Shun Li
*a,
Zhanpeng Zhua,
Yuqiao Zhanga,
Yong Liub,
Xinyue Zhang
*ab and
Kwun Nam Hui
*c
aInstitute of Quantum and Sustainable Technology (IQST), School of Chemistry and Chemical Engineering, Jiangsu University, Zhenjiang, 212013, China. E-mail: shun@ujs.edu.cn
bFoshan (Southern China) Institute for New Materials, Foshan, 528200, China. E-mail: zhangxinyue@fscinm.com
cJoint Key Laboratory of the Ministry of Education, Institute of Applied Physics and Materials Engineering, University of Macau, Avenida da Universidade, Taipa, Macau SAR, China. E-mail: bizhui@um.edu.mo
First published on 21st April 2025
Hydrogen peroxide (H2O2) plays a crucial role in various industrial sectors and everyday applications. Given the energy-intensive nature of the current anthraquinone process for its production, the quest for cost-effective, efficient, and stable catalysts for H2O2 synthesis is paramount. A promising sustainable approach lies in small-scale, decentralized electrochemical methods. Carbon nanomaterials have emerged as standout candidates, offering low costs, high surface areas, excellent conductivity, and adjustable electronic properties. This review presents a thorough examination of recent strides in engineering strategies of carbon-based nanomaterials for enhanced electrochemical H2O2 generation. It delves into tailored microstructures (e.g., 1D, 2D, porous architectures), defect/surface engineering (e.g., edge sites, heteroatom doping, surface modification), and heterostructure assembly (e.g., semiconductor–carbon composites, single-atom, dual-single-atom catalysts). Moreover, the review explores structure–performance interplays in these carbon electrocatalysts, drawing from advanced experimental analyses and theoretical models to unveil the mechanisms governing selective electrocatalytic H2O2 synthesis. Lastly, this review identifies challenges and charts future research avenues to propel carbon electrocatalysts towards greener and more effective H2O2 production methods.
Wider impactThe sustainable production of hydrogen peroxide (H2O2) is pivotal for a broad range of industrial and environmental applications, yet its current large-scale synthesis via the anthraquinone process remains energy-intensive and environmentally taxing. This review highlights recent advances in carbon-based electrocatalysts for decentralized, electrochemical H2O2 generation—a promising alternative that operates under mild conditions with minimal environmental footprint. The development of efficient, stable, and cost-effective carbon-based materials for electrochemical H2O2 production has profound implications for sustainable chemical manufacturing, decentralized wastewater treatment, and green energy applications. By providing a comprehensive analysis of cutting-edge catalyst engineering strategies, mechanistic insights, and structure–performance relationships, this work lays the foundation for the rational design of next-generation carbon-based electrocatalysts. Additionally, this review underscores critical challenges and future research directions, serving as a valuable roadmap for the transition toward scalable, eco-friendly H2O2 synthesis. The insights presented herein will not only advance fundamental scientific understanding but also accelerate practical implementation in industrial and environmental sectors. |
Since the pioneering work on nitrogen-doped carbon nanotubes for efficient electrochemical oxygen reduction reaction (ORR) in 2009,42 carbon-based metal-free electrocatalysts have garnered considerable research attention.43,44 Various carbon nanomaterials, such as graphene,45 carbon nanotubes (CNTs),46 carbon nanofibers/wires,47 carbon dots,48 fullerene,49 and porous carbon,50 have demonstrated great potential as catalysts for electrochemical production of H2O2 via the 2e− ORR pathway.51 These materials offer great advantages like low cost, abundant availability, tunable nanostructures, excellent electrical conductivity, and adjustable electronic properties.52,53 However, several challenges remain in optimizing these catalysts for H2O2 production, including (1) high overpotential, low catalytic activity and selectivity compared to precious metal catalysts; (2) unclear mechanisms involving active sites and reaction pathways; (3) poor stability and durability in acidic or alkaline environments; (4) difficulty of integration with electrochemical systems for scalable production.
To overcome these inherent limitations, many strategies have been developed in the past few decades. Several important review papers have been published on carbon-based nanomaterials for H2O2 production.53–57 As a rapidly growing field, recent studies have demonstrated notable advancements in the rational design and fabrication of carbon-based electrocatalysts. Moving beyond traditional catalyst optimization methods, innovative engineering strategies have been developed, such as precise modulation of microstructures (e.g., programed hierarchical structure), multi-component synergistic engineering (e.g., topological and doping effect), and the incorporation of atomically precise catalytic centers (e.g., dual-single-atom sites). These modification strategies have led to significantly enhanced H2O2 production activity, selectivity and stability, achieving performance levels that rival or even surpass those of the leading catalysts reported to date. Therefore, it is timely to provide a comprehensive overview of the latest breakthroughs in advanced engineering strategies for carbon-based nanomaterials in electrochemical H2O2 production (Fig. 1). First, we describe the mechanisms of the 2e− ORR and water oxidation reaction (WOR) pathways and elucidate the key factors influencing selectivity and activity. Next, we thoroughly summarize the recent progress in carbon-based electrocatalysts for H2O2 synthesis, emphasizing their underlying structure–performance relationships, particularly in the 2e− ORR pathway. Finally, we highlight the current challenges and propose future directions for the rational design and large-scale utilization of advanced carbon-based electrocatalytic materials.
![]() | ||
Fig. 1 Schematics of key challenges and corresponding modification strategies of carbon-based electrocatalysts for H2O2 production. |
![]() | ||
Fig. 2 (a) Possible reduction pathways of the ORR and WOR for electrocatalysis H2O2 production (left) and a plot of theoretical limiting potential (UL) against Gibbs free energies of binding *OOH (ΔGHOO*) and *OH (ΔGHO*) for H2O2 electrosynthesis (right) of different types of electrocatalysts. UL is the least positive (anodic) or negative (cathodic) potential at which both electron transfers are downhill in free energy.2 Copyright 2019, Nature Publishing Group. (b) Schematic illustration of crucial factors and characterization techniques for investigating the electrocatalytic mechanism. |
O2 + * + H+ + e− → *OOH | (1) |
*OOH + H+ + e− → H2O2 + * | (2) |
Consequently, ORR mechanisms can be divided into two pathways: the 2e− or 4e− pathway. The 4e− pathway is preferred as the cathode reaction in fuel cells and metal–air batteries to obtain higher voltage and energy efficiency, while the electrochemical production of H2O2 relies on the 2e− pathway (eqn (3) and (4)).
In acidic medium:
O2 + 2H+ + 2e− → H2O2 (E0 = 0.67 VRHE) | (3) |
In alkaline medium (pH > 11.6):
O2 + H2O + 2e− → HO2− + OH− (E0 = 0.76 VRHE) | (4) |
A critical aspect in determining the pathway of the ORR is the adsorption mode of O2 molecules on the catalyst's surface. In particular, when O2 is adsorbed in the Pauling mode, the O–O bond is less likely to break, thus favoring H2O2 generation.63
2H2O → H2O2 + 2H+ + 2e− (E0 = 1.76 VRHE) | (5) |
* + H2O → *OH + H+ + e− | (6) |
*OH + *OH → H2O2 | (7) |
The key issue for determining the pathway of the WOR is the recombination of *OH species on the catalyst's surface. Specifically, if *OH is not further oxidized to *O or ˙OH, H2O2 generation is favored.
