeds through multi-electron transfer steps, ultimately leading to the formation of CO2. The corresponding half-reactions are summarized as follows (eqn (10)–(12)):29,30

 
CH3OH(l) + H2O(l) ⇌ CO2(g) + 6H+ + 6e,E0 = 0.016 V vs. SHE (10)
 
CH3OH(aq) + H2O(l) ⇌ CO2(g) + 6H+ + 6e, E0 = 0.032 V vs. SHE (11)
 
CH3OH(aq) + H2O(l) ⇌ CO2(aq) + 6H+ + 6e, E0 = 0.046 V vs. SHE (12)

Furthermore, MOR can potentially yield value-added chemicals such as formic acid. The theoretical potential for the partial oxidation pathway (CH3OH → HCOOH, E0 = 0.103 V vs. SHE) is thermodynamically slightly less favorable than the complete oxidation to CO2 (E0 = 0.046 V vs. SHE). Although CO2 remains the most stable end-product, the selective formation of HCOOH is kinetically accessible under suitable catalysts and reaction conditions, thereby offering an opportunity for co-generation of chemicals alongside hydrogen. The CH3OH-to-HCOOH pathway is summarized as follows (eqn (13)):

 
CH3OH(aq) + H2O(l) ⇌ HCOOH(aq) + 4H+ + 4e, E0 = 0.103 V vs. SHE (13)

Similarly, the ethanol oxidation reaction (EOR) exhibits multiple pathways depending on the reaction conditions and the nature of the final products.31 A standard potential of 0.084 V vs. SHE for ethanol oxidation is widely referenced in the field of direct ethanol fuel cells.32 However, under aqueous conditions relevant to assisted water electrolysis, the thermodynamic behavior of ethanol differs. Therefore, the aqueous phase of both reactants and products should be considered for accurate potential calculation. The consideration of phase transitions between reactants and products results in variations in the calculated theoretical potentials. Representative half-reactions are summarized below (eqn (14)–(16)):30,33

 
CH3CH2OH(l) + 12OH ⇌ 2CO2(g) + 9H2O(l) + 12e, E0 = 0.084 V vs. SHE (14)
 
CH3CH2OH(aq) + 12OH ⇌ 2CO2(g) + 9H2O(l) + 12e, E0 = 0.090 V vs. SHE (15)
 
CH3CH2OH(aq) + 12OH ⇌ 2CO2(aq) + 9H2O(l) + 12e, E0 = 0.105 V vs. SHE (16)

Glycerol, a major organic pollutant frequently generated from biodiesel and other industrial processes, has also been widely utilized as a feedstock for assisted water electrolysis.34 Its oxidation offers the dual benefits of energy-efficient hydrogen production and simultaneous waste valorization. The following half-reaction describes the theoretical redox potential for the glycerol oxidation reaction (GOR) (eqn (17)):35

 
C3H8O3(l) + 3H2O(l) ⇌ 3CO2(g) + 14H+ + 14e, E0 = 0.003 V vs. SHE (17)

The formic acid oxidation reaction (FAOR) has also attracted considerable attention as an alternative anodic reaction in assisted water electrolysis due to its benefits in environmental remediation and energy efficiency.36 As a common pollutant in various industrial waste streams, formic acid can be selectively oxidized to carbon dioxide under mild electrochemical conditions. Importantly, FAOR also exhibits a significantly lower theoretical oxidation potential compared to the OER, making it highly advantageous for reducing the overall energy input in electrolysis systems. The FAOR proceeds according to the following half-reactions (eqn (18)):37

 
HCOOH(aq) ⇌ CO2(g) + 2H+ + 2e, E0 = −0.114 V vs. SHE (18)

The oxidation potentials were calculated using the FactSage software, excluding UOR and GOR due to the lack of thermodynamic data. These calculated potentials provide valuable guidance for predicting precise energy consumption and optimizing the system design of assisted water electrolysis.

The Pourbaix diagram provides a thermodynamic framework that maps the stable phases of reactants and products as a function of potential and pH. Such visualizations offer valuable insight into the electrochemical operation window and the prediction of possible products under given conditions. Fig. 2 presents the Pourbaix diagrams, calculated using the FactSage thermochemical software package with the FactPS database (excluding urea), for assisted water electrolysis with various molecular feedstocks.


image file: d5im00116a-f2.tif
Fig. 2 Pourbaix diagrams for (a) ammonia, (b) hydrazine, (c) urea, (d) methanol, (e) ethanol, and (f) formic acid redox at 298.15 K and 1 bar.

To calculate the theoretical potentials and construct the Pourbaix diagrams, the thermodynamic reversible voltage was determined using the equation:

 
E0 = −ΔG0/nF (19)
under standard conditions. The Pourbaix diagrams were then derived using the generalized Nernst equation:
 
αA + βB + hH+ + ze = γC + δD (20)
 
image file: d5im00116a-t1.tif(21)

Notably, the theoretical oxidation potentials of these molecules are significantly lower than that of the conventional OER (1.23 V vs. RHE), thereby enabling more energy-efficient anodic processes.38 These results of the calculations represent standard-state thermodynamics. Deviations are expected under real electrolytic environments due to concentration, ionic strength, and mass-transport effects. Nevertheless, they provide a consistent baseline for benchmarking assisted water electrolysis.

Fig. 2a illustrates the Pourbaix diagram for the AOR, where the theoretical oxidation potential remains well below that of OER across the entire pH range. Notably, a change in the slope of the potential-pH relationship is observed at pH 9.251, corresponding to the pKα of the NH4+/NH3 equilibrium. Similarly, the HzOR exhibits even lower theoretical potentials than AOR throughout the pH spectrum (Fig. 2b). As shown in the Pourbaix diagram, the slope of the potential curve shifts at pH 8.10, which corresponds to the pKα of the N2H5+/N2H4 equilibrium. These features highlight the favorable thermodynamics of HzOR and AOR for energy-saving hydrogen production through assisted water electrolysis. Similarly, UOR exhibits a low theoretical potential while generating benign products (CO2 and N2), making it a highly attractive anodic reaction.39 As shown in Fig. 2c, these products remain thermodynamically stable across a wide pH range. In contrast, the potential remains significantly lower than that of OER, highlighting the advantages of UOR in reducing energy consumption and treating urea-rich wastewater.

To further understand the behavior of MOR across the entire pH range, a Pourbaix diagram for CH3OH was constructed (Fig. 2d). The oxidation potential of MOR remains consistently below that of the conventional OER, confirming its thermodynamic advantage. Notably, the slope of the potential-pH relationship changes depending on the dominant carbonate species formed (CO2, HCO3, or CO32−), which vary depending on pH. These shifts in product speciation influence the MOR potential, with strongly alkaline conditions exhibiting a larger potential gap between MOR and OER than acidic or neutral environments. The red dashed line indicates the kinetically accessible partial oxidation pathway (CH3OH → HCOOH), whereas the solid lines represent the thermodynamically favorable complete oxidation to CO2. Similarly, Fig. 2e presents the Pourbaix diagram for EOR. The theoretical oxidation potential of EOR is slightly higher than that of MOR, primarily due to the presence of a C–C bond in ethanol, which requires greater energy to cleave during the oxidation process. In the case of FAOR, Fig. 2f displays an additional inflection in the potential curve resulting from the acid–base equilibrium between HCOOH and HCOO (pKα = 3.745). This protonation-dependent speciation results in a distinct transition in the Pourbaix diagram, further influencing the electrochemical window for FAOR operation. The thermodynamic evaluation and corresponding Pourbaix diagrams of various molecular feedstocks, including ammonia, hydrazine, urea, methanol, ethanol, glycerol, and formic acid, demonstrate their promise as viable anodic alternatives to the conventional OER in water electrolysis. Each molecule exhibits distinct pH-dependent electrochemical behavior, governed by the acid–base equilibrium and phase stability of the involved species. Notably, all of the reactions investigated exhibit significantly lower theoretical oxidation potentials compared to OER, thereby offering considerable reductions in energy input. These findings underscore the significance of selecting rational feedstocks and optimizing electrolytes in the development of high-efficiency assisted water electrolysis systems.

1.3 Types of 2D catalysts and their properties

Two-dimensional (2D) materials, defined by their atomic- or few-layer thickness, exhibit distinct structural and electronic properties that make them highly attractive for electrocatalysis.19 Compared to 1D catalysts such as nanowires and nanotubes, which offer directional electron transport, and 3D structures like metal–organic frameworks (MOFs) or porous bulk materials that provide high loading capacity, 2D materials exhibit a unique balance of surface accessibility and electronic tunability. This dimensional contrast helps highlight the distinct advantages of 2D catalysts in electrochemical applications. 2D materials offer high surface area, abundant exposed active sites, and excellent charge transport properties.40,41 In this review, we summarize recent advances in various 2D materials employed in assisted water electrolysis, with a particular focus on their roles and performance. Further, the electrochemical test protocols for each catalyst for their respective assisted water electrolysis are given in Table 1.
Table 1 Electrochemical test protocols of each catalyst for their respective assisted water electrolysis
Reaction Catalyst Electrolyte Cell system Counter electrode Reference electrode Working electrode fabrication Ref.
UOR MoSe2/NiSe2 1 M KOH + 0.5 M urea 3-Electrode Pt electrode Ag/AgCl Catalysts were grown on NF 77
NiOOH/(LDH/α-FeOOH) 1 M KOH + 0.33 M urea 3-Electrode Pt plate Hg/HgO Catalysts were grown on NF 79
Mo-FeNi LDH 1 M KOH + 0.33 M urea 3-Electrode Carbon rod Hg/HgO Ink dropping on the glassy carbon electrode 80
HzOR NbS2 1 M KOH + 0.5 M hydrazine 3-Electrode Graphite rod Ag/AgCl Ink dropping on the glassy carbon electrode 85
Co(OH)2/MoS2/CC 1 M KOH + 0.4 M hydrazine 3-Electrode Pt sheet SCE Catalysts were grown on CC 86
Ruc/NiFe-LDH 1 M KOH + 0.3 M hydrazine 3-Electrode Graphite rod Hg/HgO Catalysts were grown on NF 87
MOR NiMn-LDHs 1 M KOH + 3 M CH3OH 3-Electrode Graphite rod Ag/AgCl Catalysts were grown on NF 93
NiFe-LDH/NiFe-HAB/CF 1 M KOH + 3 M CH3OH 3-Electrode Pt foil SCE Catalysts were grown on CF 95
EOR CoFe LDH/MoS2/Ni3S2/NF 1 M KOH + 0.34 M C2H5OH 3-Electrode Graphite rod SCE Catalysts were grown on NF 94
GOR Co2[NiPcS8] 1 M KOH + 0.1 M glycerol 3-Electrode Pt mesh Ag/AgCl Ink dropping on the carbon paper 98
PtSA–NiCo LDH/NF 1 M KOH + 0.1 M glycerol 3-Electrode Carbon rod Hg/HgO Catalysts were grown on NF 99
FAOR B–PdCu-c/a 0.5 M H2SO4 + 0.5 M formic acid 3-Electrode Pt plate Ag/AgCl Ink dropping on the glassy carbon electrode 100


Layered double hydroxides (LDHs) and transition metal dichalcogenides (TMDs) are among the most representative and widely investigated 2D materials.42–48 LDHs, composed of positively charged metal hydroxide layers with intercalated anions, offer abundant active sites and enable flexible compositional engineering, making them excellent candidates not only for oxygen evolution but also for alcohol oxidation reactions. NiFe-LDH and CoFe-LDH are representative examples of LDHs.49–52 However, LDHs suffer from intrinsically low conductivity, which limits their catalytic activity. To address this, the electronic structure reconstruction of LDHs should be conducted by introducing heterostructures, forming cationic and anionic vacancies, and so on.

TMDs are layered compounds of the general formula MX2, where M is a transition metal and X is a chalcogen (S, Se, or Te), exhibiting unique electronic, optical, and catalytic properties.53–55 However, their catalytic activity primarily occurs at the edge sites with unsaturated dangling bonds.56 Thanks to their tunable electronic structure, TMDs allow for the modulation of catalytic behavior through layer control, doping, and defect or interface engineering, making them highly adaptable for various electrochemical reactions.57,58 Among TMD catalysts, MoS2, WS2, and MoSe2 have been most widely investigated for their electrochemical performance.

In addition, 2D metal–organic frameworks (MOFs) and graphitic carbon-based materials such as graphene and MXenes are also increasingly explored for their tailored structures and electronic properties.59–63 Graphene and MXenes have been frequently combined with other materials and widely studied recently to synergistically enhance electrocatalytic performance through improved conductivity and increased exposure of active sites. For instance, MoS2/graphene CoNi-MOF/MXene composites, Co2FeO4@rGO composites, Mo2Ti2C3Tx, and Mxene-based metal oxide electrocatalysts.64–69 Despite their advantages, carbon-based materials are prone to oxidation under harsh electrochemical conditions, which not only decreases their stability but also impairs catalytic performance.70 Therefore, considerable efforts such as defect control, surface passivation, introduction of oxidation-resistant termination groups, and other strategies are required to enhance their overall performance.71–73

2 2D catalysts for urea oxidation

The urea oxidation reaction (UOR) is one of the most attractive alternative anode reactions for hydrogen generation. The UOR only requires 0.37 V vs. RHE, which is significantly less than the 1.23 V vs. RHE required for OER. Furthermore, a considerable amount of the wastewater contains urea, which provokes environmental pollution.74,75 Therefore, if hydrogen is generated using urea derived from sewage, it is much more environmentally friendly. However, along with these advantages, UOR still faces a significant challenge due to the sluggish reaction caused by the 6-electron mechanism.76 Thus, the catalysts with high performance are essential for the efficient generation of hydrogen.

