Qiuyue Yang†
ab,
Jilan Zeng†
ab,
Guowei Yangab,
Xinran Sunab,
Xiahui Linab,
Kunlong Liu
*ab,
Jiayi Chen
*ab,
Sibo Wang
*ab and
Xue Feng Lu
*ab
aState Key Laboratory of Chemistry for NBC Hazards Protection, College of Chemistry, Fuzhou University, Fuzhou 350116, China. E-mail: klliu@fzu.edu.cn; jiayi9236@fzu.edu.cn; sibowang@fzu.edu.cn; luxf@fzu.edu.cn
bState Key Laboratory of Photocatalysis on Energy and Environment, College of Chemistry, Fuzhou University, Fuzhou 350116, P. R. China
First published on 18th July 2025
Water electrolysis hydrogen production technology directly generates high-purity hydrogen through electrochemical water splitting, serving as a key technology for achieving zero-carbon emission hydrogen production. Alkaline water electrolysis demonstrates marked advantages in efficiency and rapidly developing anode catalysts in an alkaline medium. Nevertheless, the sluggish kinetics of the hydrogen evolution reaction (HER) at the cathode in an alkaline environment constitute a fundamental bottleneck that restricts the extensive application of this technology. Platinum, serving as the benchmark catalyst for the HER, is limited in its large-scale development due to its scarcity and high cost. In comparison, carbon-supported platinum-based catalysts exhibit exceptional HER catalytic activity and stability, driven by their unique electronic architecture and the synergistic effect with the support. In this review, we comprehensively examine the latest progress of carbon-supported platinum-based materials for the alkaline HER, summarize the factors contributing to the slow kinetics of the HER in an alkaline environment, and then focus on the strategies for modifying the carbon substrate and synthesizing carbon-supported platinum-based nanomaterials. Finally, the review critically evaluates existing challenges and proposes targeted research directions to advance Pt-based electrocatalysts for practical alkaline hydrogen evolution systems.
Broader contextAmid global carbon neutrality efforts, hydrogen, particularly green hydrogen, is poised to play a pivotal role, with demand projected to surge sixfold by 2050. Alkaline water electrolysis, a cost-effective and stable hydrogen production technology, stands as a key enabler of this transition. Yet its widespread adoption is hampered by the sluggish kinetics of the alkaline hydrogen evolution reaction (HER). While platinum remains the benchmark HER catalyst, its scarcity and expense drive the search for high-performance, scalable alternatives. Carbon-supported Pt-based catalysts, benefiting from tailored electronic structures and carrier interactions, exhibit exceptional activity and durability, offering a viable pathway to industrial-scale implementation. This review examines recent progress in such materials, analyzes the origins of alkaline HER's kinetic limitations, and outlines strategic modifications to carbon supports and Pt nanostructures—critical steps toward unlocking the full potential of green hydrogen and achieving global decarbonization. |
While water electrolysis boasts significant merits for green hydrogen production, its industrial-scale deployment remains constrained by persistent challenges, including excessive energy input, prohibitive capital expenditure, and suboptimal system efficiency.2,8 Water electrolysis involves the hydrogen evolution reaction (HER) at the cathode and the oxygen evolution reaction (OER) at the anode, where the efficient progression of the cathode HER is a critical link for realizing green hydrogen production. However, alkaline HER systems face unique challenges due to the characteristics of the medium. In an alkaline environment, the dissociation kinetics of water molecules are slow, and the presence of hydroxide ions (OH−) significantly alters the reaction pathway and adsorption behavior of intermediates, requiring catalysts to possess excellent dual capabilities for water dissociation and hydrogen adsorption/desorption. Although research on non-noble metal catalysts has made certain progress in recent years, demonstrating hydrogen evolution performance comparable to that of platinum (Pt)-based catalysts, their rapid performance degradation and insufficient structural stability during actual device operation directly restrict their widespread industrial application.9 Therefore, in large-scale industrial hydrogen production, Pt-based catalysts remain the core elements to ensure the high current density and low-energy consumption operation of electrolyzers, and their irreplaceable performance makes them the cornerstone of industrial application for alkaline HER systems. Pt has been recognized as the benchmark HER catalyst because of its optimal hydrogen binding energy (HBE), which positions it near the apex of the HER activity volcano plot. This plot correlates experimental exchange current densities (j0) with theoretical Gibbs free energy (ΔGH*) values, revealing a distinct “volcano-shaped” relationship: the HER activity peaks when HBE approaches that of Pt, while deviations in HBE (either stronger or weaker) lead to a dramatic decline in activity.10,11 The Sabatier principle explains that while excessively weak HBE hinders hydrogen adsorption, overly strong HBE impedes hydrogen desorption, with both extremes limiting reaction efficiency.12–14
Nevertheless, the inherent scarcity of Pt and its fluctuating price dynamics fundamentally constrain the large-scale commercialization and economic feasibility of alkaline water electrolysis. Besides, under the alkaline HER conditions, the dissolution loss of active components and agglomeration-induced deactivation of Pt catalysts are the core issues restricting the long-term efficient operation of the catalytic system.15 From a materials chemistry perspective, the dissolution mechanism of Pt exhibits multi-factor synergy. Specifically, under high potential, OH− as a strong nucleophile directly attacks the Pt lattice, promoting its oxidation to form soluble hydroxides such as Pt(OH)4. Meanwhile, the coordination compounds formed by OH− and Pt significantly reduce the activation energy of the dissolution reaction, accelerating the loss of active sites. Additionally, the strong adsorption of H* on the Pt surface induces lattice distortion, destroying the crystal structure stability and further exacerbating Pt dissolution.16,17 The agglomeration of Pt nanoparticles (NPs) stems from the coupling of thermodynamic driving forces and environmental factors. First, the high surface energy of small-sized Pt NPs triggers Ostwald ripening, promoting particle size homogenization through a “dissolution–deposition” mechanism. Second, corrosion of the support material in alkaline