The electrochemical synthesis of H2O2 using carbon-based electrocatalysts involves complex reaction mechanisms that govern their performance in terms of activity, selectivity, and stability. Density functional theory (DFT) calculations (Fig. 2a) reveal that certain defect configurations in carbon-based materials appear at the pinnacle of the 2e− ORR volcano plot,2 highlighting their role as highly active sites. Based on the fundamental mechanisms, key engineering strategies are categorized into: (1) modulation of materials'surface that governs selectivity and mass transfer, (2) reaction pathways and intermediates, and (3) selection of the electrolyte. This section provides a cohesive framework for understanding the underlying principles of electrocatalysis H2O2 production (Fig. 2b), which is crucial for optimizing the reactant adsorption/activation, stabilization of intermediates, and efficient product release.
Moreover, the wettability of the reaction interface, especially hydrophilicity/hydrophobicity, plays an important role in regulating the adsorption/desorption dynamics of products on the electrode surface.68 For instance, hydrophilic surfaces may facilitate ion transport and reduce gas bubbles overpotential, while controlled hydrophobicity can promote O2 diffusion to active sites, a crucial factor for the ORR. The balance between hydrophilic and hydrophobic domains can be optimized through chemical functionalization or micro/nanostructuring to create ideal triple-phase boundaries for simultaneous O2 supply and H2O2 release.69–71 When combined with other catalyst modifications (defect engineering, heteroatom doping, etc.), wettability control creates synergistic effects that enhance both the kinetics and thermodynamics of the 2e− pathways, bridging fundamental mechanistic understanding with practical catalyst optimization for efficient H2O2 production.
Based on the ORR mechanism, three factors are crucial to ensure high H2O2 production selectivity.72,73 Firstly, a suitable O2 adsorption model is necessary to prevent the dissociative pathway. When the adsorption state is in Pauling mode, the O–O bond is unlikely to break, which is favorable for the generation of H2O2. Secondly, the catalysts should balance the strong adsorption of O2 to facilitate OOH* generation, while allowing moderate desorption to produce H2O2 rather than H2O. Thirdly, it is important to ensure rapid release of the produced H2O2 from the catalyst surface to avoid further reduction or decomposition.
As for the WOR process, the selectivity among the 1e−, 2e−, and 4e− pathways is determined using the descriptors ΔG*OH and ΔG*O, where ΔG*O – 2ΔG*OH = 0.28 eV.74,75 The selectivity of H2O2 production depends on the recombination of two *OH intermediates; If *OH is further oxidized to *O or ˙OH, the H2O2 formation is suppressed. At ΔG*OH = 2.38 eV, *OH can form either H2O2 or *O. If the adsorption energy of *O exceeds the production energy of H2O2 (3.52 eV), the reaction favors the 4e− pathway (with ΔG*O < 3.52 eV and ΔG*OH < 1.62 eV). The optimal condition for the target 2e− WOR occurs when ΔG*OH is between 1.62 and 2.38 eV, as the standard potential is more positive than that of H2O2 oxidation (E0 = 0.67 VRHE).
![]() | ||
Fig. 3 Timeline of engineering strategies for carbon-based electrocatalysts for H2O2 production: N-doped carbon nanotubes,42 reduced graphene oxide,44 hierarchically porous carbon,50 N-doped carbon nanotube cups,78 mesoporous N-doped carbon,79 mesoporous carbon hollow spheres,80 boron-doped carbon,81 single-atom Pd/carbon,82 lattice distortion carbon,59 mesoporous carbon spheres,83 CoZn dual-single-atom carbon nanotubes,84 and porous 2D structures.85 |
The catalytic performance of microstructures is heavily influenced by their dimensions and porosity, as these factors determine the accessibility of active sites and the transport of reactants and products.86,87 For example, 1D nanostructures such as nanotubes provide excellent electron transport pathways, while 2D nanostructures like graphene sheets offer high surface areas for active site exposure. Moreover, when porous architectures are introduced, such as macropores (larger pores that facilitate mass transport), mesopores (intermediate-sized pores that enhance exposure of active sites), and micropores (smaller pores that control the retention of reactants), the material's ability to regulate reaction intermediates and diffusion kinetics can be greatly promoted.88 Moreover, the 2e− ORR can be regulated by controllable pore structures through adjusting the coordination of reaction intermediates. By optimizing the exposure of active sites and retention time of H2O2 within specialized pore sizes, the overall performance of the catalyst can be finely tuned.51,89 Recently, hollow structures that typically feature lower density and larger surface area have garnered significant attention due to their unique properties compared to solid counterparts, by facilitating mass transport and enhancing the availability of active sites. The increased surface-to-volume ratio in hollow structures further contributes to higher utilization of the catalyst material, making them highly attractive for electrocatalytic H2O2 production.80,90–92 This section will mainly focus on the key latest advances in the rational design of high efficient carbon-based electrocatalysts with controllable microstructures.
Precision nanoengineering of porous 2D structures has been emerging as a promising strategy for fine-tuning catalytic reactions. Guided by finite element simulation (FEM), Tian et al.85 designed and fabricated porous 2D carbon nanomaterials by introducing mesopores with diameters of 5–10 nm to facilitate fluid acceleration (Fig. 4a–c). The resulting mesoporous carbon nanosheets exhibited exceptional electrocatalytic H2O2 production performance, achieving a high selectivity of >95% and a diffusion-limiting disk current density of −3.1 mA cm−2 (Fig. 4d–f). FEM simulations revealed that the mesoporous nanosheet significantly accelerated fluid flow within the meso-channels due to viscosity effects and the constricted flow path (Fig. 4g), validating the crucial role of mesoporous 2D structures in enhancing local diffusion. Remarkably, the electrolysis process in a flow cell achieved a high production rate of 14390 mmol g−1 h−1, yielding a medical-grade H2O2 solution. This work demonstrates an effective approach to improving the activity and selectivity of porous carbon materials by influencing local fluid transport behavior.
![]() | ||
Fig. 4 Nanoengineering of porous 2D carbon structures for ORR H2O2 production. TEM images and structural models of (a) MeCN, (b) MaCN, and (c) MiCN; (d) LSV curves recorded in an O2-saturated KOH, (e) calculated H2O2 selectivity, and (f) ORR electron transfer number at various potentials; (g) color mapping of the spatial pressure distribution, and spatial distribution of flow velocities of the mesoporous carbon sphere and nanosheet model (pore size: 5 nm); (h) proposed fluid behavior and ORR processes in micropores, mesopores, and macropores. Reproduced with permission from ref. 85. Copyright 2024, the American Chemical Society. |
Mesoporous hollow nanoreactors (MHNs), a novel class of rationally designed catalytic material, offer unique advantages of complex catalytic processes due to their hollow internal spaces and mesoporous structures.93–95 The mesoporous channels can create a confined microenvironment that enhances molecular diffusion, adsorption and surface reactions, enabling tunable catalytic pathways. For instance, Tian et al.83 demonstrated precisely engineered carbon spheres by integrating micromechanics with controllable synthesis to improve their 2e− ORR catalytic activity (Fig. 5). The mesoporous channels accelerated the fluid flow and facilitated the transport of generated H2O2 into the solution (Fig. 5a–c), thereby minimizing electro-reduction on the catalysts’ surface. Increased flow rates led to O2 enrichment in the pore channels (Fig. 5d), while the accumulation of OH− ions (Fig. 5e) elevated the local pH within the MHNs. And the surface of the MHCS0.5 electrode demonstrated the highest OH− concentrations among all samples (Fig. 5f). As a result, MHCS0.5 exhibited exceptional 2e− ORR performance under neutral electrolyte conditions (pH = 7), with a diffusion-limited disk current of −2.8mA
cm−2 at 0.2
V, an onset potential of 0.6
V (vs. RHE), and an H2O2 selectivity exceeding 85%, outperforming most carbon-based catalysts reported in the literature.