Chen et al. reported a MoSe2/NiSe2 electrocatalyst for urea-assisted water electrolysis.77 The catalyst exhibited improved conductivity and UOR activity due to the introduction of the heterostructure. As shown in Fig. 3a, the X-ray diffraction (XRD) analysis was conducted on MoSe2/NiSe2. The peaks at about 2θ = 38.0°, 41.9°, 47.5°, 56.4°, 76.4° correspond to (103), (006), (105), (110), and (205) crystal planes of MoSe2, respectively. Similarly, the peaks at approximately 2θ = 29.6°, 42.4°, 50.2°, 52.6°, and 55.0° correspond to the (200), (220), (311), (222), and (023) crystal planes of NiSe2, respectively. However, the peaks of MoSe2 were not clearly observed due to its poor crystalline structure. To ascertain the presence of MoSe2/NiSe2 heterostructure, Raman spectroscopy was performed on MoSe2/NiSe2 (Fig. 3b). In the Raman spectroscopy, the Ag mode of NiSe2 peak is located at 219.8 cm−1, and the A1g, E12g modes of MoSe2 peaks are located at 235.0 cm−1 and 286.6 cm−1, respectively. Through the XRD and X-ray photoelectron spectroscopy (XPS) results, the presence of the MoSe2/NiSe2 heterostructure was confirmed. Fig. 3c shows the high-resolution X-ray photoelectron spectroscopy (XPS) spectra of Mo 3d, providing further insight into the chemical states of the composites. In the graph, the Mo 3d5/2 and 3d3/2 of Mo4+ were confirmed at approximately 228.26 and 231.94 eV, respectively. These characteristic Mo4+ peaks confirm the presence of the MoSe2 structure in both samples. In addition, the Mo 3d peak of MoSe2/NiSe2 exhibits a negative shift compared with bare MoSe2, which is attributed to electron transfer from Ni to Mo due to the electronegativity difference between Mo (2.16) and Ni (1.91). Through the XRD, Raman spectroscopy, and XPS results, the presence of the MoSe2/NiSe2 heterostructure was confirmed. Fig. 3d displays the linear sweep voltammetry (LSV) curve for UOR in 1 M KOH with 0.5 M urea solution. To compare the catalytic activity, the bare MoSe2, NiSe2, and NiMoO4 were tested together. The MoSe2/NiSe2 exhibited the highest UOR activity, requiring a potential of only 1.41 V vs. RHE to reach a current density of 50 mA cm−2, whereas MoSe2, NiSe2, and NiMoO4 required 1.49, 1.45, and 1.45 V vs. RHE, respectively. This result represents the potential for UOR using a TMD-based catalyst.


image file: d5im00116a-f3.tif
Fig. 3 Various 2D catalysts for the urea oxidation reaction. (a) XRD patterns of MoSe2/NiSe2, (b) Raman spectrum of MoSe2/NiSe2, MoSe2, and NiSe2, (c) Mo 3d high-resolution XPS spectra, and (d) LSV curve of MoSe2/NiSe2, MoSe2, NiSe2, and NiMoO4/NF for UOR, reprinted with permission from ref. 77, copyright 2024 American Chemical Society; (e) structural model of α-FeOOH/NiOOH, (f) calculated EH value of NiOOH and α-FeOOH/NiOOH at oxygen sites, (g) required potentials at 10, 50, and 100 mA cm−2, and (h) Cdl value of catalysts, reprinted with permission from ref. 79, copyright 2022 Wiley-VCH GmbH; (i) LSV curves of Mo-FeNi LDH for UOR and OER, (j) LSV curves comparing UOR‖HER and OER‖HER performance, (k) UOR mechanisms on Mo–FeNi LDH, and (l) the free-energy pathway for the UOR, reprinted with permission from ref. 80, copyright 2023 Wiley-VCH GmbH.

Extensive research is being conducted on NiFe-based layered double hydroxides (LDHs), which exhibit exceptional performance in oxidation reactions due to their 2D structure. This is attributed to Ni oxyhydroxide (NiOOH), which is generated through a self-reconstruction process.78 However, the NiOOH can easily undergo hydrogenation due to the instability of the Ni3+. To address this challenge, Cai et al. synthesized NiOOH/(LDH/α-FeOOH), which contains stabilized NiOOH from introducing α-FeOOH into NiFe-LDH.79 Density functional theory (DFT) calculations were performed to compare the hydrogenation formation energy (EH) of the oxygen site in NiOOH and α-FeOOH/NiOOH using a composite model of α-FeOOH/NiOOH (Fig. 3e). The rightmost columns in Fig. 3f indicate the average EH value of oxygen sites. In the case of NiOOH, the average EH value was −3.18 eV, whereas it was −2.92 eV for α-FeOOH/NiOOH. A lower EH indicates a higher tendency for the reduction of Ni3+ to Ni2+. The electrochemical measurements demonstrated the highest performance for UOR in a 1 M KOH solution with 0.33 M urea (Fig. 3g). The NiOOH/(LDH/α-FeOOH) required 1.35, 1.37, and 1.40 V vs. RHE to reach current densities of 10, 50, and 100 mA cm−2, respectively. Fig. 3h shows the measured double-layer capacitance (Cdl) value for the catalysts. The Cdl values were 1.6, 2.5, 4.4, and 10.0 mF c−2 for LDH, LDH/α-FeOOH, NiOOH/LDH, and NiOOH/(LDH/α-FeOOH), respectively, indicating that NiOOH/(LDH/α-FeOOH) has a large electrochemically active surface area (ECSA).

Huo et al. synthesized high-valence metal (V, Mn, Mo) doped FeNi LDH with hollow morphology using spindle-like Fe-MIL-88A.80 The large atomic radii of high-valence metals result in lattice expansion, which alters the electronic structure. Furthermore, the hollow structure enhances mass and heat transport ability. Among the high-valence metals, the Mo–FeNi LDH exhibited the highest activity for UOR in a 1 M KOH solution with 0.33 M urea. The Mo-FeNi LDH catalyst required only 1.32 V vs. RHE, which is 172 mV lower than that of the OER (Fig. 3i). Additionally, Fig. 3j shows the difference between HER‖OER, and HER‖UOR using a Pt/C‖Mo–FeNi LDH two-electrode electrolyzer. While conventional water electrolysis requires 1.49 V vs. RHE at 10 mA cm−2, the urea-assisted water electrolysis requires 1.38 V vs. RHE, indicating that hydrogen generation occurs at a much lower potential. The DFT calculations demonstrate that Mo doping stabilizes CO* and NH* intermediates during the UOR, leading to a decrease in the energy barrier of the rate-determining step (RDS). More specifically, while the dissociation of adsorbed urea to CO* and NH* requires only 0.12 eV for Mo–FeNi LDH, the FeNi-LDH requires 3.48 eV. This result indicates that the Mo–FeNi LDH exhibits a significantly higher ability to stabilize intermediates (CO*, NH*) than FeNi-LDH. This is because the electron transfer from Ni to Mo, derived from the electronegativity difference, regulates the electronic structure (Fig. 3k and l).

UOR not only decreases the potential required for hydrogen generation but also offers environmental advantages. Thus, 2D structure-based catalysts are being actively researched due to their outstanding catalytic properties. For instance, the MoSe2/NiSe2 exhibits enhanced UOR activity by introducing a heterostructure on TMDs. Additionally, high-valence metal-doped LDH or α-FeOOH, introduced LDH, has shown high performance for UOR. These studies demonstrate the improved performance of 2D structures and suggest infinite possibilities for their application.

3 2D catalysts for hydrazine oxidation

HzOR is highly attractive as an alternative anode reaction for hydrogen generation due to its exceptionally low potential of −0.33 V vs. RHE.81 Furthermore, HzOR produces only N2 and H2, which are non-toxic, and since industrial wastewater contains hydrazine, HzOR is even more environmentally friendly. However, noble metals such as Pt and Ru are almost inevitable for high performance in HzOR.82,83 Therefore, research on catalyst synthesis for HzOR using 2D structures to reduce the use of noble metals is actively being conducted due to their exceptional properties.

Currently, various methods exist for producing 2D materials. However, most of them face challenges in regulating the thickness of nanosheets.84 Si et al. synthesized ∼3 nm-thick few-layer NbS2 nanosheets from bulk NbS2 via the electrochemical exfoliation method.85 As shown in Fig. 4a, the single-layered NbS2 forms a multi-layered structure stacked by van der Waals forces, a common phenomenon in other transition metal dichalcogenides (TMDs). Fig. 4b shows the transmission electron microscopy (TEM) image of synthesized NbS2 nanosheets. Electrochemical measurements were conducted in 1.0 M KOH with and without 0.5 M hydrazine solution. The LSV curve (Fig. 4c) demonstrates that the HzOR starts at a much lower potential than the Oxygen evolution reaction (OER). Additionally, the bifunctional catalytic activity was evaluated using a two-electrode electrolyzer, employing few-layer NbS2 nanosheets/CP as a bifunctional electrode for HER and HzOR (Fig. 4d). Fig. 4d shows that the few-layer NbS2 nanosheets without 0.5 M hydrazine require about 1.7 V vs. RHE to deliver 10 mA cm−2. However, with 0.5 M hydrazine, it requires only about 0.4 V vs. RHE to produce 10 mA cm−2.


image file: d5im00116a-f4.tif
Fig. 4 Various 2D catalysts for the hydrazine oxidation reaction. (a) Structural model of NbS2, (b) TEM image of a few-layer NbS2 nanosheet, (c) LSV curve of a few-layer NbS2 nanosheet for HzOR and OER, and (d) LSV curves of few-layer NbS2 comparing HzOR‖HER and OER‖HER performance, reprinted with permission from ref. 85, copyright 2019 American Chemical Society; (e) Co(OH)2/MoS2 synthesis scheme, (f) LSV curve of Co(OH)2/MoS2 for HzOR and OER, (g) LSV curve of CC, Co(OH)2/CC, MoS2/CC, Co(OH)2/MoS2/CC, and IrO2 for HzOR, (h) Tafel slope value of CC, Co(OH)2/CC, MoS2/CC, and Co(OH)2/MoS2/CC, reprinted with permission from ref. 86, copyright 2023 American Chemical Society; (i) required overpotential to reach 10 mV cm−2, (j) LSV curves after 5 k, 10 k CV cycles and chronoamperometry test, (k) structural model of a two-electrode flow cell, and (l) comparison of efficiency for hydrogen generation with and without hydrazine, reprinted with permission from ref. 87, copyright 2024 Wiley-VCH GmbH.

Cheng et al. reported Co(OH)2/MoS2 heterostructure electrocatalysts for HzOR, synthesized via hydrothermal and electrodeposition methods.86 The synthesis process is shown in Fig. 4e. First, MoS2 nanosheets were synthesized on the carbon cloth (CC) via a hydrothermal method. Then, the Co(OH)2 nanosheets were introduced onto MoS2/CC via the electrodeposition method, referred to as Co(OH)2/MoS2/CC. Fig. 4f indicates the comparative LSV curves for OER and HzOR. Co(OH)2/MoS2/CC exhibits high catalytic performance for HzOR. To be specific, Co(OH)2/MoS2/CC requires 1.58 V vs. RHE for OER, whereas only 0.18 V vs. RHE for HzOR is needed to deliver 100 mA cm−2, resulting in a voltage reduction of 1.4 V vs. RHE. The oxidation peak at about 1.1 V vs. RHE originated from the oxidation of Co2+ to Co3+. Additionally, Co(OH)2/MoS2/CC exhibits higher performance in HzOR than other samples, including commercial IrO2 (Fig. 4g). At 0.25 V vs. RHE, the current density reached 190 mA cm−2, which is 3.9, 15.6 times higher than MoS2/CC and Co(OH)2/CC, respectively. The Tafel slopes were calculated for Co(OH)2/MoS2/CC, Co(OH)2/CC, and MoS2/CC, revealing values of 47 mV dec−1, 167 mV dec−1, and 121 mV dec−1, respectively. The low Tafel slope of Co(OH)2/MoS2/CC indicates faster reaction kinetics due to the formation of the heterostructure with 2D material.

Zhu et al. synthesized a bifunctional catalyst Ruc/NiFe-LDH by anchoring a Ru cluster on 2D-structured NiFe LDH through the formation of a Ru–O–Ni/Fe bridge.87 The Ru–O–Ni/Fe bridge induces a widening of the d-band width in the Ru cluster. As the d-band width widens, the d-band center of the Ru cluster shifts downward. Consequently, more electrons occupy the antibonding molecular orbitals, which weakens the strong adsorption of hydrazine on the Ru surface. The weakened adsorption facilitates enhanced desorption, allowing an effective adsorption–desorption process. The electrochemical measurements are shown in Fig. 4i. While Ruc/NiFe-LDH required only 75 mV vs. RHE to reach 10 mA cm−2, NiFe-LDH, Pt/C, Ni foam, and Ru/NF required 7, −21, −19, and −24 mV vs. RHE in 1 M KOH with 0.3 M hydrazine solution, respectively. This result indicates that the Ruc/NiFe-LDH exhibits superior HzOR activity compared to the other catalysts. The LSV was conducted on Ruc/NiFe-LDH after 5 k and 10 k cyclic voltammetry (CV) cycles (Fig. 4j). There were no apparent differences between before and after CV cycles, indicating the high stability of the catalyst. Additionally, Ruc/NiFe-LDH demonstrated high stability for a 100 h chronoamperometry (CA) test at a current density of 10 mA cm−2 (Fig. 4j, inset). Furthermore, to evaluate the possibility of industrial usage, the electrochemical measurements were conducted on Ruc/NiFe-LDH using a two-electrode flow cell (Fig. 4k). Fig. 4l shows the comparison of economic efficiency for hydrogen generation between water electrolysis and hydrazine-assisted water electrolysis. The hydrazine-assisted water electrolysis exhibited higher H2 productivity (mol m−3 h−1) than water splitting and even lower energy expense (kWh mH2−3), leading to a high product value ($ m−2 h−1) and low electric cost ($ kWh−1).