environments disrupts the anchoring structure of Pt NPs, leading to migration and agglomeration of active components. Third, the ordered reconstruction of interfacial water molecules alters the electrostatic repulsion between NPs, facilitating their collision and fusion. Finally, the mechanical impact of H2 bubbles generated during the HER also accelerates the agglomeration of Pt NPs.18 These stability issues highlight the need to break through the traditional single-performance improvement paradigm and construct an “activity-stability-economy” trinity design strategy for alkaline HER catalysts. On the basis of ensuring the intrinsic catalytic activity of noble metals, it is urgent to achieve atomic-level dispersion of Pt and inhibit particle agglomeration through means such as nanostructure regulation, support interface engineering, and composite catalytic system design. Meanwhile, enhancing corrosion resistance via interfacial electronic structure optimization can ultimately achieve synergistic enhancement of catalytic efficiency and material utilization.19,20
Noble metals Pt possess superior surface free energy, usually loading on support during synthesis and reaction processes to enhance dispersion and prevent aggregation.21 Carbon-based materials are generally recognized as compelling and intriguing catalyst supports. Compared with other supports such as metal oxides, they feature high electrical conductivity, excellent chemical stability, and tuneable surface electronic structures. Firstly, the highly conductive framework facilitates charge transfer to the catalyst surface. In contrast, metal oxide supports (e.g., Ni-based) usually have poor conductivity, which easily becomes a bottleneck for charge transfer. Secondly, the extensive specific surface area and porous architecture enhance active site exposure while optimizing mass transport efficiency. However, Ni-based supports have a small specific surface area and irregular pore structures, leading to the aggregation of Pt particles and a reduction in the number of active sites.22 Thirdly, excellent electrochemical stability is maintained over a wide potential window, demonstrating tolerance to harsh alkaline electrolytic environments. Finally, defect engineering and elemental doping enable precise electronic structure modulation through strong metal–support interactions (SMSI), offering multidimensional strategies for catalytic performance optimization.23 Moreover, the surface functional groups of carbon supports can tune the adsorption energy of Pt, optimizing the adsorption and desorption of reaction intermediates, whereas the surface properties of metal oxide supports are difficult to regulate.24,25 However, carbon materials are prone to corrosion in alkaline media, leading to poor stability.26 The corrosion mechanism of carbon mainly originates from electrochemical oxidation, nucleophilic attack of OH−, galvanic corrosion, and dissolution–deposition effects. Under high potential, OH− generates ˙OH radicals that attack the carbon skeleton, oxidizing it into oxygen-containing functional groups such as carbonyl and carboxyl, which eventually decompose into soluble carbonates. OH− also directly attacks defect sites on carbon, weakening the skeleton stability.27 When carbon supports metal catalysts, the potential difference between the metal and carbon forms a micro-battery, with carbon acting as the anode to accelerate oxidation. Additionally, the metal may catalyze the generation of strong oxidizing species to indirectly corrode carbon. Furthermore, carbonates formed by corrosion redeposit with changes in pH or potential, blocking carbon pores. Meanwhile, the collapse of carbon nanostructures causes instability and aggregation of metal particles.28
Currently, significant endeavors have been committed to deciphering the mechanistic insight and application boundaries of Pt-based electrocatalysts in alkaline HER. These scholarly endeavors have been systematically analyzed in multiple authoritative reviews that map the evolution trajectory of this field. For example, Liu et al. systematically investigated the electronic structure–activity correlation mechanism in carbon-supported noble metal catalysts and proposed to improve the slow HER kinetics by optimizing the synthesis strategy and structural characterization.12 Sui et al. comprehensively reviewed the influence of support materials on noble metal single-atom catalysts (SACs), elaborating the critical roles of carriers—including anchoring effects, SMSI, and synergistic catalytic mechanisms—thereby establishing a novel theoretical paradigm to address key challenges in the scalable application of noble metal SACs.29 Ma et al. critically delineated mechanistic modulation approaches to overcome kinetic limitations in alkaline HER, while establishing design principles for advanced Pt-based electrocatalysts with record-breaking efficiency.8 Zhang et al. conducted a critical investigation into the mechanisms underlying HER activity enhancement, systematically elucidating strategies for modulating the interactions between active sites and reaction intermediates.1 Furthermore, recent advances in support materials are categorized and elaborated in detail, focusing on their anchoring mechanisms.30
The aforementioned review has comprehensively explored critical areas, including synthesis optimization, the HER mechanism, the structure–activity relationship, and practical applications. Notably, despite numerous investigations revealing that the HER kinetics in alkaline media exhibit a much slower rate, with a 2–3 orders-of-magnitude decrease relative to the kinetics in acidic media, a clear and consistent understanding of this phenomenon remains elusive. Consequently, we commence by summarizing the latest research findings regarding the sluggish reaction kinetics of the alkaline HER, aiming to shed light on this long-debated scientific conundrum. Building on this foundation, we methodically examine recent progress in carbon-supported Pt nanomaterials, focusing on three interconnected dimensions: strategic modifications of carbon matrix support, innovative synthesis approaches for carbon-supported Pt nanomaterials, and their resulting alkaline HER electrocatalytic performance in terms of activity, stability, and reaction kinetics. Finally, we critically evaluate the prevailing challenges within this field, such as limited understanding of alkaline interfacial processes and scalability constraints of advanced catalysts, while proposing feasible research directions grounded in experimental and theoretical advancements to guide the development of next-generation alkaline HER electrocatalysts (Fig. 1).