![]() | ||
Fig. 5 Nanoreactor design based on carbon nanospheres. (a)–(c) Schematic diagram of the electrochemical 2e− ORR to produce H2O2 over the designed nanoreactor; Spatial distribution of (d) O2 and (e) OH− concentration in the mesoporous carbon sphere model; (f) OH− concentrations on the surface of MHCSx; (g) LSV curves of RRDE measurements, (h) H2O2 selectivity, and (i) electron transfer number at various applied potentials; (j) structural model and TEM images of MHCSx. Reproduced with permission from ref. 83. Copyright 2024, Nature Publishing Group. |
![]() | ||
Fig. 6 Edge defect engineering of ORR activity. (a) Schematic, SEM, and TEM images of different mesoporous carbon nanofibers (MCNFs); (b) LSV curves and (c) calculated H2O2 selectivity; (d) preferred *OOH adsorption configurations on pristine graphene and different carbon defects models (blue, grey, and red represent C, H, and O atoms, respectively); (e) free-energy profiles of ORR paths at a potential of 0.7 V; (f) volcano plot for the most active structures, with the limiting potential depicted as a function of ΔG*OOH. Reproduced with permission from ref. 98. Copyright 2023, Wiley-VCH. |
By introducing asymmetric structures into a hexagonal carbon lattice, the symmetry can be broken to allow rapid electron transfer, thus promoting the electrocatalytic process. Recently, topological defects have been recognized as effective active sites toward catalyzing different electrochemical reactions. She et al.105 investigated the curvature-dependent ORR activity of surface oxidized carbon nanotubes (o-CNTs) (Fig. 7a–d). Computation modeling suggested that the curvature can alter the epoxy group geometry, exerting greater strain on the C–O bond in smaller diameter o-CNTs that leads to improved activity. Increase of R results in stronger OOH* binding and a lower ΔGOOH* (Fig. 7c), and 2e-ORR theoretical potential (UL) approaches the equilibrium potential (E0 = 0.69 V) on a model with a greater R (Fig. 7d). As predicted, the o-CNT with the smallest diameter (∼8 nm) exhibited the highest faradaic efficiency >85% and a mass activity of 161 A g−1 at 0.65 V. Very recently, Wang et al.106 fabricated carbon nanomaterials with rich topological defect sites and curved defective surface by controlling the pyrolytic shrinkage process of precursors (Fig. 7e–i). Theoretical calculations demonstrated that bending the defect sites can manipulate the local electronic structure, facilitate the charge transfer to key intermediates, and reduce the ORR energy barrier. Experimental results showed that a large kinetic current density of 22.5 mA cm−2 at 0.8 V vs. RHE was obtained for high-curvature defective carbon (HCDC), which is ∼18 times of low-curvature defective carbon (LCDC). Further increasing the defect densities of HCDC results in a dual-regulated product (HCHDC), which exhibited exceptional ORR activity in both alkaline and acidic media (half-wave potentials of 0.88 and 0.74 V, respectively), exceeding most of the reported carbon electrocatalysts. These studies highlight the crucial role of curvature effect in promoting electrocatalytic activity and offer new guidance to the design of advanced carbon nano-catalysts for H2O2 production.
![]() | ||
Fig. 7 Investigation of the curvature-dependent ORR activity. (a) Schematic illustration of the epoxy group grafted on the curvature varied carbon surface and (b) geometry optimized atomic models of different R values; (c) calculated ORR free energy diagrams at 0.69 V and (d) correlations between ΔGOOH* and UL to surface curvature. Reproduced with permission from ref. 105. Copyright 2024, the American Chemical Society. (e) TEM, HRTRM, and AC HAADF-STEM images of LCDC and HCDC (insets show the morphology and atomic structure diagrams); (f) LSV curves in O2-saturated KOH and (g) comparison of Jk (0.8 V) and E1/2; (h) LSV curves of LCDC, HCDC, and Pt/C in O2-saturated H2SO4 and (i) comparison of Jk (0.6 V) and E1/2. Reproduced with permission from ref. 106. Copyright 2024, Wiley-VCH. |
In 2021, Xia et al.81 reported a boron-doped carbon (B–C) catalyst (Fig. 8). Compared to the state-of-the-art oxidized carbon catalyst, the B–C catalyst presented significantly lowered overpotential by 210 mV under industrial-level currents (300 mA cm−2) while maintaining high H2O2 selectivity (85–90%) (Fig. 8e and f). DFT calculations revealed that the boron dopant site is responsible for high H2O2 activity and selectivity due to the reduced thermodynamic and kinetic reaction barriers (Fig. 8c and d). Integrated in a porous solid electrolyte reactor, the B–C catalyst demonstrated continuous generation of pure H2O2 solutions with high current density (∼400 mA cm−2) and selectivity (∼95%), presenting their great potential for practical applications in the future.
![]() | ||
Fig. 8 Heteroatom-doped carbon catalysts for electrocatalytic H2O2 generation. (a) XPS survey scan and (b) Raman spectroscopy for pure C and B, N, P, S-doped carbon catalysts; (c) preferred *OOH adsorption configurations on B-, P-, N-, and S-doped graphene, respectively. Green, orange, blue, yellow, gray, red, and white spheres represent B, P, N, S, C, O and H, respectively; (d) free-energy profile of O2 reduction paths (URHE![]() ![]() ![]() ![]() |
Oxygen-doped carbon quantum dots (o-CQDs) with C–O–C surface functional groups were fabricated with tunable electronic structures by varying isomerization precursors,112 presenting a remarkable H2O2 selectivity of 96.2% (n = 2.07) at 0.68 V vs. RHE along with a low Tafel slope of 66.95 mV dec−1, and maintaining consistent production stability of H2O2 over 120 h. In addition, a hierarchical porous boron-doped carbon (B-DC) electrocatalyst was synthesized from fullerene (C60) frameworks and boric oxide.107 Befitting from boron doping and abundant topological pentagon defects, the B-DC catalysts exhibited a high ORR onset potential of 0.78 V and a 2e− selectivity over 95%. Remarkably, the B-DC electrocatalyst-based device achieved a remarkable H2O2 yield rate of 247 mg L−1 h−1 and a quantitative Faraday efficiency of nearly 100%. For sulfur-doped defective nanocarbons (S-DNC), a similar high ORR onset potential (0.78 V) and selectivity (90%) was obtained, with an superior H2O2 yield rate of 690 mg L−1 h−1 in a H cell.113 In 2023, Lu et al.102 designed and fabricated a pentagonal defect-rich N-doped carbon nanomaterial (PD/N–C) via pyrolysis of C60 as the precursor followed by ammonia treatment (Fig. 9a). A great number of mesopores and micropores were created in PD/N–C, along with abundant irregular folds and curved edges (Fig. 9b–d). Moreover, the aberration-corrected scanning TEM (AC-STEM) and the corresponding fitting results (Fig. 9e–j) further demonstrate that the PD/N–C sample contains rich pentagon defects. As a result, the PD/N–C catalysts achieved excellent 2e− ORR activity, selectivity, and stability in acidic electrolytes, even outperforming the Pt–Hg alloy catalyst. The linear sweep voltammetry curves (Fig. 9k) show that the PD/N–C catalyst achieves a larger ORR current density (3.1 mA cm−2 at 0 V) along with a higher onset potential in a wide potential range from 0 to 0.6 V, as compared to the N-doped carbon nanomaterial dominated by hexagon graphene nanosheets (N–GNS). This performance is among the top of all reported metal-free carbon-based catalysts and even superior to some noble-metal-based catalysts and the benchmark PtHg4 alloy catalyst in acidic electrolytes (Fig. 9l). By comparing a series of samples, the synergetic effects of both topological defects and N-doping are responsible for the superb 2e− ORR performance of the PD/N–C catalysts (Fig. 9m and n).