HzOR has high potential as an alternative OER due to its exceptionally low onset potential. In particular, various 2D structure-based catalysts for HzOR have been reported, demonstrating excellent activity. For example, NbS2 nanosheets synthesized via electrochemical exfoliation, the Co(OH)2/MoS2 heterostructure catalyst exhibiting fast reaction kinetics, and Ruc/NiFe-LDH, in which the Ru cluster is anchored on the NiFe-LDH. These catalysts have all shown high performance not only for HzOR but also for HER, indicating their potential as bifunctional catalysts. Therefore, these advancements in 2D material-based catalysts contribute to achieving highly efficient hydrogen production.

4 2D catalysts for alcohol oxidation (methanol & ethanol)

The alcohol oxidation reactions, particularly MOR and EOR, have emerged as attractive alternatives to the OER in water electrolysis. Alcohol-assisted water electrolysis not only significantly reduces the overall energy input due to the lower oxidation potentials of alcohols, but also enables the co-generation of value-added chemicals such as formate and acetate.88 These benefits make MOR and EOR promising anodic reactions for the development of next-generation electrochemical hydrogen production platforms. In addition to their role in assisted water electrolysis, MOR and EOR have also been widely applied in direct alcohol fuel cells (DAFCs), which are regarded as efficient and portable power sources.89 DAFCs offer several advantages, including high energy density, easy fuel handling, and low operating temperatures.90 However, MOR suffers from severe catalyst poisoning by CO intermediates, which limits the long-term activity of the electrocatalyst.91 Meanwhile, EOR involves a more complex reaction pathway, leading to incomplete oxidation and the formation of by-products such as acetaldehyde and acetic acid.92 To address these challenges, 2D electrocatalysts have been widely explored that can enhance the activity, selectivity, and durability of alcohol oxidation reactions.

Nickel-based catalysts are the most widely explored among transition metals for MOR. Zhu et al. synthesized NiMn and NiFe-LDHs uniformly on nickel foam (NF) using a hydrothermal synthesis.93 Inductively coupled plasma-optical emission spectrometer analysis revealed high-Ni compositions, with Ni/Mn and Ni/Fe atomic ratios of 4.4 and 4.6, respectively. MOR performance of NiMn and NiFe-LDHs was evaluated in 1 M KOH with 3 M methanol solution (Fig. 5a). NiMn-LDH showed superior performance with a lower onset overpotential of 1.30 V vs. RHE at 2[thin space (1/6-em)]mA[thin space (1/6-em)]cm−2 than that of NiFe-LDH (1.37 V vs. RHE). At the higher current density, NiMn-LDH also showed a smaller working potential of 1.41/1.49 vs. RHE at 100/500 mA cm−2 compared to NiFe-LDH (1.45/1.62 V vs. RHE). The ECSA data showed a significant difference in the catalytic activity of the two catalysts for both MOR and OER (Fig. 5b). NiMn-LDH exhibited three times higher activity than NiFe-LDH for MOR but showed sixteen times lower activity for OER. These LSV and ECSA results indicate that NiMn-LDH is more feasible for MOR compared to NiFe-LDH. This difference was elucidated through mechanistic investigations based on density functional theory (DFT) calculations (Fig. 5c). In the case of NiMn-LDH, the MOR pathway consists of a series of exothermic steps that are thermodynamically favorable, including methanol adsorption (reaction (2)), the initial dehydrogenation step (reaction (3)), and formate formation (reaction (5)), which facilitate a rapid and stable reaction process at the early stages. In contrast, the major thermodynamic limitation arises from the formation of *OCH2 in reaction (4), identified as the potential-determining step (PDS). These findings highlight the intrinsic advantages of NiMn-LDH in promoting efficient and active MOR kinetics, providing a reasonable explanation for the superior MOR performance of NiMn-LDH compared to NiFe-LDH.


image file: d5im00116a-f5.tif
Fig. 5 Various 2D catalysts for alcohol oxidation reaction. (a) LSV curves of NiMn, NiFe LDH for MOR and OER, (b) ECSA value and specific activity for MOR, OER, and (c) DFT calculation of the MOR process, reprinted with permission from ref. 93, copyright 2023 Springer Nature Limited; (d) SEM image of CoFe LDH/MoS2/Ni3S2/NF Astructure, (e) XRD pattern of CoFe LDH/MoS2/Ni3S2/NF structure, (f) Raman spectroscopy of CoFe LDH/MoS2/Ni3S2/NF structure, and (g) comparison of LSV curves for EOR, reprinted with permission from ref. 94, copyright 2024 American Chemical Society; (h) LSV curves of NiFe-LDH/NiFe-HAB/CF and comparison group for OER and MOR, (i) Cdl values of NiFe-LDH/NiFe-HAB/CF, NiFe-HABCF, and NiFe-LDH/CF, (j) 1H NMR spectrum of the electrolyte after electrolysis for 1 h at 60 mA cm−2 in 1.0 m KOH + 3.0 m CH3OH solution, and (k) stability test and yield rate, yield, and Faradaic efficiency for HCOO- and H2, reprinted with permission from ref. 95, 2023 Wiley-VCH GmbH.

Li et al. developed a heterojunction structure composed of CoFe LDH needles on MoS2/Ni3S2/NF nanoarrays for EOR, which also contains nickel.94 Although the catalytic performance was assessed for both EOR and UOR, this review will specifically address its application in EOR. The one-dimensional (1D) CoFe LDH needles possess a large surface area and abundant active sites, offering advantages similar to those of 2D materials. CoFe LDH/MoS2/Ni3S2/NF structure incorporates both 2D MoS2 and 1D CoFe LDH, contributing to a large surface area that enhances the overall electrocatalytic performance, as evidenced by the SEM image (Fig. 5d). In contrast to CoFe LDH and Ni3S2, no peaks of MoS2 were observed in the XRD patterns, which is presumably due to a low loading amount or poor crystallinity (Fig. 5e). In this case, the presence of MoS2 can be confirmed through other characterization techniques such as Raman spectroscopy or XPS. The Raman spectrum exhibits two strong characteristic peaks at 378 and 404 cm−1, corresponding to the E2g1 and A1g modes, respectively, assigned to MoS2, thereby confirming its presence (Fig. 5f). In 1 M KOH with 0.34 M ethanol solution, CoFe LDH/MoS2/Ni3S2/NF exhibits the best performance for EOR among those catalysts showing the lowest potential of 1.484 V vs. RHE at 50 mA cm−2 (Fig. 5g). These results suggest that catalysts combining 1D and 2D structures can exhibit excellent performance in alcohol oxidation.

NiFe-LDH is considered a highly effective OER catalyst due to its layered structure and high catalytic activity.49 Therefore, when performing methanol electrooxidation to formic acid (MEtF), a type of MOR, using NiFe-LDH, the OER tends to dominate over MEtF within the potential range where OER typically occurs, leading to reduced MEtF activity. Thus, strategies to suppress OER are required. Jiang et al. synthesized NiFe-LDH as both a template and precursor for the growth of NiFe hexylaminobenzene (NiFe-HAB) coordination polymers (CPs) on carbon fibers (CFs), referred to as NiFe-LDH/NiFe-HAB/CF, which also contains Ni.95 By incorporating NiFe-HAB, the OER activity of NiFe-LDH is significantly passivated, thereby enhancing the methanol-to-formate conversion on NiFe-LDH/NiFe-HAB/CF. The LSV results in 1 M KOH solution revealed that the OER activity decreases in the order of NiFe-LDH/CF > NiFe-LDH/NiFe-HAB/CF > NiFe-HAB/CF (Fig. 5h). In contrast, MEtF performance measured in 1 M KOH with 3 M methanol solution demonstrated that NiFe-LDH/NiFe-HAB/CF outperforms the others, with NiFe-LDH/CF showing the next highest activity. NiFe-LDH/NiFe-HAB/CF exhibited a notable potential drop upon the addition of methanol. NiFe-LDH/NiFe-HAB/CF also showed a greater Cdl (3.29 mF cm−2) than that of NiFe-LDH (2.55 mF cm−2), indicating superior surface activity (Fig. 5i). After MOR at 60 mA cm−2 for 1 h, nuclear magnetic resonance (NMR) analysis revealed that formate is the only oxidation product of methanol on the NiFe-LDH/NiFe-HAB/CF electrode (Fig. 5j). This indicates the high selectivity of the catalyst toward the methanol-to-formate conversion. The chronopotentiometry measurement was conducted at 20 mA cm−2 for 28 h (Fig. 5k). During the experiment, value-added formate and hydrogen were co-produced at yield rates of 0.2 and 0.4 mmol h−1 cm−2 with Faradaic efficiencies of 98.0% and 99.3%, respectively, closely matching the theoretical formate[thin space (1/6-em)]:[thin space (1/6-em)]hydrogen = 1[thin space (1/6-em)]:[thin space (1/6-em)]2 molar ratio. In conclusion, the inherently layered nature of NiFe-LDH and the planar structure of the NiFe-HAB CPs categorize NiFe-LDH/NiFe-HAB/CF as a 2D electrocatalyst with extended surface area and abundant active sites, which is considered to contribute to its excellent performance in MEtF.

5 2D catalysts for other organic oxidation

Assisted water electrolysis has been applied not only to alcohols but also to other organic compounds, including glycerol and formic acid.96,97 Glycerol is a byproduct of biofuel production, and its oxidation in assisted water electrolysis has been studied for the production of valuable chemicals such as formic acid. The selectivity of these oxidation reactions is crucial for producing the desired target products. Additionally, formic acid has been explored as a feedstock for assisted water electrolysis due to its lower theoretical potentials compared to those of the OER. Recently, 2D materials have emerged as promising catalysts in these fields due to their intrinsic physicochemical properties, offering new opportunities for the oxidation of such organic compounds.

Huang et al. synthesized a series of 2D conjugated metal–organic frameworks (c-MOFs), denoted as M2[NiPcX8], through a solvothermal process. Co2[NiPcS8] was developed as a representative catalyst for the GOR.98 By systematically varying both the metal cations (M = Co, Ni, Cu) and the ligands (X = S, O, NH), the authors evaluated the influence of electronic structure on GOR performance (Fig. 6a). The PXRD pattern of Co2[NiPcS8] showed characteristic peaks at 2θ = 5.0°, 9.9°, 14.5°, and 25.7°, corresponding to the (100), (200), (300), and (002) lattice planes, respectively (Fig. 6b). These results matched well with DFT-simulated XRD patterns, suggesting an in-plane lattice constant of ∼1.8 nm and an interlayer spacing of ∼0.35 nm. Electrochemical measurements revealed that Co2[NiPcS8] exhibited the highest GOR activity among the M2[NiPcS8] catalysts, with the lowest onset potential of 1.2 V vs. RHE (Fig. 6c). Furthermore, comparison among Co2[NiPcX8] catalysts (X = S, O, NH) demonstrated that Co2[NiPcS8] also delivered superior performance at 10 mA cm−2 (1.35 V vs. RHE), outperforming Co2[NiPcO8] (1.44 V) and Co2[NiPc(NH)8] (1.51 V) (Fig. 6d). These experimental trends were further supported by DFT calculations, which indicated that the free energy barrier for the rate-determining step in GOR was lowest for Co2[NiPcS8] (0.93 eV), compared to 1.00 eV and 1.07 eV for Co2[NiPcO8] and Co2[NiPc(NH)8], respectively. These findings highlight the crucial role of ligand chemistry and metal–ligand interactions in determining the electrocatalytic activity of 2D c-MOFs.


image file: d5im00116a-f6.tif
Fig. 6 Various 2D catalysts for other organic oxidation reactions. (a) Structural model of M2[M′PcX8], (b) XRD patterns of experimental and simulated catalysts, LSV curves of (c) M2[NiPcS8] (M = Cu, Ni, Co) and (d) Co2[NiPcX8] (X = S, O, NH) during the GOR, reprinted with permission from ref. 98, copyright 2024 Wiley-VCH GmbH; (e) HAADF-STEM image of PtSA–NiCo LDH, (f) LSV curves of PtSA–NiCo LDH for OER and GOR, (g) Faradaic efficiency of glycerol-to-formic acid conversion at different applied potentials, and (h) the oxidation process of glycerol, reprinted with permission from ref. 9, copyright 2023 Elsevier; (i) XRD patterns of B–PdCu-c/a bimetallene catalysts, (j) LSV curves comparing FAOR‖HER and OER‖HER performance, (k) calculated density of states of B–PdCu-c/a bimetallene, and (l) the free-energy pathway for the FAOR, reprinted with permission from ref. 100, copyright 2024 American Chemical Society.

Yu et al. developed a bifunctional electrocatalyst by anchoring Pt single atoms (PtSA) onto NiCo layered double hydroxides (NiCo LDHs), forming PtSA–NiCo LDHs.99 The catalyst was synthesized via an electrodeposition process, in which oxygen vacancies within the 2D LDH matrix stabilized the Pt atoms. HAADF-STEM images revealed uniformly distributed bright dots, corresponding to dispersed Pt SAs (Fig. 6e). As shown in Fig. 6f, PtSA–NiCo LDHs exhibited excellent GOR activity, requiring only 1.298 V vs. RHE to reach 100 mA cm−2, significantly lower than the potential for OER (1.568 V). Additionally, at 1.375 V vs. RHE, the catalyst achieved 85% glycerol conversion with a high Faradaic efficiency of 88.7% for formate production (Fig. 6g). Based on product analysis via NMR, a proposed reaction pathway for glycerol-to-formate conversion involved sequential oxidation steps (Fig. 6h).