![]() | ||
Fig. 1 Schematic diagram of the principal guidelines and performance improvement strategies for carbon-support Pt-based electrocatalysts for the alkaline HER. |
In acid:
Volmer step: H3O+ + e− + M → M–H* + H2O |
Heyrovsky step: M–H* + H3O+ + e− → H2 + H2O + M |
Tafel step: 2M–H* → H2 + 2M |
In alkaline:
Volmer step: H2O + e− + M → M–H* + OH− |
Heyrovsky step: M–H* + H2O + e− → H2 + OH− + M |
Tafel step: 2M–H* → H2 + 2M |
As is well known, the reaction rate of the HER in alkaline media is generally 2–3 orders of magnitude lower than that in acidic media.34–36 Although the basic reaction mechanism of the alkaline HER has been verified to a certain extent, there are still many controversies in the research of its kinetics and energetics, and scholars in the fields of computational chemistry and electrochemistry have not yet reached a consensus. This situation has inspired researchers' strong interest in exploring the underlying reasons from multiple perspectives.37–40 For example, Yan et al. conducted tests on polycrystalline Pt in electrolytes with different pH values (ranging from 0 to 13), deeply investigating the correlation between the hydrogen oxidation/evolution activity and the experimentally measured HBE.41 Their research uncovers that HBE which varies with pH is the fundamental reason for the pH-dependent kinetics. More precisely, acidic conditions are marked by an elevated concentration of H+ ions. These H+ ions can rapidly adsorb onto the surface of Pt, facilitating the HER Volmer step. Therefore, HBE is regarded as a key descriptor for the HER. Wei et al. briefly summarized the slow reaction rate of the HER under alkaline conditions from the aspects of the barriers of water dissociation and the pH dependence of HBE, providing theoretical directions for engineering electrocatalysts to boost the efficiency of the alkaline HER activity.31 Moreover, Zheng et al. conducted an in-depth analysis of the alkaline HER from multiple dimensions, such as the energetics and kinetics of the hydrogenation reaction on the metal surface and the controversy over activity descriptors.42 As characterization techniques and computational chemistry continue to evolve, Qiao et al. revisited and systematically re-evaluated three prominent theories: the water dissociation theory, the HBE theory, and the interfacial water and/or anion transfer theory.43 This conclusion has also been confirmed by the research of Menezes et al.44 Similarly, Sun et al. systematically and thoroughly summarized the vital elements influencing the observable HER kinetics under alkaline conditions, covering the HBE theory, the role of OH*, the interfacial water, and the influence of alkali metal cations (AM+).45 Drawing on the extensive research achievements outlined above, we offer a comprehensive overview of the current key influencing factors and present our distinctive perspectives.
The water dissociation theory reveals the key regulatory role of the pH environment in HER kinetics.14 Under acidic conditions, high concentration H+ directly participates in the initial Volmer step as reactants. The sufficient supply of H+ significantly reduces the activation energy of the Volmer reaction, enabling efficient generation of H* intermediates on the catalyst surface and accelerating subsequent Heyrovsky or Tafel reactions for high-efficiency hydrogen production. In stark contrast, alkaline media lack a high concentration of free H+, and the generation of H* originates from water dissociation steps, forcing the HER process to overcome a higher energy barrier.46 The hydroxide binding energy (OHBE), defined as the binding energy between the catalyst and OH− intermediates, directly influences the difficulty of water dissociation. Excessively strong OHBE leads to over-adsorption of OH− at active sites, blocking the generation pathway of H*; whereas insufficiently weak OHBE fails to effectively promote water dissociation, resulting in inadequate proton supply.42 Therefore, under alkaline conditions, the accumulation of OH− concentration on the catalyst surface forms a localized high OH− environment, which thermodynamically and kinetically inhibits the water dissociation process. The competitive adsorption between OH* and H* on active sites reduces the number of H* adsorption sites and alters the electronic structure of active sites, further hindering the reaction progress.39 Furthermore, OH* and H* exhibit strong competitive adsorption on the limited active sites of the catalyst. On one hand, this competition reduces the number of active sites available for H adsorption and reaction. On the other hand, adsorbed OH* may alter the electronic structure or steric hindrance of active sites, indirectly weakening the adsorption strength of H* or hindering subsequent Heyrovsky or Tafel steps.40 This view has been verified by constructing Pt (111) catalysts with additional water dissociation sites (such as (M(OH)2 clusters, where M = Fe, Co, Ni)).47
The HBE theory provides a critical theoretical interpretation of the HER mechanism from a kinetic perspective.9 Ideal HER catalytic materials need to tune their binding energy with H* to an optimal range–an excessive HBE will significantly raise the activation energy barrier of the reaction, especially in high-pH systems, where this effect triggers a drastic decrease in the HER rate and exerts a remarkable inhibitory effect on catalytic efficiency.48 At the essence of catalysis, an excessive HBE leads to hindrance in the desorption process of H* from the catalyst surface, significantly increasing the energy barrier of the rate-determining step. In an alkaline environment, the competitive adsorption of OH− on the catalyst active sites and the change in proton supply pathway further amplify the negative impact of excessive HBE on reaction kinetics, resulting in a non-linear decay trend of catalytic efficiency.49 This theoretical framework provides a clear thermodynamic-kinetic optimization direction for the design of efficient HER catalysts under high-pH conditions. The effect of interfacial water plays an indispensable role in the HER kinetics. The configuration of interfacial water is deeply intertwined with the pH-dependent changes in HER activity.43 As the pH value gradually increases, the kinetic process of alkaline HER and the hydrogen generation rate show a decreasing trend, the charge transfer resistance correspondingly increases, and the adsorption rate of hydrogen ions also slows down accordingly. From a theoretical mechanism perspective, the correlation between the electrode potential and the potential of zero free charge (PZFC) has a significant impact on the reorganization energy of interfacial water molecules, thereby exerting a regulatory