![]() | ||
Fig. 9 Pentagonal defect-rich N-doped carbon nanomaterial (PD/N–C) using fullerene (C60). (a) Schematic of the synthesis of PD/N–C; (b) SEM, (c) and (d) TEM, and (e), (g) and (i) AC-STEM images and (f), (h) and (j) the corresponding Fourier transform fitting results of PD/N–C; (k) RRDE polarization curves and H2O2 currents at the ring electrode of PD/N–C and N–GNS; (l) comparison of the 2e− ORR performance of representative electrocatalysts; (m) RRDE polarization curves and H2O2 currents at the ring electrode of N–CNT, N–CNS, and PD/N–C-2; (n) onset potential (E0) as a function of the D3/G ratio of a series of samples. Reproduced with permission from ref. 102. Copyright 2023, the American Chemical Society. |
In addition to single-element doping, dual element co-doping on carbon materials has received increased research attention toward the electrocatalytic 2e− ORR due to the unique synergistic effect.114 For instance, a yolk–shell B/N co-doped hollow carbon nanosphere, with oxygen-vacancy decorated reduced graphene oxide coating (B/N-HCNS@VO-G) was reported as an efficient metal-free electrocatalyst for the 2e− ORR with high durability.115 Such a dual-doping electrocatalyst leads to excellent electrocatalytic H2O2 production performance with a high selectivity of 91% and yield of 56 ppm (0.7 V), allowing in situ antibiotic and dye degradation for on-site wastewater remediation. Very recently, pentagon-S and pyrrolic-N coordinated (SNC) graphene with in-plane topological defects was synthesized through a two-step hydrothermal and nitridation procedure (Fig. 10).99 The SNC is composed of stained hexagons with sporadic pentagons (Fig. 10d–g), suggesting that the hexagonal topological structure of the carbon matrix is significantly distorted. In addition, the dual-doping of S and N elements introduces unsymmetrical dumbbell-like S–C–N motifs, which can effectively tune the electronic structures of graphene (Fig. 10h and i). Theoretical calculations further unveiled that the defective S–C–N motifs can effectively optimize the binding strength to the OOH* intermediate and reduce the energy barrier for the ORR to H2O2 (Fig. 10j). As a result, the SNC catalyst exhibited ultrahigh H2O2 production rates of 8100, 7300, and 3900 mmol g−1 h−1 in alkaline, neutral, and acidic electrolytes, respectively.
![]() | ||
Fig. 10 Pentagon-S and pyrrolic-N coordinated (SNC) graphene with in-plane topological defects. (a) Schematic illustration of synthesis for graphene pentagon-S and pyrrolic-N coordinated (SNC); (b)–(g) XRD patterns, Raman spectra, TEM images, and EDS mappings; (h) computational models, (i) Bader charge, (j) activity-volcano plot (UL = 0.70 V), and (k) ORR reaction pathway diagrams and corresponding energy barriers for each step on Basal 2. Reproduced with permission from ref. 99. Copyright 2024, Wiley-VCH. |
It is worth mentioning that heteroatom doping into the carbon matrix could also induce surface functional groups. For example, oxygen functionalization onto the parent carbon structures has been reported to be effective to improve 2e− ORR performance.21,116,117 For instance, oxygen-doped carbon dots (O-CDs) were synthesized using a solvent engineering approach and exhibited excellent catalytic activity.118 By tuning the ratio of ethanol and acetone solvents during the synthesis, the surface electronic structure of the resulting O-CDs can be precisely modulated. The selectivity and activity of the O-CDs were strongly dependent on the amount of edge active C–O groups, achieving the highest H2O2 selectivity of up to ∼96.6% (n = 2.06, 0.65 V vs. RHE) and an ultralow Tafel slope of 64.8 mV dec−1.
Despite tremendous works on doped carbon materials for electrocatalysis H2O2 production, the dynamic structural transformation of these materials during the electrocatalysis reaction process has received less attention. Active site identification in carbon materials is crucial for understanding the mechanism, but resolving precise configuration of active sites remains a huge challenge. In 2023, Wu et al.119 manipulated the defect density and oxygen groups on graphene, and revealed the oxygen group redistribution and positive correlation relationship between the defect density and their ORR performance (Fig. 11). The dynamic evolution processes of defects were monitored through in situ Raman, FTIR and XPS technologies, combined with theoretical simulations. The results clarified the configuration of major active sites (carbonyl on pentagon defect) and key intermediates (*OOH), providing deep understanding of the catalytic mechanism for doped defective carbon materials.
![]() | ||
Fig. 11 In situ investigation of dynamic active sites in oxygen modified defective graphene (O-DG) electrocatalysts. iDPC-STEM image of O-DG-30 in (a) perfect domain and (b) and (c) defective domain; in situ Raman spectra of (d) DG-30 and (e) O-DG-30 on a flow cell in O2-saturated 0.1 M KOH; (f) the calculated Raman spectra of three possible atomic structures of the O-groups and relevant surface species on defective graphene; (g) schematic diagram of the possible electrocatalytic mechanism of O-DG. Reproduced with permission from ref. 119. Copyright 2023, Nature Publishing Group. |
Recently, Xue et al.121 reported a novel p–n heterojunction nanocomposite consisting of N-doped carbon and Co3O4 (NC@Co3O4), for efficient electrocatalytic H2O2 production (Fig. 12). Increasing the Co content results in a more positive flat band potential (Efb), indicating a decrease of the Fermi level. In addition, the differential charge density calculation demonstrated that the electrons of the NC are transferred to the Co3O4 and promotes the affinity of the O atom of H2O2 at the electron deficient carbon sites in NC, which facilitates the cleavage of O–O bonds. Consequently, the ˙OH generation rate catalyzed by NC@Co3O4 was 6.5 times of that by bare NC. This work highlights the promising potential of constructing carbon/semiconductor nanocomposites toward efficient H2O2 generation.
![]() | ||
Fig. 12 N-doped carbon and Co3O4 (NC@Co3O4) p–n heterojunction nanocomposite for electrocatalytic H2O2 production. XPS spectra of (a) N 1s and (b) C 1s; (c) FTIR spectra of NC and NC@Co3O4; (d) the flat band potential (Efb) estimated from Mott–Schottky plots; (e) Fermi level measured by UPS; (f) differential charge density of the NC@Co3O4 heterojunction; (g) schematic illustration of the energy band and electron transfer of the NC@Co3O4 p–n heterojunction. Reproduced with permission from ref. 121. Copyright 2024, Elsevier. |
Gao et al.147 fabricated transition metal (Mn, Fe, Co, Ni, and Cu) SAC anchored N-doped graphene and systematically investigated their electrocatalytic performance for synthesizing H2O2 via the ORR (Fig. 13a). Theoretical simulations demonstrated that the Co-SAC possesses optimal d-band center, activity-volcano plot, and adsorption energy of the *OOH intermediate among all M-SAC samples (Fig. 13b–d). As a result, the kinetic current for H2O2 production using the Co SAC catalyst can reach up to ∼1 mA cm−2 (0.6 V vs. RHE in 0.1 M HClO4) with Faraday efficiency over 90% (Fig. 13e and f), which even outperforms state-of-the-art noble-metal-based electrocatalysts in acidic media. Moreover, kinetic and in situ X-ray absorption analysis demonstrated that the N-coordinated single Co single-atom function as the active site for the reaction, which is rate-limited by the first electron transfer step.