Zeng et al. developed a PdCu bimetallene featuring abundant crystalline/amorphous (c/a) interfaces and boron doping (B–PdCu-c/a), introduced via a solvothermal process with NaBH4 post-treatment.100 The incorporation of the p-block element boron induced d–sp (d–p) orbital hybridization with Pd-based metals, effectively modulating the electronic structure and enhancing the catalytic activity toward the FAOR. The XRD patterns (Fig. 6i) revealed that, following NaBH4 treatment, the diffraction peaks of B–PdCu-c/a shifted to lower 2θ angles, indicative of lattice expansion due to boron incorporation into the PdCu matrix. Electrochemical measurements under FAOR‖HER and OER‖HER configurations demonstrated the superior performance of the B–PdCu-c/a catalyst. As shown in Fig. 6j, the cell voltage at 10 mA cm−2 in the FAOR‖HER system was significantly reduced to 0.19 V, in contrast to 1.65 V under conventional OER‖HER conditions. DFT calculations revealed that B doping promotes strong d–sp orbital hybridization, resulting in the upward shift of the d-band center and enhanced intermediate adsorption (Fig. 6k). Furthermore, the free energy pathway analysis (Fig. 6l) indicated that B–PdCu-c/a possesses a lower energy barrier (0.15 eV) for the C–H bond cleavage step, the rate-determining step in FAOR, compared to PdCu (0.20 eV), thereby facilitating improved catalytic kinetics.

Various 2D electrocatalysts have demonstrated significant potential for assisted water electrolysis, extending beyond the oxidation of alcohols to include glycerol and formic acid. The studies discussed various strategies to enhance catalytic activity and selectivity, such as metal–ligand modulation in 2D c-MOFs, the incorporation of atomically dispersed noble metals on LDH supports, and electronic structure engineering through dopant-induced orbital hybridization. These approaches reduce the anodic potential compared to the traditional OER. Furthermore, in the case of GOR, they enable the synthesis of value-added products. These developments highlight the promise of 2D catalysts as efficient and tunable materials for electrochemical energy conversion systems.

6 Industrial implementation of assisted water electrolysis

To bridge laboratory-scale research with practical deployment, it is important to consider the cell and reactor configurations employed for industrial hydrogen generation. Among these, the zero-gap electrolyzer system has been most widely adopted, particularly under alkaline or anion-exchange conditions. A typical zero-gap configuration consists of stacked anode and cathode plates (current collectors), gaskets, porous transport layers (PTLs), electrocatalysts, an anion-exchange membrane (AEM), and an electrolyte circulation system such as an integrated peristaltic pump (Fig. 7a and d).129,130 Compared to conventional three-electrode configurations such as H-type or one-pot cells, zero gap two-electrode systems offer distinct advantages for scale-up, including a large effective reaction area, reduced ohmic resistance, and improved energy efficiency.
image file: d5im00116a-f7.tif
Fig. 7 Industrial H2 production using alternative anodic reactions. (a) Schematic of the cell configuration, and (b) Overpotentials of NiO/Co3O4‖NiCoP AEMWE with and without urea under high-current conditions, reprinted with permission from ref. 129, copyright 2025 Springer Nature; (c) stability test of N–Ni–MoO2/NF catalysts at 500 mA cm−2 for 100 h, reprinted with permission from ref. 131, copyright 2023 Elsevier; (d) schematic of the cell configuration, and (e) stability test of Ni(OH)2/NMO catalyst with 0.1 M methanol for over 130 h at high current densities, reprinted with permission from ref. 130, copyright 2024 Wiley-VCH GmbH; (f) stability test of NiMoPx@Ni4P5 pair catalyst at 500 mA cm−2 for 100 h, reprinted with permission from ref. 132, copyright 2024 Wiley-VCH GmbH; (g) Overpotentials of CoFe hydroxide catalysts with and without 0.5[thin space (1/6-em)]M glycerol under high-current conditions, and (h) stability tests of CoFe hydroxide catalyst at 500 and 1000 mA cm−2 for 120 h, reprinted with permission from ref. 133, copyright 2025 Elsevier.

Recent studies have demonstrated that assisted water electrolysis can be successfully integrated into zero-gap electrolyzers. For example, Yang et al. studied AEMWE performance for UOR using NiO/Co3O4 as anode and NiCoP as the cathode.129 The UOR‖HER system achieved energy savings of 210, 230, and 210 mV compared with overall water splitting at 500, 1000, and 1500 mA cm−2, respectively (Fig. 7b). In addition, Qian et al. investigated N–Ni–MoO2/NF catalysts, which exhibited stable operation (Vinitial/Vfinal = 98.3%) at 500 mA cm−2 for 100 h in the presence of urea, with the electrolyte refreshed every 20 h (Fig. 7c).131 Similarly, Ni(OH)2/NMO catalysts coupled with 0.1 M methanol achieve stable operation for over 130 h at high current densities (Fig. 7e).130 NiMoPx@Ni4P5 pairs sustain 500 mA cm−2 for 100 h under 6.0 M KOH and 1.0 M CH3OH at 65 °C, with electrolyte refreshed every 20 h (Fig. 7f).132 Moreover, CoFe hydroxide catalysts not only exhibit significantly reduced overpotentials when operated with glycerol (Fig. 7g) but also maintain stability at 500 and 1000 mA cm−2 for 120 h (Fig. 7h).133 Collectively, these demonstrations highlight the scalability and durability of 2D catalyst-based assisted electrolysis under industrially relevant conditions.

Nevertheless, further challenges remain in reactor design (e.g., flow-field optimization, gas–liquid management), long-term electrode stability under fluctuating loads, and techno-economic viability considering feedstock availability and by-product valorization. Importantly, unlike conventional OER, assisted water electrolysis inherently relies on the continuous supply of organic or nitrogenous feedstocks. Therefore, evaluation at the industrial scale must necessarily include feedstock delivery and management processes, such as reactant feeding strategies, concentration control, and by-product separation. Without such considerations, performance metrics cannot be directly extrapolated to practical hydrogen generation.

7 Summary and outlook

Assisted water electrolysis has emerged as a promising strategy for not only reducing the energy demand of hydrogen production but also enabling the co-generation of valuable chemicals and the remediation of environmental pollutants. A wide range of alternative oxidation reactions, including HzOR, UOR, MOR, EOR, GOR, and FAOR, have been extensively studied in combination with diverse catalytic systems to enhance overall electrolysis efficiency.

In particular, 2D materials such as LDHs, TMDs, MXenes, metallenes, and graphene-based materials have received considerable attention owing to their high surface area, electrical conductivity, and tunable catalytic activity. This review provides a comprehensive overview of recent developments in 2D-based catalysts for various assisted anodic reactions in water electrolysis. Table 2 provides specific electrochemical performance values such as the Tafel slope, potential (at 10, 50, and 100 mA cm−2), and long-term stability. Each parameter plays a major role in evaluating the performance of catalysts. In comparison, Fig. 8 highlights the electrochemical performance of these reactions, showing that HzOR exhibits significantly lower overpotentials than UOR, MOR, and GOR. This indicates that HzOR is a highly favorable anodic alternative from an energy efficiency perspective.

Table 2 Comparison of 2D-based catalysts for assisted water electrolysis
Reaction Catalyst Electrolyte Potential at 10 mA (V vs. RHE) Potential at 50 mA (V vs. RHE) Potential at 100 mA (V vs. RHE) Tafel slope (mV dec−1) Target products/FE (product) Stability performance (CV, CA, CP) Ref.
HzOR NbS2 1 M KOH + 0.5 M hydrazine ∼0.37 ∼0.53 10 mA cm−2 85
10 h
Co(OH)2/MoS2/CC 1 M KOH + 0.4 M hydrazine 0.18 11 ∼0.14 V vs. RHE 86
11 h
N–NiZnCu LDH/rGO 1 M KOH + 0.5 M hydrazine 0.014 0.021 0.045 20 3000 cycle 101
CV
Ruc/NiFe-LDH 1 M KOH + 0.3 M hydrazine −0.075 40.4 10[thin space (1/6-em)]000 cycle 87
CV
AR-Co(OH)2/Ti3C2(OH)x 1 M KOH + 0.5 M hydrazine 0.995 56 1.15 V vs. RHE 102
50 h
S–CuNiCo-LDH 1 M KOH + 0.02 M hydrazine ∼0.29 ∼0.65 73.3 0.7 vs. RHE 103
20 h
CC@WS2/Ru-450 1 M KOH + 0.5 M hydrazine −0.074 42.2 10 mA cm−2 104
100 h
NiCo@C/MXene/CF 1 M KOH + 0.5 M hydrazine −0.096 −0.025 73 100 mA cm−2 105
30 h
UOR CoFe LDH/MoS2/Ni3S2/NF 1 M KOH + 0.5 M urea 1.423 50 mA cm−2 94
50 h
N-NiZnCu LDH/rGO 1 M KOH + 0.5 M urea 1.304 1.426 1.467 29 3000 cycle 101
CV
FeCuCoNiZn-LDH/CC 1 M KOH + 0.33 M urea 1.326 120 1.53 V vs. RHE 106
∼70 h
Ni0.67Co0.33(OH)2/CC 1 M KOH + 0.5 M urea 1.23 40 1.40 V vs. RHE 107
10 h
MoS4-LDH/NF 1 M KOH + 0.33 M urea 1.34 29 10 mA cm−2 108
24 h
FQD/CCoNi-LDH/NF 1 M KOH + 0.5 M urea 1.36 1.42 17 1.40 V vs. RHE 109
15 h
NiS/MoS2@CC 1 M KOH + 0.5 M urea 1.36 1.38 24.2 1.36 vs. RHE 110
∼27 h
MoSe2/NiSe2 1 M KOH + 0.5 M urea 1.35 1.41 1.47 68 100 mA cm−2 77
100 h
Mo–FeNi LDH 1 M KOH + 0.33 M urea 1.31 16.9 50 mA cm−2 80
12 h
MoS2/Ni3S2 1 M KOH + 0.5 M urea 1.44 3000 cycle 111
CV
CoS2–MoS2 1 M KOH + 0.5 M urea 1.29 32 ∼1.29 V vs. RHE 112
30 h
NiMoV LDH 1 M KOH + 0.33 M urea 1.4 24.29 1.40 V vs. RHE 113
30 h
NiS2–MoS2 1 M KOH + 0.33 M urea ∼1.54 29.9 1.6 V vs. RHE 114
10 h
Co1Mn1 LDH/NF 1 M KOH + 0.33 M urea 1.326 73 1.326 V vs. RHE 115
10 h
NiS/MoS2@FCP 1 M KOH + 0.4 M urea 1.42 1.43 31 1000 cycle 116
CV
NiOOH/(LDH/α-FeOOH) 1 M KOH + 0.33 M urea 1.35 1.37 1.4 30.1 79
MOR NiMn-LDHs 1 M KOH + 3 M CH3OH 1.33 1.41 39.4 Formate/97.3 100 mA cm−2 93
20 h
NiFe-LDH/NiFe-HAB/CF 1 M KOH + 3 M CH3OH 1.44 1.46 Formate/98 20 mA cm−2 95
28 h
Ni0.33Co0.67(OH)2/NF 1 M KOH + 0.5 M CH3OH 1.33 17 Formate/∼100 1.35 V vs. RHE 117
20 h
NiSe/MoSe2/CC 1 M KOH + 1.0 M CH3OH 1.38 14 Formate/− 1.41 V vs. RHE 118
120 h
NiFexP@NiCo-LDH/CC 1 M KOH + 0.5 M CH3OH 1.42 1.425 Formate/∼100 0.96 V vs. RHE 119
10 h
NiCo-LDH-E 1 M KOH + 1.0 M CH3OH ∼1.40 28.7 120
Cu0.33CoCo-LDH/CF 1 M KOH + 3 M CH3OH 1.28 67.8 Formate/99 20 mA cm−2 121
24 h
CoxP@NiCo-LDH/NF 1 M KOH + 0.5 M CH3OH 1.24 1.32 1.34 Formate/∼100 1.35 V vs. RHE 122
20 h
EOR CoFe LDH/MoS2/Ni3S2/NF 1 M KOH + 0.34 M C2H5OH 1.484 50 mA cm−2 94
50 h
Pt/N–Ti3C2Tx 0.5 M H2SO4 + 1.0 M C2H5OH 275.92 123
NiAl-LDH-NSs 1 M NaOH + 1.0 M C2H5OH ∼1.444 124
NiCo-LDH-E 1 M KOH + 1.0 M C2H5OH ∼1.40 50.13 120
U-NiFe LDH 1 M KOH + 1.0 M C2H5OH 1.344 32.7 100 mA cm−2 125
10 h
PtSe2 0.1 M KOH + 0.5 M C2H5OH ∼0.58 177 126
Pd/DB-Ti3C2 1 M KOH + 1.0 M C2H5OH ∼0.51 158 2000 cycle 127
CV
GOR Co2[NiPcS8] 1 M KOH + 0.1 M glycerol 1.35 102 Formate/>85 10 mA cm−2 98
12 h
PtSA–NiCo LDH/NF 1 M KOH + 0.1 M glycerol 1.298 68.6 Formate/88.7 99
NiVRu-LDHs NAs/NF 1 M KOH + 0.1 M glycerol 1.24 1.3 1.33 40.7 Formate/97 1.40 V vs. RHE 96
10 h
FAOR PtTe2 NSs/C 0.5 M H2SO4 + 0.5 M formic acid ∼0.3 208 97
Pd/Ti3C2Tx-rGO 0.5 M H2SO4 + 0.5 M formic acid ∼0.311 128



image file: d5im00116a-f8.tif
Fig. 8 Comparative analysis of electrochemical performance in assisted water electrolysis systems utilizing various molecular feedstocks.

These insights underscore the pivotal role of feedstock selection and 2D catalyst design in enhancing the energy efficiency and practicality of assisted electrolysis systems. Building upon this foundation, several future research directions are proposed to accelerate the development and implementation of 2D-based electrocatalysts in real-world AWE applications:

(1) Various 2D-based materials have been employed as catalyst supports to improve distribution and structural stability for main catalysts, and synergistic effects. However, when used alone, these materials often exhibit limited intrinsic catalytic activity, indicating the need for coupling with additional active catalysts, such as noble metals. To overcome this limitation, high-entropy (HE) strategies have recently gained traction in the field of catalysis. High-entropy alloy (HEA) catalysts, characterized by multiple principal elements, offer four main effects, including the cocktail effect, lattice distortion, sluggish diffusion, and high configurational entropy, that enable tunable electronic structures and improved catalytic behavior.134,135 In particular, high-entropy transition metal dichalcogenides (HE-TMDs) as 2D catalysts have demonstrated enhanced activity for oxidation reactions without requiring noble metals, owing to their modulated electronic environments.136–139 This approach is not limited to TMDs, other 2D materials (LDHs, MXenes, and metallenes) can also be tailored using high-entropy designs to optimize electrocatalytic performances.140,141 Moreover, the inherent thermodynamic stability of high-entropy configurations contributes to improved durability, bringing such materials closer to practical application and commercialization.