effect on the hydrogen adsorption process and reaction activity through this mechanism.41 As a key descriptor of activity, the structure of the interfacial water will change with the variations in water concentration, the type of supporting salt, and the pH value. Specifically, different water concentrations and supporting salts affect the reduction activity of water. Typically, organic solvents are added or supporting salts are regulated to alter the water concentration. At higher water concentrations, water–water interactions are enhanced, and the formed water nanodomains promote the reaction through the Grotthuss diffusion mechanism, corresponding to higher HER activity.44 Additionally, as the pH increases, the first layer of interfacial water transitions from a proton acceptor to a donor, while the activation energy for water dissociation changes in a specific order. The AM+ theory further expands our understanding of the HER kinetics.45 On the one hand, the complex of hydroxyl-water-alkali metal cations adsorbed on the surface of the catalyst can promote the transfer process of OH*, and AM+ itself can also significantly change the transfer kinetics of OH*, thus having an important impact on the HER under alkaline conditions. On the other hand, the interaction between the metal cations and the interfacial water will change the architecture of the interfacial water, and further affect the HER rate. It merits attention that as a result of the concentration distribution of OH* and AM+ is closely related to the pH value, the pH environment on the surface of the catalyst becomes a key factor in controlling the HER kinetics.50
Although there are multiple explanations for the sluggish alkaline HER kinetics, controversies still exist in the related theoretical research, with a unified system and universal model yet to be established. Future research should focus on three primary aspects: (1) optimizing catalyst design by deepening the understanding of reaction mechanisms, which involves regulating metal cation-related factors, optimizing catalyst–intermediate interactions, reducing water dissociation energy barriers, and enhancing reaction rates; (2) combining multi-disciplinary methods through the integration of experimental and theoretical approaches, specifically using in situ characterization techniques and density functional theory (DFT) calculations to investigate intermediate adsorption and charge transfer mechanisms, thereby providing theoretical support for catalyst design; (3) exploring new materials such as transition–metal-based composites and metal–organic frameworks (MOFs) to improve alkaline HER kinetics. Notably, carbon materials have demonstrated exceptional potential in alkaline water electrolysis, where carbon modification synergistically optimizes the kinetics of the HER through multi-dimensional regulation.8–12 At the morphological regulation level, nanostructures such as carbon nanosheets and porous networks increase the exposure of active sites, control the size and dispersion of Pt particles, and shorten the ion diffusion path by optimizing the spatial topology. This synergizes with the pore structure to enhance the accessibility of active sites. In defect engineering optimization, defect sites on the carbon support anchor Pt particles to inhibit agglomeration while inducing interfacial charge polarization, reducing the energy barrier for water dissociation and enhancing the intrinsic catalytic activity. For electronic structure regulation, heteroatom doping (e.g., N, S) leverages electronegativity differences to reconstruct the electron cloud on the carbon surface, optimizing the ΔGH* of Pt active sites via SMSI and accelerating the Volmer step kinetics. In terms of surface wettability regulation, hydrophilic/hydrophobic modifications optimize the contact efficiency of the three-phase interface (solid–liquid–gas), where hydrophobicity promotes H2 bubble detachment and hydrophilicity enhances electrolyte penetration. For interfacial synergistic catalysis, the electronic coupling between the carbon support and Pt forms bifunctional catalytic sites that cooperatively promote water dissociation and hydrogen recombination, enhancing stability. As summarized in Table 1, carbon-supported Pt-based HER electrocatalysts exhibit competitive overpotentials at 10 mA cm−2, favorable Tafel slopes, and enhanced stability in alkaline electrolytes. The following sections will systematically elaborate on recent advancements in carbon-supported Pt-based materials in alkaline HER systems.
Catalyst | Pt loading | η10 (mV) | Tafel slope (mV dec−1) | Mass activity (V vs. RHE) | Stability | Ref. |
---|---|---|---|---|---|---|
Pt/G–CNFs | 3.4 μg cm−2 | 177 | 72 | 0.626 A mgPt−1 at −0.1 V | 90 h at 10 mA cm−2 | 58 |
NiPt–NC | 0.27 mg cm−2 | 48 | 50.3 | 0.67 A mgPt−1 at −0.02 mV | 30 h at 1.56 V | 59 |
Ru–PtFeNiCuW/CNTs | 7.67 wt% | 16 | 27.9 | — | 50 h at 10 mA cm−2 | 61 |
Pt/VGAr-5 | 0.71 μg cm−2 | 124 | 36 | 44.8 A mgPt−1 at −0.2 V | Over 24 h at 100 mA cm−2 | 63 |
Pt–Ni4Mo/CNT | 0.14 mg cm−2 | 18.6 | 37.4 | 6.97 A mgPt−1 at −0.07 V | 200 h at 100 mA cm−2 | 67 |
Ru1Pt2@rGO | 10.81 wt% | 9 | 20.9 | 2.67 A mgmetal−1 at −0.05 V | 3000 cycles | 69 |
Pt–NiOx–H2/C | 50 μg cm−2 | 39.6 | 24.49 | 6.07 A mgPt−1 at −0.1 V | 25 h at 10 mA cm−2 | 81 |
Pt/8-NCNT | 0.14 mg cm−2 | 17 | 33.3 | 3.13 A mgPt−1 at −0.1 V | 100 h at 500 mA cm−2 | 82 |
(Ru–P)#Pt/C | 0.76 wt% | 17 | 27 | — | 24 h at 10 mA cm−2 | 86 |
P–Pt1Ni2NH/NGA | 9.1 μg cm−2 | 15 | 37 | 13.4 A mgPt−1 at −0.1 V | 20 h at 10 mA cm−2 | 90 |
Pt-30/NCM | 3.39 wt% | 19.5 | 40.8 | 1.91 A mgPt−1 at −0.05 V | 3000 cycles | 91 |
Pt/NSC | 7.677 wt% | 17.8 | 30.59 | 4.90 A mgPt−1 at −0.1 V | 50 h at 50 mA cm−2 | 93 |
PtMo–NC | 13.52 wt% | 35 | 32 | 3.68 A mgPt−1 at −0.05 V | 5000 cycles | 109 |
Pt/Ce–N–C | 32 μg cm−2 | 22 | 33.8 | 7.16 A mgPt−1 at −0.05 V | 200 h at 100 mA cm−2 | 110 |
Pt5/HMCS | 7.6 μg cm−2 | 46.2 | 48.1 | 3.23 A mgPt−1 at −0.07 V | 3000 cycles | 112 |
Pt@CoN4–G | 12 μg cm−2 | 39 | 29 | 1.15 A mgPt−1 at −0.05 V | 10![]() |
113 |
Pt–AC/Cr–N–C | 1.72 wt% | 19 | 30 | 7.9 A mgPt−1 at −0.05 V | 24 h at 10 mA cm−2 | 114 |
Pt/TC2 | — | 58 | 65.5 | 1.115 A mgPt−1 at −0.05 V | 10 h at 10 mA cm−2 | 118 |
Pt/C60-2 | 0.053 mg cm−2 | 25 | 55 | — | 100 h at 10 mA cm−2 | 119 |
c-PRPDWNT/C | 0.1 mg cm−2 | 10 | 30.8 | 47.8 A gmetal−1 at −0.02 V | 100 h at 50 mA cm−2 | 120 |
Pt3Co@NCNT | 0.4 mg cm−2 | 36 | 34.8 | — | 100 h at 100 mA cm−2 | 123 |
Con–Pt1@NPC | 6.47 wt% | 20 | 29.07 | 1.36 A mgPt−1 at −0.05 V | 100 h at 10 mA cm−2 | 124 |
PtC60 | 0.4 mg cm−2 | 24.3 | 30 | 1.55 A mgPt−1 at −0.1 V | 50 h at 10 mA cm−2 | 125 |
PtRu/CNT@MO2−x | 4.6 μg cm−2 | 53 | 48 | 12.3 A mgmetal−1 at −0.07 V | 2000 cycles | 130 |