![]() | ||
Fig. 13 Transition metal SACs (M-SAC) (M = Mn, Fe, Co, Ni, and Cu) anchored in N-doped graphene electrocatalysts for producing H2O2. (a) Schematic of the ORR; (b) binding energy of *OOH, *O, and *OH on M-SAC and d-bond center of different M atoms in M-SAC; (c) volcano curves of the ORR via the 2e− and 4e− pathways; (d) free energy diagrams of the 2e− ORR on different M-SACs (0.7 V versus RHE); (e) LSV curves and (f) faradaic efficiency as a function of potential. Reproduced with permission from ref. 147. Copyright 2020, Cell Press. |
Compared to well-studied SACs, dual-atom catalysts (DACs) have received considerable attention due to the synergistic interactions between two adjacent metal sites, leading to remarkably enhanced activity, selectivity, and stability. The dual sites facilitate multi-step reactions and improve binding energies for reactants and products. In addition, DACs are more resistant to aggregation and can offer tunable properties through metal pair selection, making them ideal for complex catalytic processes. Recent advancements in DAC design have demonstrated their superior electrocatalytic performance for H2O2 production through the 2e− ORR pathway.
In-N-C DACs were proposed by Du et al.148 as effective 2e− ORR catalysts for producing H2O2 in acidic media. The DFT calculations indicate that the valence electron number and the d-band center of the Co 3d orbital can be modulated by OH-blocked In, which leads to moderate adsorption of the OOH intermediate on neighboring Co and favors 2e− ORR kinetics of Co/In-N-C DACs. As a result, a partial current density of 1.92 mA cm−2 (0.65V in the RRDE test) was obtained, with a H2O2 production rate as high as 9680 mmol g−1 h−1 in a three-phase flow cell. Very recently, Yang et al.84 fabricated heteronuclear CoZn DACs confined in N,O-doped hollow carbon nanotube reactors (CoZnSA@CNTs) (Fig. 14). The differential partial charge density calculation demonstrated that the dual-atom centers in CoZnSA@CNTs possess more charges than CoSA@CNTs (Fig. 14a), indicating a rearrangement of the local charge distribution to facilitate the adsorption of OOH* after introducing Zn. The decrease of the d-band center in CoZnSA@CNTs resulted in the appropriate adsorption of OOH* for H2O2 generation (Fig. 14b), while inhibiting the breakage of O–O bond to produce H2O, ultimately facilitating the 2e− ORR. In addition, the CoZnSA@CNT model exhibited a more favorable ΔGOOH* than CoSA@CNTs (Fig. 14c and d). The well-designed CoZnSA@CNT nanocomposite displayed outstanding electrocatalytic reactivity/selectivity for generating H2O2 in the whole pH range (Fig. 14e–g), with a higher 2e− ORR selectivity for H2O2 production than CoSA@CNTs and ZnSA@CNTs. In a H-type cell, CoZnSA@CNTs/carbon fiber felt reached nearly 100% H2O2 selectivity in the range of 0.2–0.65 V (vs. RHE) with a yield rate of 1500 mmol g−1 h−1, surpassing most of the reported SACs. These studies highlight the great advantages of using DACs/carbon for highly efficient and selective H2O2 production.
![]() | ||
Fig. 14 CoZn DACs confined in N,O-doped hollow carbon nanotube reactors (CoZnSA@CNTs) for electrocatalytic synthesis of H2O2. (a) Illustration of the model and their differential partial charge densities with OOH* adsorption (side view); (b) PDOS of the Co 3d orbital and (c) Gibbs free energy diagrams of the two-electron ORR of CoSA@CNTs and CoZnSA@CNTs, respectively; (d) schematic of the optimized electronic structure; electrochemical performance of CoZnSA@CNTs and CoSA@CNTs: (e) CV curves, (f) LSV curves based on RRDE, and (g) H2O2 selectivity and electron transfer number (n). Reproduced with permission from ref. 84. Copyright 2024, Wiley-VCH. |
In addition, Jing et al.89 demonstrated that N,O co-doped carbon nanosheets (N,O-CNS) with a hierarchical micro/mesoporous structure can create an oxygen-rich, locally alkaline-like microenvironment, significantly promoting the 2e− ORR pathway in a neutral medium. Their study revealed that the hierarchical architecture not only elevates the local pH near the active sites but also facilitates the formation of intermediates (*O2 and *OOH), thereby enhancing H2O2 selectivity and yield. Remarkably, the optimized N,O-CNS0.5 catalyst achieved an exceptional H2O2 production rate of 6705 mmol g−1 h−1 in a flow cell, setting a new benchmark for neutral-media electrocatalysis. This work highlights the critical role of synergistic microenvironment engineering, which combines pore structure control, heteroatom doping, and local pH modulation, in designing high-performance electrocatalysts for sustainable H2O2 synthesis.
![]() | ||
Fig. 15 DFT calculation on various metal SAC catalysts anchored on nitrogen-doped carbon (M–NC) for H2O2 production via 2e− ORR and 2e− WOR. (a) Optimized MN4 structure with different pyrrole N coordination numbers; activity volcano relation for (b) 2e− ORR and (c) 2e− WOR between intermediates' formation energy and limiting potential; (d) scaling plots between ΔG*OH and ΔG*OOH; (e) relationship between pyrrole N coordination numbers and limiting potential for 2e− ORR and 2e− WOR of all Ru-NC catalysts. Reproduced with permission from ref. 151. Copyright 2023, Elsevier. |
Moreover, as a notable example, Dong et al.140 developed a novel Mn SAC-coordinated boundary-rich porous carbon-based electrocatalyst (Fig. 16a and b), in which the secondary coordinated epoxide, hydroxyl groups surrounding the Mn–N3–O centers, and the boundary-rich morphology together lead to the predominant selectivity and efficiency for H2O2 production through the 2e− ORR pathway. The catalysts exhibited nearly 100% faradaic efficiency and with a reaction rate up to 15100 mmol gcat−1 h−1 (0.1 V vs. RHE), achieving the record activity for the Mn-based electrocatalyst in the electrosynthesis of H2O2 (Fig. 16c–e). Mechanistic investigations indicate that the epoxide and hydroxyl groups surrounding the Mn(II) centers promote the spin state and tailor the adsorption of the *OOH intermediate, and multiscale simulations reveal that the high-curvature boundaries can promote adsorption and local enrichment of O2 (Fig. 16f–j). Such synergy highlights the importance of holistic material design, where defects, doping, and atomic-scale engineering work in concert to achieve greatly improved performance.