(2) It is well-established that edge sites in 2D catalysts exhibit significantly higher catalytic activity than their in-plane counterparts, due to their unsaturated coordination and favorable adsorption properties.56 While conventional doping strategies have primarily focused on in-plane substitution, edge-site doping is emerging as a promising approach to enhance performance further. Doping at edge sites not only activates additional reaction centers but also offers thermodynamic advantages, as the formation energy of dopants is generally lower at edge sites compared to basal planes.142 This enhances both the stability and activity of doped atoms, enabling synergistic improvements in overall catalytic performance.

(3) A lot of studies on 2D-based catalysts report high catalytic performance under laboratory conditions, particularly at elevated concentrations of the molecular feedstocks. While one of the key advantages of assisted water electrolysis is the remediation of organic and nitrogen-containing pollutants, the concentrations used in experimental setups often far exceed those found in real wastewater streams.143,144 To enhance the practical relevance of these studies, electrochemical testing under realistic, low-concentration conditions is essential. In addition, most current research is conducted under strongly acidic or alkaline environments, which are not always compatible with industrial or environmental systems.24,145–149 Future efforts should focus on catalyst evaluation under neutral or near-neutral pH conditions to better reflect practical application scenarios and long-term operational stability.

(4) Furthermore, in real-world wastewater streams, multiple organic and nitrogen-containing compounds often coexist, presenting a complex chemical environment that differs significantly from simplified laboratory conditions.150,151 However, most current studies on assisted water electrolysis focus on single-component feedstocks, limiting the applicability of these systems to practical situations.76,152 Future research should therefore prioritize the development of multi-component oxidation strategies, wherein catalysts are designed to selectively and simultaneously activate different types of molecules. This could involve tandem or cascade oxidation pathways, where intermediates from one oxidation reaction participate in subsequent transformations, potentially enabling the formation of more complex and valuable chemical products. Such approaches would not only improve energy efficiency and pollutant removal under realistic conditions but also expand the functional scope of assisted electrolysis by coupling hydrogen production with advanced molecular synthesis.

2D materials have demonstrated exceptional promise as catalysts for assisted water electrolysis, enabling energy-efficient hydrogen production while simultaneously achieving environmental remediation and the synthesis of value-added chemicals. This review has provided a comprehensive overview of the fundamental principles, representative oxidation reactions, and recent advances in the design of 2D-based catalysts. By bridging the gap between theoretical understanding and practical implementation, these insights contribute to the global goal of sustainable hydrogen technologies and carbon neutrality.

Author contributions

Hyojung Lim: conceptualization, data curation, investigation, methodology, visualization, writing—original draft preparation, writing—review, and editing. Seonghyeon Park: conceptualization, investigation, writing—original draft preparation, writing—review, and editing. Jinuk Choi: conceptualization, investigation, writing—original draft preparation, writing—review, and editing. Junho Shim: resources, visualization, writing—review, and editing. Subramani Surendran: supervision, validation, visualization, writing—review, and editing. Uk Sim: conceptualization, investigation, visualization, supervision, funding acquisition, validation, writing—original draft preparation, writing—review, and editing.

Conflicts of interest

There are no conflicts to declare.

Data availability

No primary research results, software, or code have been included, and no new data were generated or analysed as part of this review.

Acknowledgements

This work was supported by the Korea Institute of Energy Technology Evaluation and Planning (KETEP) and the Ministry of Trade, Industry and Energy (MOTIE) of the Republic of Korea (No. 20224000000320 and No. RS-2025-07852969). This work was also supported by the Korea Institute of Energy Technology Evaluation and Planning (KETEP) grant funded by the Korea government (MOTIE) (No. RS-2024-00421291, Clean Hydrogen and Ammonia Innovation Research Center).