Pt–Co2P/CC | 3.33 wt% | 2 | 44 | — | 200 h at 100 mA cm−2 | 131 |
Pt@DG | 1.57 wt% | 37 | 53 | 26.05 A gPt−1 at −0.1 V | 5000 cycles | 132 |
0D carbon nanomaterials have garnered significant research interest owing to their tunable electronic structures, superior electrical conductivity, and exceptional chemical stability, as evidenced by recent breakthroughs in alkaline HER electrocatalysis.55,56
For instance, Dang et al. synthesized CDs/Pt-PANI catalysts by leveraging the strong bonding between the surface functional groups of carbon dots (CDs) and polyaniline nanosheets (PANI), which is confirmed by Fig. 2a.57 Thanks to the synergistic effect of CDs, Pt, and PANI, these catalysts manifest superior efficiency: only 56 mV is needed to reach 10 mA cm−2 in 1 M KOH, outperforming 20 wt% Pt/C which requires 86 mV at an identical current density. To further investigate the influence of CDs on catalysts, Wang et al. proposed a heterostructure composed of graphene–carbon nanofibers (G–CNF) loaded with Pt nanoclusters and studied the effects of different contents of Pt nanoclusters and CDs on the HER performance.58 Using a mixture of calcium lignosulfonate (LS) and CDs as the precursor, they formed G–CNF through electrospinning and subsequent carbonization. As shown in Fig. 2b, the conductivity of the prepared G/CNFs gradually improved with the increase of the added CD content, which contributed to the enhancement of the alkaline HER activity. In addition to regulating the HER activity by optimizing the electron transport performance, CDs can also use surface defect sites as anchoring sites to load Pt NPs. By triggering the SMSI effect, the intrinsic activity and structural stability of the catalyst are synergistically improved. This strategy regulates the electron coupling between the support and metal active sites through defect engineering, which can not only promote the adsorption–desorption kinetics of reaction intermediates but also inhibit the agglomeration and sintering of Pt particles through interfacial electron reconstruction, realizing the dual optimization of catalytic performance. Liu et al. anchored Pt on Ni and N co-doped CDs (Ni–NC) with rich defect structures and atomically dispersed Ni–N5 sites to prepare NiPt–NC electrocatalysts with high dispersion and low Pt loading.59 Fig. 2c revealed that the interaction between Pt NPs and the surrounding Ni–N5 sites led to electron enrichment around Pt, optimizing the desorption of OH* and the adsorption/dissociation of H2O (volmer step), enhancing the intrinsic activity of Pt for the alkaline HER, with a 48 mV overpotential at 10 mA cm−2. These studies provide constructive guidance for the subsequent study of 0D carbon materials as supports.
![]() | ||
Fig. 2 (a) FTIR spectra of CDs/Pt-PANI-4, CDs-PANI, and PANI-Na2SO4. Reproduced with permission.57 Copyright 2019, Elsevier. (b) Electronic conductivity curves of G–CNFs with different contents of CDs. The pyrolysis temperature is 1400 °C. Reproduced with permission.58 Copyright 2022, Wiley-VCH GmbH. (c) Free energy diagrams of the elementary steps in the HER. Reproduced with permission.59 Copyright 2025, Elsevier. (d) HER polarization curves of CNT, Pt/C, and M–Ru/CNT (M = Rh, Pd, Ir) catalysts in N2-saturated 1.0 M KOH solution. (e) ΔGH* of Pt–Ru/CNT, Rh–Ru/CNT, Pd–Ru/CNT, Ir–Ru/CNT, and Ru/CNT. Reproduced with permission.60 Copyright 2021, Wiley-VCH GmbH. (f) HER performance in 1.0 M KOH electrolytes. Reproduced with permission.61 Copyright 2024, Wiley-VCH GmbH. (g) TEM image of Pt3Co/NG-700 catalyst. Reproduced with permission.62 Copyright 2020, American Chemical Society. (h) Calculated Gibbs free energy of the alkaline HER reaction pathway on Pt/graphene edge, Pt/basal graphene, Pt (111) surface, and a zigzag graphene edge. Reproduced with permission.63 Copyright 2022, Wiley-VCH GmbH. (i) SEM image of 3D porous WGA. Inset: Optical picture of WGA. Reproduced with permission.64 Copyright 2021, Wiley-VCH GmbH. |
Carbon nanotubes (CNTs) and carbon nanofibers (CNFs), as prototypical 1D carbon nanomaterials, have garnered significant attention in alkaline HER electrocatalysis owing to their exceptional electrical conductivity, superior mechanical robustness, remarkable chemical stability, and tunable electronic structures.65,66 For instance, Zhang et al. prepared a series of surfactant-free noble metal doped Ru/CNT (M–Ru/CNT, M = Pt, Rh, Pd, Ir, CNT stands for carbon nanotube) by microwave reduction method, using CNTs with large specific surface areas, strong conductivities and high mechanical strengths as the carrier to make the active ingredients highly uniformly dispersed and increase the number of active sites.60 DFT demonstrates that Pt–Ru/CNT can significantly optimize ΔGH* and accelerate the release of H2 (Fig. 2e), thereby promoting alkaline HER kinetics, requiring only 4 mV overpotential to drive 10 mA cm−2 in 1.0 M KOH (Fig. 2d). Liu et al. prepared small-sized (3–4 nm) M–Ni4Mo NPs (M = Pt, Ru, Pd) supported on CNT via a simple, rapid, and solvent-free microwave reduction method.67 The Pt–Ni4Mo/CNT catalyst with optimal atomic ratio and loading exhibits exceptional HER performance in 1 M KOH electrolyte, achieving an overpotential of 18.6 mV at 10 mA cm−2. Benefiting from the SMSI effect, the catalyst also demonstrates excellent long-term stability, maintaining catalytic activity without significant decay after continuous operation at a high current density of 100 mA cm−2 for 200 hours. This performance advantage stems from the synergistic regulation of the interfacial electronic structure of the catalyst. On one hand, the electronic coupling between the Ni4Mo alloy and the CNT support optimizes the hydrogen adsorption energy barrier; on the other hand, the SMSI induced interfacial charge reconstruction effectively inhibits the agglomeration and sintering of Pt active sites, achieving a dual improvement in catalytic activity and structural stability. Moreover, Wang et al. prepared Ru-doped PtFeNiCuW octahedral high-entropy alloy (HEA) nanocrystals on CNTs (Ru–PtFeNiCuW/CNT).61 High-angle annular dark field-scanning transmission electron microscopy (HAADF-STEM) images demonstrate that the HEA octahedrons are homogeneously distributed on the CNTs. Fig. 2f reveals that the catalyst exhibits a low overpotential of 16 mV in an alkaline electrolyte (at a current density of 10 mA cm−2). In a two-electrode alkaline system, it can operate stably for 1200 h at a current density of 50 mA cm−2. The above work demonstrates that CNTs, with their large specific surface area and excellent stability, serve as superior supports for loading Pt, holding profound significance for advancing the commercial application of carbon-supported Pt-based catalysts.