![]() | ||
Fig. 16 (a) The schematic of the heterogeneous catalyst composed of boundary-rich carbons supported active Mn(II) centers (Mn–NO–CH) with selective oxygen functional groups. (b) SEM, TEM and HAADF-STEM images of Mn–NO–CH. (c) TOFs, (d) H2O2 selectivity, and (e) correlation between H2O2 selectivity (0.62 V vs. RHE) and the proportion of oxygen species. (f) The DFT models for Mn–N3O1. (g) The calculated free energy at 0 V vs. RHE for the 2e-ORR pathway. (h) The correlation between the magnetic moment and the free energy for Mn–N3O1 after *OOH adsorption. (i) MD simulation for the minimum distance between O2 molecules and C atoms. (j) Electronic localization function mapping images with different curvatures. Reproduced with permission from ref. 140. Copyright 2024, Wiley-VCH. |
To comprehensively evaluate the advantages and limitations of carbon-based electrocatalysts for H2O2 production, it is essential to compare them with non-carbon-based catalyst systems, such as metal oxides and noble-metal catalysts. Table 1 summarizes the critical metrics (onset potential, H2O2 selectivity, production rate, and stability) across different representative catalyst types. Although Pt/Pd-based noble metal-based catalysts are recognized as efficient catalysts with small overpotential and high selectivity for the 2e− ORR, their industrial-scale application is greatly limited by scarcity, high cost, and susceptibility to poisoning.24,41,153,154 Metal oxide electrocatalysts demonstrate excellent stability, while achieving high selectivity remain challenging.155 Although some recently reported metal oxides such as h-SnO2 demonstrate excellent electrocatalytic activity,156 they still suffer from poor stability in electrolyte. Furthermore, synthesizing highly active metal oxide catalysts often requires complex procedures and precise control over their structure and composition. It is also worth noting that the emerging MOF/COF materials show high production activity but poor stability.
Type | Catalysts | Selectivity (%) | H2O2 production rate | Electrolyte | Stability | Ref. |
---|---|---|---|---|---|---|
Noble metals | PtP2 | 98.5 | 2.26 mmol h−1 cm−2 | 0.1 M HClO4 | 120 h | 157 |
Pt–Hg/C | >90 | — | 0.1 M HClO4 | 8000 cycles | 24 | |
Pt1–CuSx | 92–96 | 546 mmol g−1 h−1 | 0.1 M HClO4 | 10 h | 158 | |
Pt1/CoSe2 | — | 110.02 mmol g−1 h−1 | 0.1 M HClO4 | 60 h | 159 | |
Oxides | h-SnO2 | 99.9 | 3885.26 mmol g−1 h−1 | 0.1 M Na2SO4 | 20 h | 156 |
α-Fe2O3 | 80.5 | 546.8 mmol g−1 h−1 | 0.1 M KOH | 48 h | 155 | |
Carbon-based | Single atoms-carbon | 96.5 | 12![]() |
0.1 M KOH | 12 h | 139 |
N/S-CNTs | 90.0 | 30![]() |
1.0 M KOH | 200 h | 160 | |
B-doped carbon | 88.7 | 24![]() |
0.1 M KOH | 100 h | 161 | |
Other systems | MOF nanosheets | 99 | 6500 mmol | 0.1 M KOH | 11 h | 162 |
Boron nanosheets | ∼90 | 25![]() |
1.0![]() |
140 h | 163 |
We have summarized the most representative carbon-based materials for electrocatalytic H2O2 production reported in the past few years (Table 2), systematically classified by materials design categories: microstructure tailoring, defect engineering (including intrinsic and heteroatom doping), surface functional group modulation, single-atomic site configurations, and synergistic effects. In comparison, carbon-based electrocatalysts offer distinct advantages, including low-cost and wide availability, making them highly attractive for large-scale applications. Up till now, the performance of carbon-based electrocatalyst is comparable or even outperforming those of many noble metal-based systems, maintaining cost and environmental benefits. Carbon-supported single-atom catalysts (SACs) achieve high production rates of ∼500–30000 mmol g−1 h−1. However, the activity and stability of carbon catalysts in neutral or weakly acidic environments still need to be further improved, as well as mitigating degradation during long-term electrochemical operation.
Type | Catalysts | Electrolyte | Onset potential (V vs. RHE) | H2O2 selectivity (%) | Production rate | Voltage (V vs. RHE) | Faradaic efficiency (%) | Tafel slope (mV dec−1) | Stability | Ref. |
---|---|---|---|---|---|---|---|---|---|---|
Microstructure modulation | Carbon mesoporous nanoreactors (MHCS) | 0.1 M KOH | 0.6 | 85 | 3.36 mmol L−1 h−1 | 0.5 | 97 | 73 | 12 h | 83 |
Graphitic ordered mesoporous carbon (O-GOMC) | 0.1 M KOH | 0.75 | ∼92 | 63.33 mg L−1 h−1 | 0.6 | 99.2 | 42 | 168 h | 164 | |
O-GOMC | 0.1 M KOH | 0.73 | 93 ± 1 | 24 mmol L−1 | — | 99 | 59 | 16 h (3 mA) | 165 | |
Honeycomb carbon nanofibers (HCNFs) | 0.1 M KOH | 0.87 | 97.3 | 6.37 mmol L−1 h−1 (0.05 mg) | 0.5 | — | 75.6 | 12 h | 166 | |
Porous carbon (PCC900) | 0.1 M KOH | 0.83 | 95 | 1696.8 mmol g−1 h−1 | — | 90 | 38 | 10![]() |
167 | |
Hollow mesoporous carbon spheres (HMCSs) | 0.1 M KOH | 0.82 | 95 | — | 0.53 | 87 | — | 10 h | 168 | |
Mesoporous carbon-nanofibers (MCNF) | 0.1 M KOH | 0.68 | >90 | — | 0.4 | >85 | 43 | 12 h | 98 | |
Holey graphene | 0.1 M KOH | 0.56 | 95 | 2360 mmol g−1 h−1 | 0.1 | 97 | — | 12 h | 169 | |
Defect engineering | Oxidized carbon nanotubes | 0.1 M KOH | 0.795 | >90 | — | 4 | >85 | 38 | 96 h | 170 |
Oxygen-doped carbon quantum dots | 0.1 M KOH | — | 96.2 | 338.7 mmol g−1 h−1 | −1.33 | 92 | 66.95 | 120 h | 171 | |
Oxygen modified defective graphene | 0.1 M KOH | 0.9 | 98.38 | 41.45 mg cm−2 h−1 | 0.5 | 95 | — | 10 | 172 | |
Doped carbon nanohorns | 0.1 M KOH | 0.85 | >80 | 740 mmol g−1 h−1 | 0.65 | 50–100 | 49 | 12 h | 173 | |
B-doped carbon | 0.1 M KOH | 0.780 | 88.7 | 24![]() |
— | 82 | 51 | 100 h (100![]() |
161 | |
B-doped nanocarbon | 0.1 M KOH | 0.78 | 95 | 247 mg L−1 h−1 | 0.5 | 100 | — | 10 h | 174 | |
B-doped carbon | 0.1 M KOH | 0.773 | 95 | 7.36 mmol cm−2 h−1 | — | 90 | 78 | 30 h (200![]() |
81 | |
N-doped carbon spheres | 0.1 M KOH | 0.7 | 91.9 | 618.5 mmol g−1 h−1 | 0.4 | 85.1 | — | 10 h | 90 | |