References

  1. S. C. Jesudass, S. Surendran, D. J. Moon, S. Shanmugapriya, J. Y. Kim, G. Janani, K. Veeramani, S. Mahadik, I. G. Kim, P. Jung, G. Kwon, K. Jin, J. K. Kim, K. Hong, Y. I. Park, T.-H. Kim, J. Heo and U. Sim, Defect engineered ternary metal spinel-type Ni-Fe-Co oxide as bifunctional electrocatalyst for overall electrochemical water splitting, J. Colloid Interface Sci., 2024, 663, 566–576 CrossRef CAS PubMed.
  2. G. Janani, S. Surendran, D.-K. Lee, S. Shanmugapriya, H. Lee, Y. Subramanian and U. Sim, Aggregation induced edge sites actuation of 3D MoSe2/rGO electrocatalyst for high-performing water splitting system, Aggregate, 2024, 5, e430 CrossRef CAS.
  3. S. Mahadik, S. Surendran, D. J. Moon, J. Y. Kim, G. Janani, S. C. Jesudass, K. Veeramani, H. Choi, S. Shanmugapriya, I. G. Kim, P. Jung, Y. I. Park, J. Heo, T.-H. Kim, K. Hong and U. Sim, Structurally engineered highly efficient electrocatalytic performance of 3-dimensional Mo/Ni chalcogenides for boosting overall water splitting performance, Chemosphere, 2024, 352, 141233 CrossRef CAS PubMed.
  4. Y.-H. Wang, L. Li, J. Shi, M.-Y. Xie, J. Nie, G.-F. Huang, B. Li, W. Hu, A. Pan and W.-Q. Huang, Oxygen defect engineering promotes synergy between adsorbate evolution and single lattice oxygen mechanisms of OER in transition metal-based (oxy)hydroxide, Adv. Sci., 2023, 10, 2303321 CrossRef CAS PubMed.
  5. K. Zhang and R. Zou, Advanced transition metal-based OER electrocatalysts: Current status, opportunities, and challenges, Small, 2021, 17, 2100129 CrossRef CAS.
  6. K. Jin, J. Park, J. Lee, K. D. Yang, G. K. Pradhan, U. Sim, D. Jeong, H. L. Jang, S. Park, D. Kim, N.-E. Sung, S. H. Kim, S. Han and K. T. Nam, Hydrated manganese(II) phosphate (Mn3(PO4)2·3H2O) as a water oxidation catalyst, J. Am. Chem. Soc., 2014, 136, 7435–7443 CrossRef CAS PubMed.
  7. G. Janani, S. Surendran, D. J. Moon, P. S. Ramesh, J. Y. Kim, Y. Lim, K. Veeramani, S. Mahadik, S. C. Jesudass, J. Choi, I. G. Kim, P. Jung, H. Choi, G. Kwon, K. Jin, J. K. Kim, Y. I. Park, J. Heo, K. Hong, Y. S. Kang and U. Sim, Ambipolar nature accelerates dual-functionality on Ni/Ni3N@NC for simultaneous hydrogen and oxygen evolution in electrochemical water splitting system, Adv. Sustainable Syst., 2024, 8, 2400059 CrossRef CAS.
  8. Y. Lim, S. Surendran, W. So, S. Shanmugapriya, C. Jo, G. Janani, H. Choi, H. S. Han, H. Choi, Y.-H. Yun, T.-H. Kim, M.-J. Kim, K. Jin, J. K. Kim and U. Sim, In situ decorated Cu2FeSnS4 nanosheet arrays for low voltage hydrogen production through the ammonia oxidation reaction, Mater. Chem. Front., 2023, 7, 5843–5857 RSC.
  9. H. Choi, S. Surendran, D. Kim, Y. Lim, J. Lim, J. Park, J. K. Kim, M.-K. Han and U. Sim, Boosting eco-friendly hydrogen generation by urea-assisted water electrolysis using spinel M2GeO4 (M = Fe, Co) as an active electrocatalyst, Environ. Sci.: Nano, 2021, 8, 3110–3121 RSC.
  10. P. Arora, K. Bhadauriya, L. Singh, A. Goyal, S. Verma, B. Singh and A. Draksharapu, 3d–2p–5d orbital synergy in electrocatalytic hydrazine oxidation assisted water splitting with industrial scale current density, Inorg. Chem., 2025, 64, 5069–5076 CrossRef CAS PubMed.
  11. F. Arshad, T. u. Haq, A. Khan, Y. Haik, I. Hussain and F. Sher, Multifunctional porous NiCo bimetallic foams toward water splitting and methanol oxidation-assisted hydrogen production, Energy Convers. Manage., 2022, 254, 115262 CrossRef CAS.
  12. F. Arshad, T.u. Haq, I. Hussain and F. Sher, Recent advances in electrocatalysts toward alcohol-assisted, energy-saving hydrogen production, ACS Appl. Energy Mater., 2021, 4, 8685–8701 CrossRef CAS.
  13. B. Zhu, Z. Liang and R. Zou, Designing advanced catalysts for energy conversion based on urea oxidation reaction, Small, 2020, 16, 1906133 CrossRef CAS PubMed.
  14. Y. Tian, Z. Mao, L. Wang and J. Liang, Green chemistry: Advanced electrocatalysts and system design for ammonia oxidation, Small Struct., 2023, 4, 2200266 CrossRef CAS.
  15. Q. Qian, J. Zhang, J. Li, Y. Li, X. Jin, Y. Zhu, Y. Liu, Z. Li, A. El-Harairy, C. Xiao, G. Zhang and Y. Xie, Artificial heterointerfaces achieve delicate reaction kinetics towards hydrogen evolution and hydrazine oxidation catalysis, Angew. Chem., Int. Ed., 2021, 133, 6049–6058 CrossRef.
  16. J. Ren, L. Xing, T. Lai and Y. Zhao, Boosting hydrogen production coupling with electrochemical methanol oxidation to formate using monolayered layered double hydroxide nanosheets, Ind. Eng. Chem. Res., 2023, 62, 17553–17561 CrossRef.
  17. G. Wu, X. Dong, J. Mao, G. Li, C. Zhu, S. Li, A. Chen, G. Feng, Y. Song, W. Chen and W. Wei, Anodic glycerol oxidation to formate facilitating cathodic hydrogen evolution with earth-abundant metal oxide catalysts, Chem. Eng. J., 2023, 468, 143640 CrossRef CAS.
  18. C. Jo, S. Surendran, M.-C. Kim, T.-Y. An, Y. Lim, H. Choi, G. Janani, S. Cyril Jesudass, D. Jun Moon, J. Kim, J. Young Kim, C. Hyuck Choi, M. Kim, J. Kyu Kim and U. Sim, Meticulous integration of N and C active sites in Ni2P electrocatalyst for sustainable ammonia oxidation and efficient hydrogen production, Chem. Eng. J., 2023, 463, 142314 CrossRef CAS.
  19. J. Choi, H. Lim, S. Surendran, X. Lu, K. Jin, H. Park, H.-Y. Jung and U. Sim, Dichalcogenides as emerging electrocatalysts for efficient ammonia synthesis: A focus on mechanisms and theoretical potentials, Adv. Funct. Mater., 2025, 35, 2422585 CrossRef CAS.
  20. R. Zhang, S. Zhang, Y. Guo, C. Li, J. Liu, Z. Huang, Y. Zhao, Y. Li and C. Zhi, A Zn–nitrite battery as an energy-output electrocatalytic system for high-efficiency ammonia synthesis using carbon-doped cobalt oxide nanotubes, Energy Environ. Sci., 2022, 15, 3024–3032 RSC.
  21. X. Qiang, Y. Yao, J. Yin, P. Da, Z. Mu, K. Shen, Y. Sun, Y. Zhang, P. Li, Z. Li, P. Xi and C.-H. Yan, Activating La–O–Ni bridge in ordered macroporous interface for electrochemical urea wastewater purification, Angew. Chem., Int. Ed., 2025, 64, e202424014 CrossRef CAS PubMed.
  22. V. S. Protsenko, L. S. Bobrova, T. E. Butyrina and O. D. Sukhatskyi, Thermodynamics of electrochemical urea oxidation reaction coupled with cathodic hydrogen evolution reaction in an alkaline solution: Effect of carbonate formation, Int. J. Hydrogen Energy, 2024, 59, 354–358 CrossRef CAS.
  23. V. S. Protsenko, Thermodynamic aspects of urea oxidation reaction in the context of hydrogen production by electrolysis, Int. J. Hydrogen Energy, 2023, 48, 24207–24211 CrossRef CAS.
  24. A. Song, Y. Wei, X. Jin, Y. Ma, Y. Wang and J. Yang, Decoupled water reduction and hydrazine oxidation by fast proton transport MoO3 redox mediator for hydrogen production, Small, 2025, 21, 2407783 CrossRef CAS PubMed.
  25. F. H. Gül, H. A. Deveci, A. Deveci, O. Akyıldırım and M. L. Yola, Hydrazine imprinted electrochemical sensor based on cobalt-barium stannate nanoparticles incorporated-functionalized MWCNTs nanocomposite for hydrazine determination in tap water samples, Microchim. Acta, 2025, 192, 124 CrossRef PubMed.
  26. J. Shi, Q. Sun, J. Chen, W. Zhu, T. Cheng, M. Ma, Z. Fan, H. Yang, F. Liao, M. Shao and Z. Kang, Nitrogen contained rhodium nanosheet catalysts for efficient hydrazine oxidation reaction, Appl. Catal., B, 2024, 343, 123561 CrossRef CAS.
  27. A. B. Anderson and H. A. Asiri, Reversible potentials for steps in methanol and formic acid oxidation to CO2; adsorption energies of intermediates on the ideal electrocatalyst for methanol oxidation and CO2 reduction, Phys. Chem. Chem. Phys., 2014, 16, 10587–10599 RSC.
  28. M. Singh, H. M. Sharma, J. Kaur, D. K. Das, M. Ubaidullah, R. K. Gupta and A. Kumar, Engineering of electrocatalysts for methanol oxidation reaction: Recent advances and future challenges, Mol. Catal., 2024, 557, 113982 CAS.
  29. C. Lamy, B. Guenot, M. Cretin and G. Pourcelly, Kinetics analysis of the electrocatalytic oxidation of methanol inside a DMFC working as a PEM electrolysis cell (PEMEC) to generate clean hydrogen, Electrochim. Acta, 2015, 177, 352–358 CrossRef CAS.
  30. Z. H. N. Al-Azri, W.-T. Chen, A. Chan, V. Jovic, T. Ina, H. Idriss and G. I. N. Waterhouse, The roles of metal co-catalysts and reaction media in photocatalytic hydrogen production: Performance evaluation of M/TiO2 photocatalysts (M=Pd, Pt, Au) in different alcohol–water mixtures, J. Catal., 2015, 329, 355–367 CrossRef CAS.
  31. P. Kalluru, G. A. B. Azizar, R. K. Pramadewandaru, T. G. Kim, J. Yu, D. I. Kang, J. Jung, Y. W. Lee and J. W. Hong, Synergistic function of Bi and B in interstitial ternary Pd–Bi–B alloy nanocrystals for highly active and durable electrochemical ethanol oxidation, J. Mater. Chem. A, 2025, 13, 10019–10027 RSC.
  32. Y. Wang, M. Zheng, Y. Li, C. Ye, J. Chen, J. Ye, Q. Zhang, J. Li, Z. Zhou, X.-Z. Fu, J. Wang, S.-G. Sun and D. Wang, p–d orbital hybridization induced by a monodispersed Ga site on a Pt3Mn nanocatalyst boosts ethanol electrooxidation, Angew. Chem., Int. Ed., 2022, 61, e202115735 CrossRef CAS PubMed.
  33. Y. Liu, M. Wei, D. Raciti, Y. Wang, P. Hu, J. H. Park, M. Barclay and C. Wang, Electro-oxidation of ethanol using Pt3Sn alloy nanoparticles, ACS Catal., 2018, 8, 10931–10937 CrossRef CAS.
  34. W. Trisap, S. Photaworn and P. Khongprom, Impact of hydrocyclone geometry on glycerol separation efficiency in biodiesel purification, Int. J. Chem. Eng., 2025, 2025, 1758062 CrossRef CAS.
  35. M. Simões, S. Baranton and C. Coutanceau, Electrochemical valorisation of glycerol, ChemSusChem, 2012, 5, 2106–2124 CrossRef PubMed.
  36. X. Tan, J. Wang, Y. Xiao, Y. Guo, W. He, B. Du, H. Cui and C. Wang, Engineering topological and chemical disorder in Pd sites for record-breaking formic acid electrocatalytic oxidation, Adv. Mater., 2025, 37, 2414283 CrossRef CAS PubMed.
  37. P. A. Kempler and A. C. Nielander, Reliable reporting of Faradaic efficiencies for electrocatalysis research, Nat. Commun., 2023, 14, 1158 CrossRef CAS PubMed.
  38. Y. Chen, J. Meng, M. Xu, L. Qiao, D. Liu, Y. Kong, X. Hu, Q. Liu, M. Chen, S. Lyu, R. Tong and H. Pan, Adaptive active site tuning for superior OER and UOR on Ir–Ni3N catalyst, Adv. Funct. Mater., 2025, 35, 2413474 CrossRef CAS.
  39. S. Xu, X. Ruan, M. Ganesan, J. Wu, S. K. Ravi and X. Cui, Transition metal-based catalysts for urea oxidation reaction (UOR): Catalyst design strategies, applications, and future perspectives, Adv. Funct. Mater., 2024, 34, 2313309 CrossRef CAS.
  40. A. Kumar and Q. Xu, Two-dimensional layered materials as catalyst supports, ChemNanoMat, 2018, 4, 28–40 CrossRef CAS.
  41. J. Choi, S. Im, J. Choi, S. Surendran, D. J. Moon, J. Y. Kim, J. K. Kim and U. Sim, Recent advances in 2D structured materials with defect-exploiting design strategies for electrocatalysis of nitrate to ammonia, Energy Mater., 2024, 4, 400020 CAS.
  42. S. Roy, A. Joseph, X. Zhang, S. Bhattacharyya, A. B. Puthirath, A. Biswas, C. S. Tiwary, R. Vajtai and P. M. Ajayan, Engineered two-dimensional transition metal dichalcogenides for energy conversion and storage, Chem. Rev., 2024, 124, 9376–9456 CrossRef CAS PubMed.
  43. Q. Wang and D. O'Hare, Recent advances in the synthesis and application of layered double hydroxide (LDH) nanosheets, Chem. Rev., 2012, 112, 4124–4155 CrossRef CAS PubMed.
  44. M. Xu and M. Wei, Layered double hydroxide-based catalysts: Recent advances in preparation, structure, and applications, Adv. Funct. Mater., 2018, 28, 1802943 CrossRef.
  45. M. Gong, Y. Li, H. Wang, Y. Liang, J. Z. Wu, J. Zhou, J. Wang, T. Regier, F. Wei and H. Dai, An advanced Ni–Fe layered double hydroxide electrocatalyst for water oxidation, J. Am. Chem. Soc., 2013, 135, 8452–8455 CrossRef CAS PubMed.
  46. W. Choi, N. Choudhary, G. H. Han, J. Park, D. Akinwande and Y. H. Lee, Recent development of two-dimensional transition metal dichalcogenides and their applications, Mater. Today, 2017, 20, 116–130 CrossRef CAS.
  47. R. Yang, Y. Fan, Y. Zhang, L. Mei, R. Zhu, J. Qin, J. Hu, Z. Chen, Y. Hau Ng, D. Voiry, S. Li, Q. Lu, Q. Wang, J. C. Yu and Z. Zeng, 2D transition metal dichalcogenides for photocatalysis, Angew. Chem., Int. Ed., 2023, 62, e202218016 CrossRef CAS PubMed.
  48. H.-W. Choi, D. H. Seo, J. W. Heo, S.-I. Kim and T. Kim, Thickness-dependent electrical and optoelectrical properties of SnSe2 field-effect transistors, Electron. Mater. Lett., 2025, 21, 154–161 CrossRef CAS.
  49. P. M. Bodhankar, P. B. Sarawade, G. Singh, A. Vinu and D. S. Dhawale, Recent advances in highly active nanostructured NiFe LDH catalyst for electrochemical water splitting, J. Mater. Chem. A, 2021, 9, 3180–3208 RSC.
  50. Y. Lin, H. Wang, C.-K. Peng, L. Bu, C.-L. Chiang, K. Tian, Y. Zhao, J. Zhao, Y.-G. Lin, J.-M. Lee and L. Gao, Co-induced electronic optimization of hierarchical NiFe LDH for oxygen evolution, Small, 2020, 16, 2002426 CrossRef CAS PubMed.
  51. R. Yang, Y. Zhou, Y. Xing, D. Li, D. Jiang, M. Chen, W. Shi and S. Yuan, Synergistic coupling of CoFe-LDH arrays with NiFe-LDH nanosheet for highly efficient overall water splitting in alkaline media, Appl. Catal., B, 2019, 253, 131–139 CrossRef CAS.