Graphene, a representative 2D carbon-based material, boasts unique structural advantages such as a large surface area, high electrical conductivity, and good chemical stability. These properties endow graphene-based electrocatalysts with enhanced alkaline HER performance.68 Lin et al. anchored Pt3Co NPs onto nitrogen-doped graphene (NG) (Pt3Co/NG-700) through a simple pyrolysis method.62 The SMSI between NG and Pt3Co led to the formation of small-sized and ordered Pt3Co NPs during the high-temperature treatment, as demonstrated in Fig. 2g. Owing to the synergistic coupling effect between the ordered Pt3Co nanocrystals and the NG support, the mass activity of Pt3Co/NG-700 reaches 51.8 times compared with the commercial Pt/C. This synthesis strategy regulates the growth kinetics of NPs through the SMSI effect, not only achieving high-density uniform distribution of active sites but also optimizing the hydrogen adsorption energy barrier through the electronic coupling between the support and the metal, providing new ideas for the design of efficient electrocatalytic systems. Tsounis et al. utilized the enhanced reactivity of graphene edges to anchor ultra-low mass-loaded Pt SACs onto the edges of edge-rich, vertically oriented graphene.63 This enhanced reactivity enables the anchoring and tuning of the electronic states of SACs. Specifically, the edge-anchored Pt SACs have a firm coupling with the π electrons of graphene. This interaction results in a higher occupancy of the Pt 5d orbitals, shifting the d-band center towards the Fermi level, thus improving the adsorption of H* in the HER (Fig. 2h). To arrive 10 mA cm−2, 1 μg Pt/C and Pt/VG Ar-5 requires overpotentials of 199 mV and 124 mV, respectively. Besides, Yang et al. prepared Ru–Pt heterogeneous bimetallic NPs (RuxPty@rGO) on reduced graphene oxide (rGO) via a simple “simultaneous reduction” hydrothermal method, simplifying the catalyst synthesis process.69 The rGO as a support can significantly enhance the catalyst's electrical conductivity. Ru1Pt2@rGO with the optimal Ru/Pt ratio exhibits a faster Volmer–Tafel mechanism in 1 M KOH solution, with an overpotential of 9 mV at a current density of 10 mA cm−2, outperforming other Ru-based or Pt-based monometallic catalysts reported in the literature. In the future, graphene is expected to leverage its unique structure to innovate carbon-supported Pt-based catalysts by optimizing electron conduction and enhancing interfacial stability.
Compared with low-dimensional carbon materials with a simple structure, 3D carbon materials have a complex structure and possess significant advantages such as high mechanical strength, strong designability, and excellent environmental adaptability.70–72 Li et al. electrochemically deposited Pt NPs on a tungsten oxide/reduced graphene oxide aerogel (WGA) support.64 The WGA exhibits a typical 3D interconnected porous structure (Fig. 2i), featuring abundant oxygen vacancies and hierarchical pore sizes, which plays the role of anchoring Pt NPs, providing continuous mass transport and electron transfer channels, and regulating the electronic state of the Pt surface. Therefore, Pt/WGA exhibits high HER activity and durability, with an overpotential of 42 mV and 40 h chronopotentiometry at 10 mA cm−2. Park et al. designed a new type of CNT-based electrode that can produce a 3D porous structure when the CNT array is cross-woven.73 The 3D porous structure contains square pores that provide high conductivity and a large surface area, and the HER electrocatalyst can be dispersed in a hierarchical porous CNT framework, which is called the M_CNTs (M: HER catalyst) electrode. The CNT-based NiFeOx and Pt electrodes exhibited sustained performance without any performance loss during anion-exchange-membrane water electrolysis (AEMWE) operation for 40 h. 3D carbon materials show extremely broad application prospects in the field of carbon-supported Pt-based catalysts for alkaline HER, and their unique structural advantages will continue to provide support for catalytic performance optimization. In terms of activity improvement, the large specific surface area of 3D carbon materials can significantly increase the loading sites of Pt, promote the uniform dispersion of Pt particles, reduce agglomeration, and thus increase the number of active sites of the catalyst. In terms of stability, 3D carbon materials have good mechanical and chemical stability, which can provide stable support for Pt in an alkaline environment, inhibit the dissolution and agglomeration of Pt during the reaction, and prolong the service life of the catalyst.
The morphological dimensions of carbon materials considerably impact catalyst performance. 0D carbon materials leverage quantum confinement effects and high surface activity; adjusting particle size and functional groups optimizes catalyst adsorption and dispersion. 1D carbon nanotubes, characterized by high aspect ratios, conductivity, and hollow structures, enhance catalyst loading via diameter control. 2D graphene utilizes its large surface area and stability, boosting catalyst anchoring through defect introduction. 3D carbon materials with porous architectures enable uniform catalyst dispersion and protection. Combining multi-dimensional carbon materials with precise control represents a critical future direction for catalyst performance improvement.