N-doped graphene | 0.1 M KOH | 0.764 | >82 | 224.8 mmol g−1 h−1 | 0.3 | >43.6 | — | 4 h | 175 | |
N-doped carbon | 0.1 M HClO4 | 0.6 | 80–98 | 2923 mg L−1 h−1 | 0.3 | 100 | 136.6 | 10 h | 176 | |
N-doped carbon | 0.5 M NaCl | 0.61 | 95 | 631.2 mmol g−1 h−1 | 0.51 | 79.8 | 79 | 10 h | 177 | |
N,O co-doped carbon nanosheets | 0.1 M K2SO4 | 0.65 | >90 | 6705 mmol g−1 h−1 | 0.2 | 91 | 52 | 24 h | 89 | |
Surface modification | Polydopamine modified carbon (CB-PDA-A) | 0.1 M KOH | 0.8 | 80 | 1800 mmol g−1 h−1 | 1.5 | 95 | 70 | 250 h | 60 |
CoPc-carbon | 0.1 M KOH | 0.7 V | 99% | 10![]() |
0.2 | 93 | 54 | 24 h | 178 | |
Polymerization of acrylonitrile (PNAC-F) | 0.1 M KOH | 0.78 | 93 | 816 mmol g−1 h−1 | 0.78 | — | — | 40 h | 179 | |
Carbon black | 0.1 M KOH | 0.75 | 96 | — | 0.33 | — | 60 | 10 h | 180 | |
Activated coke@carbon cloth | 0.1 M Na2SO4 | — | 100 | 30.41 mg h−1 cm−2 | — | 80 | 104.1 | 10 h | 181 | |
N-B-OH-graphene quantum dots | 0.1 M KOH | 0.7 | 90 | 709 mmol g−1 h−1 | 0.2 | 81 | — | 12 h | 108 | |
Single-atom—carbon hybrids | Pb-carbon dot (Pb SAs/OSC) | 0.10 M KOH | 0.65 | >90 | 6.9 mmol cm−1 h−1 | — | 92.7 | 49 | 100 h (50 mA cm−2) | 144 |
Single atoms cellulose-carbon (FeSAs/ACs-BCC) | 0.1 M KOH | 0.78 | 96.5 | 12![]() |
0.4 | 89.4 | 73 | 12 h | 139 | |
In single atoms on carbon (In SAs/NSBC) | 0.1 M KOH | 0.66 | 95 | 6490 mmol g−1 h−1 | 0.3 | 75 | 30.3 | 12 h | 146 | |
Pt single atoms-g-C3N4 (Pt0.21/CN) | 0.1 M KOH | 0.81 | 98 | 767 mmol g−1 h−1 | 0.6 | 96 | — | 10 h | 142 | |
Metal–nitrogen–carbon | 0.1 M HClO4 | 0.75 | >70 | 688 mmol g−1 h−1 | 0.3 | 93 | 113 | 12 h | 182 | |
Synergistic effects | CoZn–N/O-carbon nanotube (CoZnSA@CNTs) | 0.1 M KOH | — | 90–100 | 4770 mmol g−1 h−1 | 0.2 | 95 | 76.3 | 5 h | 84 |
Co–O4 porous graphene-like carbon (Co–O4@PC) | 0.1 M KOH | 0.73 | 98.8 | 250 mmol g−1 h−1 | 0.5 | — | 72 | 10 h | 137 | |
Co–N4, C–O–C epoxide groups | 0.1 M KOH | 0.801 | 91.3 | 6912 mmol g−1 h−1 | 0.09 | — | 51.3 | 10 h | 135 | |
Atomic metal–nitrogen–carbon (F/Co–N–G SAC) | 0.1 M KOH | 0.810 | 94.2 | 12![]() |
0.1 | 70 | 45.2 | 10 h | 183 | |
S,N-coordinated Ni SAC (Ni-N3S) | 0.1 M KOH | 0.80 | 90.0 | 17![]() |
0.2 | 85 | — | 12 h | 184 | |
Bipyridyl N-carbonyl (Pd-N2O2-C) | 0.1 M KOH | 0.82 | 95 | 1.5 × 107 mmol g−1 h−1 | 0.4 | — | — | 10 h | 185 | |
Pd atoms on nitrogen-doped carbon (Pd-NC) | 0.1 M KOH | 0.8 | 95 | 30 mmol g−1 h−1 | 0.5 | — | 56 | 8 h | 82 | |
Carbon-based single atom (Zn–N2O2–S SAC) | 0.1 M KOH | — | 96 | 6924 mmol g−1 h−1 | — | 93.1 | 62 | 65 h (80 mA cm−2) | 131 | |
Co SACs-carbon nanofiber (Co@EO-ACNF) | 0.1 M KOH | — | 85.21 | 15![]() |
0.6 | 84.5 | — | 48 h | 136 | |
W SAC on O,N-doped carbon (W1/NO-C) | 0.1 M KOH | 0.815 | >90 | 1230 mmol g−1 h−1 | 0.6 | 95 | 115 | 20 h | 141 | |
Pd-oxidized carbon nanotubes | 0.1 M HClO4 | 0.7 | 95 | 1700 mmol g−1 h−1 | 0.1 | 93 | — | 8 h | 73 | |
Defective carbon (Mo-CDC-ns) | 0.1 M KOH | 0.75 | 90 | 455 mmol g−1 h−1 | 0.55 | 85.1 | 95–99 | 12 h | 100 | |
Co@carbon nanocages | 0.1 M HClO4 | 0.60 | 94 | 57 mmol g−1 h−1 | 0.31 | — | 107.9 | 10 h | 152 | |
Ni SAC-carboxyl-multiwall carbon nanotubes (N4Ni1O2/OCNTs) | 1 M KOH | 0.68 | — | 5.7 mmol cm−2 h−1 | — | 96.1 | 107 | 24 h (200 mA cm−2) | 132 | |
Oxygen-vacancy-graphene armor (B/N-HCNS@V-G) | 1 M KOH | 0.79 | 91 | 55 ppm | 0.5 | 94 | 63 | 24 h | 115 | |
Carbon nanotubes with N,S-rich tips (N, S-TCNTs) | 1 M KOH | 0.78 | 90.0 | 30![]() |
— | >90 | 55.5 | 200 h (100 mA) | 160 |
The stability of carbon-based electrocatalysts for H2O2 production varies significantly across different categories, with durations ranging from a few hours to ∼250 h. Microstructure-modulated catalysts like graphitic ordered mesoporous carbon (O-GOMC) exhibit high stability (168 h) due to robust porous frameworks, while defect-engineered materials such as B-doped carbon demonstrate durability (100 h) under high current densities. Surface modifications, including polymer coatings or metal complexes, demonstrated relatively low stability as evidenced by CoPc-carbon (24 h). Although single-atom catalysts generally exhibit very short lifetime (∼10 h), recently reported synergistic systems such as Pb SAs/OSC (100 h) and N,S-rich carbon nanotubes (200 h) are highly promising. Challenges still remain in standardizing testing protocols and achieving long-term stability for industrial applications. Future research should focus on elucidating degradation mechanisms and optimizing catalyst designs for extended operational lifetimes.
Despite the considerable progress achieved, challenges remain for the industrial-scale implementation of electrocatalytic H2O2 production using carbon-based materials, including improving electrocatalytic activity, ensuring long-term stability, and developing practical designs for carbon-based electrodes and reactors. Therefore, advancing carbon electrocatalyst design to achieve high activity and selectivity has become a primary focus in this field. Furthermore, innovation in catalyst preparation, optimization of the activity and selectivity, as well as deeper understanding of mechanisms are crucial to further advancing H2O2 production in an environmentally benign manner more effectively. To date, carbon-based electrocatalysts have demonstrated remarkable progress, driven by advances in engineering strategies that have significantly improved their performance over traditional methods. In the following section, we will summarize and discuss the challenges and future perspectives in this research area (Fig. 17), including the rational design of carbon catalysts (e.g., atomic level engineering, multiple-synergistic effects), advanced synthesis methods, in situ characterization techniques, and system-level device design considerations for industrial applications.