  52. S. Nagappan, A. Karmakar, R. Madhu, S. S. Selvasundarasekar, S. Kumaravel, K. Bera, H. N. Dhandapani, D. Sarkar, S. M. Yusuf and S. Kundu, 2D CoFe-LDH nanosheet-incorporated 1D microfibers as a high-performance OER electrocatalyst in neutral and alkaline media, ACS Appl. Energy Mater., 2022, 5, 11483–11497 CrossRef CAS.
  53. Q. Fu, J. Han, X. Wang, P. Xu, T. Yao, J. Zhong, W. Zhong, S. Liu, T. Gao, Z. Zhang, L. Xu and B. Song, 2D transition metal dichalcogenides: Design, modulation, and challenges in electrocatalysis, Adv. Mater., 2021, 33, 1907818 CrossRef CAS PubMed.
  54. R. Lv, J. A. Robinson, R. E. Schaak, D. Sun, Y. Sun, T. E. Mallouk and M. Terrones, Transition metal dichalcogenides and beyond: Synthesis, properties, and applications of single- and few-layer nanosheets, Acc. Chem. Res., 2015, 48, 56–64 CrossRef CAS PubMed.
  55. B. H. Kim, D. W. Kim, S. H. Kwon, H. Yoon and Y. J. Yoon, Contact and interface engineering of MoS2-based photodetectors using electron-beam irradiation, Electron. Mater. Lett., 2023, 19, 564–570 CrossRef CAS.
  56. Z. Zheng, L. Yu, M. Gao, X. Chen, W. Zhou, C. Ma, L. Wu, J. Zhu, X. Meng, J. Hu, Y. Tu, S. Wu, J. Mao, Z. Tian and D. Deng, Boosting hydrogen evolution on MoS2 via co-confining selenium in surface and cobalt in inner layer, Nat. Commun., 2020, 11, 3315 CrossRef CAS PubMed.
  57. B. Chamlagain, S. S. Withanage, A. C. Johnston and S. I. Khondaker, Scalable lateral heterojunction by chemical doping of 2D TMD thin films, Sci. Rep., 2020, 10, 12970 CrossRef CAS PubMed.
  58. M. Chhowalla, Z. Liu and H. Zhang, Two-dimensional transition metal dichalcogenide (TMD) nanosheets, Chem. Soc. Rev., 2015, 44, 2584–2586 RSC.
  59. X. Ren, G. Liao, Z. Li, H. Qiao, Y. Zhang, X. Yu, B. Wang, H. Tan, L. Shi, X. Qi and H. Zhang, Two-dimensional MOF and COF nanosheets for next-generation optoelectronic applications, Coord. Chem. Rev., 2021, 435, 213781 CrossRef CAS.
  60. M. Nemiwal, T. C. Zhang and D. Kumar, Graphene-based electrocatalysts: Hydrogen evolution reactions and overall water splitting, Int. J. Hydrogen Energy, 2021, 46, 21401–21418 CrossRef CAS.
  61. H. Wang and J.-M. Lee, Recent advances in structural engineering of MXene electrocatalysts, J. Mater. Chem. A, 2020, 8, 10604–10624 RSC.
  62. H. Kim, Multimodal MXene artificial synapses realizing optoelectronic, olfactory, and tactile neuromorphic memory in wearable devices, Electron. Mater. Lett., 2025, 1–13 Search PubMed.
  63. Y. Zhang, X. Wang, M. Chen, P. He and Z. Kong, Two-dimensional leafy Fe/N-doped carbon nanomaterials derived from vitamin C-modified ZIF-L for efficient oxygen reduction reaction, Electron. Mater. Lett., 2024, 20, 592–602 CrossRef CAS.
  64. S. Jing, X. Peng, S. Li, L. Yuan, S. Lu, Y. Zhang and H. Fan, Synergistic realization of fast polysulfide redox kinetics and stable lithium anode in Li-S battery from CoNi-MOF/MXene derived CoNi@TiO2/C heterostructure, Chin. Chem. Lett., 2024, 110732 Search PubMed.
  65. C. S. Manikandababu, S. Navaneethan, M. I. S. Kumar, S. Ramkumar, K. Muthukannan and P. Siva Karthik, Construction of MoS2@RGO hybrid catalyst: An efficient and highly stable electrocatalyst for enhanced hydrogen generation reactions, Chem. Phys. Impact, 2025, 10, 100874 CrossRef.
  66. A. Hanan, D. Shu, U. Aftab, D. Cao, A. J. Laghari, M. Y. Solangi, M. I. Abro, A. Nafady, B. Vigolo, A. Tahira and Z. H. Ibupoto, Co2FeO4@rGO composite: Towards trifunctional water splitting in alkaline media, Int. J. Hydrogen Energy, 2022, 47, 33919–33937 CrossRef CAS.
  67. A. Hanan, F. Bibi, G. Elsa, A. Numan, R. Walvekar and M. Khalid, A comprehensive review of double transition metal MXene (Mo2Ti2C3Tx) in energy storage, conversion, and harvesting, Mater. Today Energy, 2025, 51, 101905 CrossRef CAS.
  68. M. N. Lakhan, A. Hanan, Y. Wang, H. K. Lee and H. Arandiyan, Integrated MXene and metal oxide electrocatalysts for the oxygen evolution reaction: Synthesis, mechanisms, and advances, Chem. Sci., 2024, 15, 15540–15564 RSC.
  69. A. Hanan, H. T. A. Awan, F. Bibi, R. R. R. Sulaiman, W. Y. Wong, R. Walvekar, S. Singh and M. Khalid, MXenes and heterostructures-based electrocatalysts for hydrogen evolution reaction: Recent developments and future outlook, J. Energy Chem., 2024, 92, 176–206 CrossRef CAS.
  70. Q. Wang, Y. Cheng, H. B. Tao, Y. Liu, X. Ma, D.-S. Li, H. B. Yang and B. Liu, Long-term stability challenges and opportunities in acidic oxygen evolution electrocatalysis, Angew. Chem., Int. Ed., 2023, 62, e202216645 CrossRef CAS PubMed.
  71. Y. Tang, C. Yang, X. Xu, Y. Kang, J. Henzie, W. Que and Y. Yamauchi, MXene nanoarchitectonics: Defect-engineered 2D MXenes towards enhanced electrochemical water splitting, Adv. Energy Mater., 2022, 12, 2103867 CrossRef CAS.
  72. A. Iqbal, J. Hong, T. Y. Ko and C. M. Koo, Improving oxidation stability of 2D MXenes: Synthesis, storage media, and conditions, Nano Convergence, 2021, 8, 9 CrossRef CAS PubMed.
  73. H. Li, N. Cheng, Y. Zheng, X. Zhang, H. Lv, D. He, M. Pan, F. Kleitz, S. Z. Qiao and S. Mu, Oxidation stability of nanographite materials, Adv. Energy Mater., 2013, 3, 1176–1179 CrossRef CAS.
  74. X. Gao, S. Zhang, P. Wang, M. Jaroniec, Y. Zheng and S.-Z. Qiao, Urea catalytic oxidation for energy and environmental applications, Chem. Soc. Rev., 2024, 53, 1552–1591 RSC.
  75. S.-K. Geng, Y. Zheng, S.-Q. Li, H. Su, X. Zhao, J. Hu, H.-B. Shu, M. Jaroniec, P. Chen, Q.-H. Liu and S.-Z. Qiao, Nickel ferrocyanide as a high-performance urea oxidation electrocatalyst, Nat. Energy, 2021, 6, 904–912 CrossRef CAS.
  76. L. Wang, Y. Zhu, Y. Wen, S. Li, C. Cui, F. Ni, Y. Liu, H. Lin, Y. Li, H. Peng and B. Zhang, Regulating the local charge distribution of Ni active sites for the urea oxidation reaction, Angew. Chem., Int. Ed., 2021, 60, 10577–10582 CrossRef CAS PubMed.
  77. Y. Chen, J. Ge, Y. Wang, X. Zhao, F. Zhang and X. Lei, Nanostructured MoSe2/NiSe2 electrocatalysts with heterojunctions for hydrogen evolution coupling urea oxidation, ACS Appl. Nano Mater., 2024, 7, 12091–12100 CrossRef CAS.
  78. Q. Lin, G. Nan, D. Fu and L. Xie, Oxygen evolution reaction on NiFe-LDH/(Ni,Fe)OOH: Theoretical insights into the effects of electronic structure and spin-state evolution, Phys. Chem. Chem. Phys., 2025, 27, 4926–4933 RSC.
  79. M. Cai, Q. Zhu, X. Wang, Z. Shao, L. Yao, H. Zeng, X. Wu, J. Chen, K. Huang and S. Feng, Formation and stabilization of NiOOH by introducing α-FeOOH in LDH: Composite electrocatalyst for oxygen evolution and urea oxidation reactions, Adv. Mater., 2023, 35, 2209338 CrossRef CAS PubMed.
  80. J.-M. Huo, Y. Wang, J.-N. Xue, W.-Y. Yuan, Q.-G. Zhai, M.-C. Hu, S.-N. Li and Y. Chen, High-valence metal doping induced lattice expansion for M–FeNi LDH toward enhanced urea oxidation electrocatalytic activities, Small, 2024, 20, 2305877 CrossRef CAS PubMed.
  81. T. Y. Burshtein, Y. Yasman, L. Muñoz-Moene, J. H. Zagal and D. Eisenberg, Hydrazine oxidation electrocatalysis, ACS Catal., 2024, 14, 2264–2283 CrossRef CAS.
  82. W. Zhu, A. Gandi Naidu, Q. Wu, H. Yan, M. Zhao, Z. Wang and H. Liang, Simultaneous electrocatalytic hydrogen production and hydrazine removal from acidic waste water, Chem. Eng. Sci., 2022, 258, 117769 CrossRef CAS.
  83. X. Fu, D. Cheng, A. Zhang, J. Zhou, S. Wang, X. Zhao, J. Chen, P. Sautet, Y. Huang and X. Duan, Ag–Ru interface for highly efficient hydrazine assisted water electrolysis, Energy Environ. Sci., 2024, 17, 2279–2286 RSC.
  84. J. N. Coleman, M. Lotya, A. O'Neill, S. D. Bergin, P. J. King, U. Khan, K. Young, A. Gaucher, S. De, R. J. Smith, I. V. Shvets, S. K. Arora, G. Stanton, H.-Y. Kim, K. Lee, G. T. Kim, G. S. Duesberg, T. Hallam, J. J. Boland, J. J. Wang, J. F. Donegan, J. C. Grunlan, G. Moriarty, A. Shmeliov, R. J. Nicholls, J. M. Perkins, E. M. Grieveson, K. Theuwissen, D. W. McComb, P. D. Nellist and V. Nicolosi, Two-dimensional nanosheets produced by liquid exfoliation of layered materials, Science, 2011, 331, 568–571 CrossRef CAS PubMed.
  85. J. Si, Q. Zheng, H. Chen, C. Lei, Y. Suo, B. Yang, Z. Zhang, Z. Li, L. Lei, Y. Hou and K. Ostrikov, Scalable production of few-layer niobium disulfide nanosheets via electrochemical exfoliation for energy-efficient hydrogen evolution reaction, ACS Appl. Mater. Interfaces, 2019, 11, 13205–13213 CrossRef CAS PubMed.
  86. X. Cheng and Y. Tong, Interface coupling of cobalt hydroxide/molybdenum disulfide heterostructured nanosheet arrays for highly efficient hydrazine-assisted hydrogen generation, ACS Sustainable Chem. Eng., 2023, 11, 3219–3227 CrossRef CAS.
  87. Y. Zhu, Y. Chen, Y. Feng, X. Meng, J. Xia and G. Zhang, Constructing Ru–O–TM bridge in NiFe-LDH enables high current hydrazine-assisted H2 production, Adv. Mater., 2024, 36, 2401694 CrossRef CAS PubMed.
  88. Q. Wang, J. Liu, W. Zhang, T. Li, Y. Wang, H. Li and A. Cabot, Branch-regulated palladium–antimony nanoparticles boost ethanol electro-oxidation to acetate, Inorg. Chem., 2022, 61, 6337–6346 CrossRef CAS PubMed.
  89. K. Elsaid, S. Abdelfatah, A. M. Abdel Elabsir, R. J. Hassiba, Z. K. Ghouri and L. Vechot, Direct alcohol fuel cells: Assessment of the fuel's safety and health aspects, Int. J. Hydrogen Energy, 2021, 46, 30658–30668 CrossRef CAS.
  90. A. Ahmadi, H. Ebrahimifar and M. B. Askari, Alcohol electrooxidation on three-component NiO/La2O3/MWCNTs catalyst for DAFC application, Electrochem. Commun., 2025, 176, 107936 CrossRef CAS.
  91. H. Yang, C. Li, L. Lü, Z. Li, S. Zhang, Z. Huang, R. Ma, S. Liu, M. Ge, W. Zhou and X. Yuan, Electronegativity- induced cobalt-doped platinum hollow nanospheres with high CO tolerance for efficient methanol oxidation reaction, J. Colloid Interface Sci., 2025, 678, 300–308 CrossRef CAS PubMed.
  92. W. Zhang, Y. Zhao, J. Li, Y. Miao, P. He, Z. Wu, Y. Zhang, G.-R. Xu and L. Wang, C–C bond cleavage driven by lattice oxygen during ethanol oxidation process, Adv. Funct. Mater., 2025, 2421763 CrossRef CAS.
  93. B. Zhu, B. Dong, F. Wang, Q. Yang, Y. He, C. Zhang, P. Jin and L. Feng, Unraveling a bifunctional mechanism for methanol-to-formate electro-oxidation on nickel-based hydroxides, Nat. Commun., 2023, 14, 1686 CrossRef CAS PubMed.
  94. C. Li, H. Pang, R. Xu, J. Fan, E. Liu and T. Sun, CoFe-layered double hydroxide needles on MoS2/Ni3S2 nanoarrays for applications as catalysts for hydrogen evolution and oxidation of organic chemicals, ACS Appl. Nano Mater., 2024, 7, 6449–6459 CrossRef CAS.
  95. S. Jiang, T. Xiao, C. Xu, S. Wang, H.-Q. Peng, W. Zhang, B. Liu and Y.-F. Song, Passivating oxygen evolution activity of NiFe-LDH through heterostructure engineering to realize high-efficiency electrocatalytic formate and hydrogen co-production, Small, 2023, 19, 2208027 CrossRef CAS PubMed.
  96. Q. Qian, X. He, Z. Li, Y. Chen, Y. Feng, M. Cheng, H. Zhang, W. Wang, C. Xiao, G. Zhang and Y. Xie, Electrochemical biomass upgrading coupled with hydrogen production under industrial-level current density, Adv. Mater., 2023, 35, 2300935 CrossRef CAS PubMed.
  97. J. Bao, H. Sun, W. Yan, S. Liu, W. Xu, J. Fan, C. Zhan, W. Liu, X. Huang and N. Chen, Large-scalable CO-tolerant ultrathin PtTe2 nanosheets for formic acid oxidation, Small Methods, 2025, 2402155 CrossRef CAS PubMed.
  98. X. Huang, M. Wang, H. Zhong, X. Li, H. Wang, Y. Lu, G. Zhang, Y. Liu, P. Zhang, R. Zou, X. Feng and R. Dong, Metal–phthalocyanine-based two-dimensional conjugated metal–organic frameworks for electrochemical glycerol oxidation reaction, Angew. Chem., Int. Ed., 2025, 64, e202416178 CrossRef CAS PubMed.
  99. H. Yu, W. Wang, Q. Mao, K. Deng, Z. Wang, Y. Xu, X. Li, H. Wang and L. Wang, Pt single atom captured by oxygen vacancy-rich NiCo layered double hydroxides for coupling hydrogen evolution with selective oxidation of glycerol to formate, Appl. Catal., B, 2023, 330, 122617 CrossRef CAS.
  100. S. Zeng, D. Qu, H. Sun, Y. Chen, J. Wang, Y. Zheng, J. Pan, J. Cao and C. Li, 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 CrossRef CAS PubMed.
  101. S. Hu, Y. Tan, C. Feng, H. Wu, J. Zhang and H. Mei, Synthesis of N doped NiZnCu-layered double hydroxides with reduced graphene oxide on nickel foam as versatile electrocatalysts for hydrogen production in hybrid-water electrolysis, J. Power Sources, 2020, 453, 227872 CrossRef CAS.
  102. K. Xiang, Y. Wang, Z. Zhuang, J. Zou, N. Li, D. Wang, T. Zhai and J. Jiang, Self-healing of active site in Co(OH)2/MXene electrocatalysts for hydrazine oxidation, J. Mater. Sci. Technol., 2024, 203, 108–117 CrossRef CAS.
  103. W. Liu, J. Xie, Y. Guo, S. Lou, L. Gao and B. Tang, Sulfurization-induced edge amorphization in copper–nickel–cobalt layered double hydroxide nanosheets promoting hydrazine electro-oxidation, J. Mater. Chem. A, 2019, 7, 24437–24444 RSC.
  104. J. Li, Y. Li, J. Wang, C. Zhang, H. Ma, C. Zhu, D. Fan, Z. Guo, M. Xu, Y. Wang and H. Ma, Elucidating the critical role of ruthenium single atom sites in water dissociation and dehydrogenation behaviors for robust hydrazine oxidation-boosted alkaline hydrogen evolution, Adv. Funct. Mater., 2022, 32, 2109439 CrossRef CAS.