![]() | ||
Fig. 3 (a) Raman spectra of Co1NC and NC. (b) Raman spectra of Pt1/Co1NC and PtNP/NC. Reproduced with permission.78 Copyright 2022, Elsevier. (c) The corresponding Raman spectra of NG after N removal (i.e., DG). (d) DFT calculations with the models of PG, DG585, DG7557, DG5665, and DGvacancy. Reproduced with permission.79 Copyright 2020, American Chemical Society. (e) Water contact angles of (up) Pt SASs/AG and (down) Pt/C. Reproduced with permission.80 Copyright 2019, The Royal Society of Chemistry. (f) Pt L3-edge EXAFS spectra (inset: the magnified pre-edge XANES region). (g) Mass activity of catalysts in 1.0 M KOH solution. Reproduced with permission.81 Copyright 2025, Wiley-VCH GmbH. (h) The adsorption free energies of H*. (i) The free energy profiles for the HER in an alkaline medium from H2O to H2. Reproduced with permission.82 Copyright 2022, Wiley-VCH GmbH. |
As expected, the increase in hydrophilicity of the carbon support surface contributes to the enhancement of HER activity. Additionally, metal oxides show strong potential in functionalizing carbon supports, which can significantly enhance their interaction with Pt species. Sun's group reported a surface functionalization strategy to provide specific catalytic activity and enhance the interaction with metal species by modifying transition metal oxide clusters on commercial carbon black.81 Specifically, NiOx clusters were modified on commercial carbon black, and then Pt clusters were synthesized by the alkali-ethylene glycol method and anchored on the functionalized carbon black surface. As verified in Fig. 3f, Pt–NiOx/C exhibited higher white line intensity than Pt/C, indicating that the interaction at the heterostructure interface contributes to the increase in the average oxidation state of Pt. The Ni K-edge X-ray Absorption Near-Edge Structure (XANES) spectrum confirmed that the valence state of Ni was reduced due to the transfer of electrons from Pt clusters to NiO clusters at the heterostructure interface. In the case of NiOx clusters decorated with carbon black, a strongly coupled cluster–cluster heterostructure (Pt–NiOx/C) consisting of Pt clusters and NiOx clusters is formed, which can greatly enhance the evolution kinetics of alkaline hydrogen, achieving 39.6 mV at 10 mA cm−2, while Pt/C requires 131 mV (Fig. 3g). Furthermore, the synergistic effect of different functional groups with Pt has been demonstrated to accelerate the HER reaction kinetics. For instance, Yu et al. integrated the advantages of Pt-based catalysts and nitrogen-doped carbon materials, and designed a series of Pt/X-NCNT (X = 4, 8, 12, 16) catalysts by virtue of the SMSI effect.82 DFT calculations demonstrate that the Pt surface with pyridinic-N and pyrrolic-N functional groups exhibits a slightly negative ΔGH*, which is much lower than that of Pt/CNT (Fig. 3h), revealing the electronic coupling between Pt active sites and nitrogen functional groups. The HER free energy curve from H2O to H2 in the alkaline medium can be seen in Fig. 3i, further proving the fast HER kinetics, with an overpotential of 17 mV at 10 mA cm−2. This provides a new paradigm for constructing electrocatalytic systems with both high activity and stability through the synergistic regulation of functional groups and metal active centers. The above works reveal that functionalization of carbon surfaces can enhance the performance of catalysts in alkaline HER by strengthening their interaction with Pt and optimizing the interface.
In conclusion, surface functionalization is an efficient strategy for modifying carbon supports. It can notably improve the hydrophilicity of carbon supports, enabling their uniform dispersion in aqueous solutions and facilitating the loading and dispersion of active components. Moreover, surface modification can regulate properties such as the electronic structure and defect density of carbon supports, enhancing the interaction with active components and thus optimizing the activity, stability, and selectivity of catalysts.
![]() | ||
Fig. 4 (a) Schematic summary of the heteroatom doping configurations: (top row, from left to right) pr-N, py-N, g-N, N–O, B–2C–O, B–3C, B–C–2O, P–3C(–O) and P–2C(–2O); (bottom row, from left to right) th-S, S–2O, py-O, C–O–C, C–OH, C![]() |
For example, Tian et al. precisely controlled the dispersion and coordination of Pt on hierarchical nitrogen-doped carbon nanocages (hNCNCs) through thermally driven Pt migration.91 The hNCNC support has a large specific surface area (620 m2 g−1), coexisting micro-meso-macroporous structures, a high N content (9.6 atomic percentage), and high intrinsic conductivity (1362 S m−1), affecting the HER performance by influencing the Pt coordination environment. Due to differences in the coordination environment, Pt–N2Cl2 shows the optimal HER activity (6 mV at 10 mA cm−2), while Pt–N4C4−x exhibits poor HER activity (321 mV at 10 mA cm−2). Yang et al. constructed hierarchical porous phosphide Pt–Ni nano-hexa-pods/nitrogen-doped graphene aerogel (P–PtNiNH/NGA) through an oxidation–phosphatization controlled reconstruction strategy (Fig. 4b).92 The hierarchical porous structure of NGA and its excellent anchoring effect on P–PtNiNH prevent the aggregation of Pt NPs and Ostwald ripening. As displayed in Fig. 4c, the dispersion of P–PtNiNH in NGA confirms this point. P–PtNiNH/NGA exhibits excellent alkaline HER activity and remarkable stability, with an ultra-low overpotential of 15 mV at a current density of 10 mA cm−2. Li et al. uniformly anchored Pt nanoclusters (∼1.6 nm) onto an ordered macroporous nitrogen-doped carbon support (Pt-30/NCM) through the polystyrene sphere templating method followed by an impregnation method.93 NCM features a uniform polyhedral nitrogen-doped carbon framework (∼675 nm) and an arrangement of interconnected ordered macropores (∼120 nm) (Fig. 4d), which can form channels for the rapid transport of electrons, ions, and molecules. Moreover, the DFT result in Fig. 4e confirms that the active Pt nanoclusters can interact with the nitrogen-doped carbon, inducing a favorable electron distribution, which is beneficial for the dissociation of water in an alkaline medium and the subsequent desorption of intermediate products (OH* and H*). Pt-30/NCM exhibits outstanding HER activity (η10 = 19.5 mV) in 1.0 M KOH, with only 5 mV attenuation at 10 mA cm−2 after 3000 cycles. In summary, N doping in carbon can regulate the electronic structure and surface properties of carbon supports, enhancing the anchoring effect on Pt species, inhibiting their agglomeration and dissolution, and promoting the dissociation kinetics of H2O in alkaline HER. N doping also improves the electrical conductivity of carbon supports, synergistically enhancing the activity and stability of catalysts to achieve efficient catalysis under low Pt loading.