The catalytic activity of carbon-based materials still has significant potential for further improvement. Multiple synergistic modification strategies can be employed, including defect engineering (e.g., creating vacancies and edge sites to promote O2 adsorption), chemical doping (e.g., introducing N, B, or S heteroatoms to tune their electronic structure), facet control (e.g., exposing specific crystal planes with optimal *OOH intermediate binding energy), surface active site design (e.g., atomically dispersed metal–N4 moieties), interface engineering (e.g., constructing hybrid metal oxide/carbon junctions to facilitate charge transfer), and microstructural modulation (e.g., developing hierarchically porous architectures to improve mass transport). Although existing strategies have made notable progress in the electrosynthesis of H2O2, the precise control of defects and single-atomic sties in carbon materials requires further investigation to optimize the balance between the intrinsic activity and selectivity of active sites. To precisely engineer defects in carbon-based materials, several methodologies can be utilized: (1) plasma irradiation (Ar, N2, or O2) for controlled introduction of vacancies and edge defects; (2) mechanical exfoliation techniques such as ball-milling or ultrasonication to create strain-induced defects; and (3) pyrolysis of metal–organic frameworks (MOFs) to derive defect-rich carbon matrices with atomic-level precision. Moreover, to promote the mass transfer process, tailoring the morphology of the catalysts with well-designed 3D hierarchically porous architectures can significantly increase the availability of active sites and accelerate the diffusion of reactants. The pore sizes and distribution need to be carefully designed to balance the retention time and selective H2O2 production, reactant diffusion, resistance and reaction kinetics. Specific methods include the use of sacrificial templates (e.g., mesoporous silica, PMMA spheres, and amphiphilic block copolymers), combination with resins (e.g., resorcinol–formaldehyde) for ordered mesopores, or addition of macroporogen (e.g., emulsion droplets) for triple porosity. In addition, it is crucial to introduce surface functional groups (e.g., –COOH or –OH) to tune hydrophilicity for improving electrolyte accessibility and modifying intermediate binding strength through dipole interactions. For single-atom carbon catalysts, the uncontrollable selectivity and poor stability remain critical challenges. The implementation of atomic layer deposition (ALD) for depositing protective layer (e.g., Al2O3 or TiO2 sub-nanolayers) can be used to enhance the stability while allowing reactant access; creating proximal dual-atom sites is effective to promote the selectivity by modulating the O–O bond cleavage. Critically, the synergistic integration of these strategies represents a pivotal research direction for achieving sustained high selectivity in the 2e− oxygen reduction pathway. Beyond catalyst design, the role of the electrolyte in H2O2 production also deserves attention, as it can influence the stability of the key intermediate *OOH, thereby regulating the overall efficiency of H2O2 formation. And the ORR activity through the 2e− pathway remains limited for many carbon-based materials in neutral or weakly acidic environments.
Beyond materials design, elucidating the fundamental electrocatalytic mechanisms remains a critical research priority. At this moment, the reaction mechanisms of carbon-based electrocatalysts remain insufficiently understood. It is worth noting that the active sites in carbon catalysts typically undergo dynamic evolution during electrochemical reactions, necessitating the study of these changes to fully explore the reaction mechanisms. To this end, a comprehensive understanding of reaction pathways, active site dynamics under operational conditions, and interfacial phenomena is essential for rationally guiding catalyst optimization. Therefore, various in situ/operando characterization techniques are needed to directly probe reactive intermediates, active sites, and their interactions during the electrochemical H2O2 generation process. For instance, in situ XRD and X-ray absorption spectroscopy (XAS) can be employed to analyze the local coordination environment, electronic structure, and oxidation state of catalysts in real time. In addition, in situ Raman and FTIR spectroscopy are valuable for real-time detection of reaction intermediates during electrocatalysis reactions. Further improvement in spatial and temporal resolution for in situ detection of dynamic electrocatalytic processes is also desirable. To gain deeper insights into electrocatalytic mechanisms, advanced theoretical calculations are needed to quantitatively correlate the energetics of intermediates with reaction kinetics. With ongoing advances in computational power, algorithms, and big data from both experiments and simulations, machine learning could play a transformative role in predicting structure–performance relationships at previously unattainable scales. At present, the 2e− WOR has received less attention and is in its early stages of development compared to the 2e− ORR. Thus, the development of novel computational approaches to screen and guide the design of advanced carbon-based 2e− WOR electrocatalysts is essential. A fundamental mechanistic understanding will elucidate the critical interplay between: (1) reaction intermediate energetics, (2) surface-dependent selectivity determinants (e.g., *OOH binding strength, local coordination environments), and (3) electrochemical microenvironment effects (pH, potential, double-layer structure), ultimately enabling rational design of efficient H2O2 production systems.
While catalyst design has advanced significantly, it is worth noting that the development of electrochemical reactors has lagged behind catalyst design in the past decade. Therefore, system-level optimization is essential for practical implementation of carbon-based electrocatalysts. Particular emphasis should be placed on the design of the overall electrochemical system and the operational reliability of the industrial-grade device for practical applications. Key considerations include O2 mass transfer, device design, electrolyte selection, ion exchange membranes, and reactor configurations. First, the mass transfer and diffusion of reactant species within the interface microenvironment of the carbon electrode needs further investigation.181 Specific optimization strategies involve gas diffusion electrodes with hierarchical porosity (macro/meso/micro pores) to balance O2 transport and flow-through membrane reactors with rotational cathodes. The use of a porous solid electrolyte (PSE) reactor for producing high-concentration and high-purity H2O2 may meet the high stability requirment.186 In addition, optimizing local reaction environments—such as water permeation/wetting, hydrophobicity/hydrophilicity, and reactant concentration—can enhance the performance of carbon-based electrocatalysts. Significant research efforts should focus on advancing reactor engineering, particularly for flow cell configurations and dual-PEM solid electrolyte systems. These reactor designs must simultaneously address two critical challenges: (1) corrosion resistance against both acidic/alkaline electrolytes and concentrated H2O2 (up to 10 wt%), and (2) maintenance of stable three-phase interfaces under industrial current densities (>200 mA cm−2). Innovative architectures incorporating corrosion-resistant materials (e.g., PTFE-coated titanium flow fields, stabilized carbon-PTFE gas diffusion electrodes) and optimized hydrodynamics are crucial for achieving economically viable H2O2 electrosynthesis on a large scale. Furthermore, the H2O2 products are typically generated in a mixture, with solutes in traditional liquid electrolytes ranging from acidic to alkaline pH. Extra separation processes to recover pure H2O2 solutions are therefore required. Using a solid-state electrolyte can avoid contamination of the product solution by extraneous ions.2 Producing H2O2 in neutral solutions offers practical advantages by avoiding pH-related complications, such as electrode degradation or the need for neutralization steps. Seawater, as an abundant and naturally neutral electrolyte, directly supports this goal by providing a sustainable medium for electrochemical H2O2 production via the 2e− ORR pathway. The dissolved oxygen serves as a reactant for carbon-based electrocatalysts, which have demonstrated high efficacy in neutral media. Moreover, the use of seawater eliminates the costs and environmental impacts associated with synthetic pH adjustments, enhancing both scalability and the sustainability of the process. However, challenges such as chloride interference may compromise catalyst stability and selectivity, necessitating the development of robust carbon-based materials and more in-depth mechanistic studies in simulated seawater electrolytes to fully realize these benefits. Consequently, exploring suitable and stable carbon-based catalysts and understanding the catalytic mechanism in neutral simulated seawater electrolytes are urgent and promising areas of research.187 Moreover, sustainable H2O2 production coupled with the high value added chemical synthesis is worth exploring, such as oxidative valorization of glycerol188 and upgrading of cellulosic biomass into valued formic acid.189
This journal is © The Royal Society of Chemistry 2025 |