  105. F. Sun, J. Qin, Z. Wang, M. Yu, X. Wu, X. Sun and J. Qiu, Energy-saving hydrogen production by chlorine-free hybrid seawater splitting coupling hydrazine degradation, Nat. Commun., 2021, 12, 4182 CrossRef CAS PubMed.
  106. M. Hao, J. Chen, J. Chen, K. Wang, J. Wang, F. Lei, P. Hao, X. Sun, J. Xie and B. Tang, Lattice-disordered high-entropy metal hydroxide nanosheets as efficient precatalysts for bifunctional electro-oxidation, J. Colloid Interface Sci., 2023, 642, 41–52 CrossRef CAS PubMed.
  107. G. Wang and Z. Wen, Self-supported bimetallic Ni–Co compound electrodes for urea- and neutralization energy-assisted electrolytic hydrogen production, Nanoscale, 2018, 10, 21087–21095 RSC.
  108. A. Nadeema, V. Kashyap, R. Gururaj and S. Kurungot, [MoS4]2−-intercalated NiCo-Layered double hydroxide nanospikes: An efficiently synergized material for urine to direct H2 generation, ACS Appl. Mater. Interfaces, 2019, 11, 25917–25927 CrossRef CAS PubMed.
  109. Y. Feng, X. Wang, J. Huang, P. Dong, J. Ji, J. Li, L. Cao, L. Feng, P. Jin and C. Wang, Decorating CoNi layered double hydroxides nanosheet arrays with fullerene quantum dot anchored on Ni foam for efficient electrocatalytic water splitting and urea electrolysis, Chem. Eng. J., 2020, 390, 124525 CrossRef CAS.
  110. C. Gu, G. Zhou, J. Yang, H. Pang, M. Zhang, Q. Zhao, X. Gu, S. Tian, J. Zhang, L. Xu and Y. Tang, NiS/MoS2 Mott-Schottky heterojunction-induced local charge redistribution for high-efficiency urea-assisted energy-saving hydrogen production, Chem. Eng. J., 2022, 443, 136321 CrossRef CAS.
  111. Y. Ren, C. Wang, W. Duan, L. Zhou, X. Pang, D. Wang, Y. Zhen, C. Yang and Z. Gao, MoS2/Ni3S2 Schottky heterojunction regulating local charge distribution for efficient urea oxidation and hydrogen evolution, J. Colloid Interface Sci., 2022, 628, 446–455 CrossRef CAS PubMed.
  112. C. Li, Y. Liu, Z. Zhuo, H. Ju, D. Li, Y. Guo, X. Wu, H. Li and T. Zhai, Local charge distribution engineered by Schottky heterojunctions toward urea electrolysis, Adv. Energy Mater., 2018, 8, 1801775 CrossRef.
  113. Z. Wang, W. Liu, J. Bao, Y. Song, X. She, Y. Hua, G. Lv, J. Yuan, H. Li and H. Xu, Modulating electronic structure of ternary NiMoV LDH nanosheet array induced by doping engineering to promote urea oxidation reaction, Chem. Eng. J., 2022, 430, 133100 CrossRef CAS.
  114. S. Wang, L. Zhao, J. Li, X. Tian, X. Wu and L. Feng, High valence state of Ni and Mo synergism in NiS2-MoS2 hetero-nanorods catalyst with layered surface structure for urea electrocatalysis, J. Energy Chem., 2022, 66, 483–492 CrossRef CAS.
  115. Z. Wang, Y. Hu, W. Liu, L. Xu, M. Guan, Y. Zhao, J. Bao and H. Li, Manganese-modulated cobalt-based layered double hydroxide grown on nickel foam with 1D–2D–3D heterostructure for highly efficient oxygen evolution reaction and urea oxidation reaction, Chem. – Eur. J., 2020, 26, 9382–9388 CrossRef CAS PubMed.
  116. Y. Zheng, P. Tang, X. Xu and X. Sang, POM derived UOR and HER bifunctional NiS/MoS2 composite for overall water splitting, J. Solid State Chem., 2020, 292, 121644 CrossRef CAS.
  117. M. Li, X. Deng, K. Xiang, Y. Liang, B. Zhao, J. Hao, J.-L. Luo and X.-Z. Fu, Value-added formate production from selective methanol oxidation as anodic reaction to enhance electrochemical hydrogen cogeneration, ChemSusChem, 2020, 13, 914–921 CrossRef CAS PubMed.
  118. X. Peng, S. Xie, X. Wang, C. Pi, Z. Liu, B. Gao, L. Hu, W. Xiao and P. K. Chu, Energy-saving hydrogen production by the methanol oxidation reaction coupled with the hydrogen evolution reaction co-catalyzed by a phase separation induced heterostructure, J. Mater. Chem. A, 2022, 10, 20761–20769 RSC.
  119. Y. Zhang, X. Wu, G. Fu, F. Si, X.-Z. Fu and J.-L. Luo, NiFexP@NiCo-LDH nanoarray bifunctional electrocatalysts for coupling of methanol oxidation and hydrogen production, Int. J. Hydrogen Energy, 2022, 47, 17150–17160 CrossRef CAS.
  120. H. Wang, L. Wang, Y. Jia, X. Li, H. Yang, X. Zhu, Q. Bu and Q. Liu, Engineering lattice planes of NiCo-LDH ultrathin sheets for boosting methanol/ethanol oxidation performance, Inorg. Chem., 2023, 62, 11256–11264 CrossRef CAS PubMed.
  121. B. Liu, T. Xiao, X. Sun, H.-Q. Peng, X. Wang, Y. Zhao, W. Zhang and Y.-F. Song, Hierarchical trace copper incorporation activated cobalt layered double hydroxide as a highly selective methanol conversion electrocatalyst to realize energy-matched photovoltaic-electrocatalytic formate and hydrogen co-production, J. Mater. Chem. A, 2022, 10, 19649–19661 RSC.
  122. M. Li, X. Deng, Y. Liang, K. Xiang, D. Wu, B. Zhao, H. Yang, J.-L. Luo and X.-Z. Fu, CoxP@NiCo-LDH heteronanosheet arrays as efficient bifunctional electrocatalysts for co-generation of value-added formate and hydrogen with less-energy consumption, J. Energy Chem., 2020, 50, 314–323 CrossRef.
  123. T. T. Huynh, Q. Huynh, A. Q. K. Nguyen and H. Q. Pham, Strong component-interaction in N-doped 2D Ti3C2T-supported Pt electrocatalyst for acidic ethanol oxidation reaction, Adv. Sustainable Syst., 2025, 9, 2400995 CrossRef CAS.
  124. L. Xu, Z. Wang, X. Chen, Z. Qu, F. Li and W. Yang, Ultrathin layered double hydroxide nanosheets with Ni(III) active species obtained by exfoliation for highly efficient ethanol electrooxidation, Electrochim. Acta, 2018, 260, 898–904 CrossRef CAS PubMed.
  125. H. Wang, M. Huo, Y. Liang, K. Qin, Q. Li, W. Liu, Z. Xing and J. Chang, Significant reduction of anode reaction overpotential in alkaline water electrolysis by ultrathin NiFe-layered double hydroxide in ethanol-added electrolyte, ChemCatChem, 2025, 17, e202401950 CrossRef CAS.
  126. L. Wang, J. Qi, Y. Zhang, Y. Dai, K. Bao, W. Wang, J. Wu, C. Ma, Z. Yin, C. Ma, Y. Chen, J. Bao, R. Ye, Y. Liu, Z. Lin, Z. Wang and Q. He, Surface engineering of PtSe2 crystal for highly efficient electrocatalytic ethanol oxidation, Adv. Mater., 2025, 37, 2502047 CrossRef CAS PubMed.
  127. Z. Chen, F. Jing, M. Luo, X. Wu, H. Fu, S. Xiao, B. Yu, D. Chen, X. Xiong and Y. Jin, Local coordination and electronic interactions of Pd/MXene via dual-atom codoping with superior durability for efficient electrocatalytic ethanol oxidation, Carbon Energy, 2024, 6, e443 CrossRef CAS.
  128. C. Yang, H. He, Q. Jiang, X. Liu, S. P. Shah, H. Huang and W. Li, Pd nanocrystals grown on MXene and reduced graphene oxide co-constructed three-dimensional nanoarchitectures for efficient formic acid oxidation reaction, Int. J. Hydrogen Energy, 2021, 46, 589–598 CrossRef CAS.
  129. C. Yang, H. Pang, X. Li, X. Zheng, T. Wei, X. Ma, Q. Wang, C. Wang, D. Wang and B. Xu, Scalable electrocatalytic urea wastewater treatment coupled with hydrogen production by regulating adsorption behavior of urea molecule, Nano-Micro Lett., 2025, 17, 159 CrossRef CAS PubMed.
  130. Z. Li, Y. Zheng, W. Zu, L. Dong and L. Y. S. Lee, Molybdate-modified NiOOH for efficient methanol-assisted seawater electrolysis, Adv. Sci., 2025, 12, 2410911 CrossRef CAS PubMed.
  131. G. Qian, T. Lu, Y. Wang, H. Xu, X. Cao, Z. Xie, C. Chen and D. Min, N-induced compressive strain in Ni-MoO2 heterostructure with micro-nano array for improving high-current-output urea-assisted water electrolysis performance, Chem. Eng. J., 2024, 480, 147993 CrossRef CAS.
  132. B. Zhu, J. Xiong, S. Wu, K. You, B. Sun, Y. Liu, M. Chen, P. Jin and L. Feng, Core@Shell heterostructured NiMoPx@Ni5P4 nanorod arrays promoting direct electro-oxidation of methanol and hydrogen evolution under industry conditions, Adv. Funct. Mater., 2024, 34, 2407236 CrossRef CAS.
  133. S. M. Hu, Y. N. Chen, J. K. Liu, X. Y. Zhang, W. H. Jiang, Y. Zhou, H. Y. Yuan, P. F. Liu, Q. Guo, H. G. Yang, F. Wang and G. Yu, Inhibiting the peroxidation of Co(III) oxyhydroxide for stable and ampere-level glycerol oxidation, Appl. Catal., A, 2026, 381, 125811 CrossRef CAS.
  134. S. Im, D. Kim, S. Surendran, J. Choi, D. J. Moon, J. Y. Kim, H. Lee, D.-H. Nam and U. Sim, High entropy alloy: From theoretical evaluation to electrocatalytic activity of hydrogen evolution reaction, Curr. Opin. Electrochem., 2023, 39, 101293 CrossRef CAS.
  135. D. Kim, S. Surendran, Y. Jeong, Y. Lim, S. Im, S. Park, J. Y. Kim, S. Kim, T.-H. Kim, B. Koo, K. Jin and U. Sim, Electrosynthesis of low Pt-loaded high entropy catalysts for effective hydrogen evolution with improved acidic durability, Adv. Mater. Technol., 2023, 8, 2200882 CrossRef CAS.
  136. J. Cavin, A. Ahmadiparidari, L. Majidi, A. S. Thind, S. N. Misal, A. Prajapati, Z. Hemmat, S. Rastegar, A. Beukelman, M. R. Singh, K. A. Unocic, A. Salehi-Khojin and R. Mishra, 2D high-entropy transition metal dichalcogenides for carbon dioxide electrocatalysis, Adv. Mater., 2021, 33, 2100347 CrossRef CAS PubMed.
  137. Z. Wang, X. Chen, Y. Ding, X. Zhu, Z. Sun, H. Zhou, X. Li, W. Yang, J. Liu, R. He, J. Luo, T. Yu, M. Zeng and L. Fu, Synthesis of two-dimensional high-entropy transition metal dichalcogenide single crystals, J. Am. Chem. Soc., 2025, 147, 1392–1398 CrossRef CAS PubMed.
  138. S. K. Nemani, M. Torkamanzadeh, B. C. Wyatt, V. Presser and B. Anasori, Functional two-dimensional high-entropy materials, Commun. Mater, 2023, 4, 16 CrossRef CAS.
  139. J. Qu, A. Elgendy, R. Cai, M. A. Buckingham, A. A. Papaderakis, H. de Latour, K. Hazeldine, G. F. S. Whitehead, F. Alam, C. T. Smith, D. J. Binks, A. Walton, J. M. Skelton, R. A. W. Dryfe, S. J. Haigh and D. J. Lewis, A Low-temperature synthetic route toward a high-entropy 2D hexernary transition metal dichalcogenide for hydrogen evolution electrocatalysis, Adv. Sci., 2023, 10, 2204488 CrossRef CAS PubMed.
  140. F. Wang, P. Zou, Y. Zhang, W. Pan, Y. Li, L. Liang, C. Chen, H. Liu and S. Zheng, Activating lattice oxygen in high-entropy LDH for robust and durable water oxidation, Nat. Commun., 2023, 14, 6019 CrossRef CAS PubMed.
  141. S. K. Nemani, B. Zhang, B. C. Wyatt, Z. D. Hood, S. Manna, R. Khaledialidusti, W. Hong, M. G. Sternberg, S. K. R. S. Sankaranarayanan and B. Anasori, High-entropy 2D carbide MXenes: TiVNbMoC3 and TiVCrMoC3, ACS Nano, 2021, 15, 12815–12825 CrossRef CAS PubMed.
  142. Y. Ying, K. Fan, Z. Lin and H. Huang, Facing the “Cutting Edge:” Edge site engineering on 2D materials for electrocatalysis and photocatalysis, Adv. Mater., 2025, 37, 2418757 CrossRef CAS PubMed.
  143. A. Shaban, M. E. Basiouny and O. A. AboSiada, Comparative study of the removal of urea by electrocoagulation and electrocoagulation combined with chemical coagulation in aqueous effluents, Sci. Rep., 2024, 14, 30605 CrossRef CAS PubMed.
  144. Y. Chen, H. Chen, Z. Chen, Z. Zhu and X. Wang, The nitrogen removal performance and mechanisms for urea wastewater by simultaneous urea hydrolysis, partial nitritation and anammox in one reactor, J. Cleaner Prod., 2022, 332, 130124 CrossRef CAS.
  145. Z. Wu, M. Fan, H. Jiang, J. Dai, K. Liu, R. Hu, S. Qin, W. Xu, Y. Yao and J. Wan, Harnessing the unconventional cubic phase in 2D LaNiO3 perovskite for highly efficient urea oxidation, Angew. Chem., Int. Ed., 2025, 64, e202413932 CrossRef CAS PubMed.
  146. P. Mannu, R. K. Dharman, T. T. T. Nga, A. Mariappan, Y.-C. Shao, H. Ishii, Y.-C. Huang, A. Kandasami, T. H. Oh, W.-C. Chou, C.-L. Chen, J.-L. Chen and C.-L. Dong, Tuning of oxygen vacancies in Co3O4 electrocatalyst for effectiveness in urea oxidation and water splitting, Small, 2025, 21, 2403744 CrossRef PubMed.
  147. Y. Wang, G. Qian, Z. Xie, H. Yu, L. Li, J. Li, C. Chen, M. Lu and P. Tsiakaras, Strengthening CuNiCoMo medium-entropy alloy by tuning local lattice tensile strain and built-in electric field for boosting urea oxidation reaction, Appl. Catal., B, 2025, 365, 124841 CrossRef CAS.
  148. P. Wang, X. Gao, M. Zheng, M. Jaroniec, Y. Zheng and S. Z. Qiao, Urine electrooxidation for energy–saving hydrogen generation, Nat. Commun., 2025, 16, 2424 CrossRef CAS PubMed.
  149. H. Qin, G. Lin, J. Zhang, X. Cao, W. Xia, H. Yang, K. Yuan, T. Jin, Q. Wang and L. Jiao, Enhanced cooperative generalized compressive strain and electronic structure engineering in W-Ni3N for efficient hydrazine oxidation facilitating H2 production, Adv. Mater., 2025, 37, 2417593 CrossRef CAS PubMed.
  150. M. Akhtar, M. Sarfraz, M. Ahmad, N. Raza and L. Zhang, Use of low-cost adsorbent for waste water treatment: Recent progress, new trend and future perspectives, Desalin. Water Trea, 2025, 321, 100914 CrossRef CAS.
  151. P. Rajasulochana and V. Preethy, Comparison on efficiency of various techniques in treatment of waste and sewage water – A comprehensive review, Resour.-Effic. Technol., 2016, 2, 175–184 Search PubMed.
  152. T. Yu, G. Liu, T. Nie, Z. Wu, Z. Song, X. Sun and Y.-F. Song, Pt-loaded CoFe-layered double hydroxides for simultaneously driving HER and HzOR, ACS Catal., 2024, 14, 14937–14946 CrossRef CAS.

Footnote

These authors contributed equally to this work.

This journal is © Institute of Process Engineering of CAS 2025
Click here to see how this site uses Cookies. View our privacy policy here.