Sulfur (S) atoms are extensively utilized for doping carbon materials, leveraging their distinctive attributes: a large atomic radius, low electronegativity, abundant electrons, and the ability to form diverse covalent bonds. These characteristics could synergistically enhance the catalytic performance of resultant catalysts.94–96 The electronegativity difference between S and carbon endows S doping with unique electron-regulating ability, which can significantly alter the surface electronic structure of carbon materials.97 Additionally, S doping can enhance the SMSI between the support and metal, significantly accelerating charge transfer kinetics through interfacial electron reconstruction. This effect originates from the charge redistribution on the support surface induced by the electron-withdrawing property of S atoms: (1) optimizing the electron density of metal active sites; (2) shortening the charge migration path by enhancing interfacial coupling.98,99 Yan et al. reported metal size-dependent charge transfer of Pt single atoms and nanocluster catalysts based on mesoporous S-doped carbon (S–C) (Fig. 4f). Specifically, the S–C support captures electrons from Pt single atoms through strong chemical Pt–S interactions. However, when the Pt size increases to nanoclusters, the direction of electron transfer reverses, with electrons flowing from S–C to Pt.100 As a result, the PtSA/S–C catalyst with atomically dispersed Pt atoms exhibits relatively poor HER activity, requiring an overpotential of 53 mV at 10 mA cm−2. In contrast, the PtNC/S–C catalyst demonstrates significantly higher activity, achieving an ultra-low overpotential of only 11 mV at 10 mA cm−2. Moreover, Li et al. innovatively proposed a S doping strategy to achieve the synergistic regulation of confined encapsulation of Pt NPs and SMSI by constructing a nitrogen–S co-doped carbon-based carrier (NSC).101 On the one hand, the introduction of S atom doping provides more anchoring sites, enhances the coordination with the active component Pt, significantly improves Pt capture efficiency, and enables uniform deposition to form Pt NPs. On the other hand, the S atom doping activation strategy enhances the regulatory mechanism of the metal–support interaction and promotes charge redistribution at the interface simultaneously. This interfacial engineering not only precisely reconstructs the coordination microenvironment and electronic structure of Pt species but also constructs a directional charge transfer channel at the heterojunction interface (Fig. 4g), only requiring 17.8 mV to reach 10 mA cm−2 in 1.0 M KOH. Experimental results demonstrate that S-doped carbon materials can significantly enhance the SMSI between the support and metal components by regulating the surface electronic structure of carbon-based matrices, ultimately achieving leapfrog improvement in catalytic performance.
In addition to single-atom doping, many studies have shown that multi-atom doping can also enhance catalytic performance.103–105 For example, Zhao et al. anchored Pt atomic clusters on isolated Co atoms and N co-doped porous carbon (Pt–ACs/CoNC).102 The isolated CoN4 species provided sufficient anchoring sites to stabilize the Pt clusters. The strong interaction between CoN4 and Pt AC led to charge redistribution, realizing excellent stability (over 100 h of HER, Fig. 4h). Wang et al. proposed the synthesis of a Pt SA and NC coupled catalyst (Pt SAs–Pt NCs/NSC) anchored by N, S co-doped carbon (NSC) derived from coal pitch rich in vacancy defects by using the defect-induced electron redistribution effect.106 The co-doped S and N atoms in the carbon material can not only serve as anchoring sites to stabilize the Pt atoms but also regulate the electronic structure of the Pt atoms, promoting the HER process, and achieving an overpotential at 26 mV@10 mA cm−2. Carbon doping with multiple atoms or co-doping can modulate the electronic structure of carbon supports through synergistic effects, enhance the interfacial interaction with Pt, improve the electrical conductivity, hydrophilicity and mechanical strength of supports, and optimize reaction kinetics and durability. This provides an important direction for designing highly active alkaline HER catalysts with low Pt loading. Heteroatom doping proves to be an exceptionally potent strategy for the modification of carbon materials. Through the meticulous adjustment of their electronic structure and the deliberate induction of vacancy defects, this method generates highly favorable sites that enable the robust and efficient anchoring of co-catalysts. Consequently, it greatly bolsters the overall catalytic performance, thereby presenting substantial potential for diverse catalytic applications.
To conclude, for carbon-supported Pt-based electrocatalysts, carbon materials suffer from corrosion in alkaline media due to electrochemical oxidation, nucleophilic attack of OH−, galvanic corrosion, and dissolution–deposition effects. Therefore, a series of advanced carbon support engineering strategies, covering two major aspects of morphology engineering and surface engineering (defect engineering, surface functionalization, and heteroatom doping), have been applied to reduce carbon corrosion and further enhance the activity and stability of Pt-based catalysts. Next, we will elaborate on the fabrication strategies of carbon-supported Pt-based catalysts to further illustrate the interaction between carbon supports and Pt and the performance enhancement effects.
Lou's group reported the synthesis of Pt5/HMCS electrocatalyst via pyrolysis of a prefabricated Pt5(GS)10/HMCS (GS = deprotonated glutathione, HMCS = hollow mesoporous carbon spheres) composite in an Ar/H2 (95/5) atmosphere, which was prepared by sonicating Pt5(GS)10 clusters with HMCS in aqueous suspension at ambient temperature, followed by centrifugal collection and vacuum drying (Fig. 5a).112 During the annealing process, the GS ligands were removed, and the Pt clusters were confined in the pore channels of HMCS, maximizing the use of precious Pt atoms. As shown in Fig. 5b, the Pt atoms are evenly distributed on the porous carbon support, and the catalyst exhibits excellent alkaline HER activity, requiring 46.2 mV overpotential at a current density of 10 mA cm−2 (Fig. 5c). However, the stability of this catalyst in long-term alkaline electrolysis still requires systematic verification. Among them, whether the Pt clusters confined in the pores of HMCS will undergo agglomeration and dissolution due to the dissolution–deposition process caused by electrochemical corrosion or the size evolution of nanoclusters due to Ostwald ripening has not been sufficiently supported by experimental evidence. What is more, Zhang et al. used g-C3N4 as a sacrificial template and nitrogen-rich polymer to obtain nitrogen-doped graphene and then anchored the rich Co atomic sites on the nitrogen-doped graphene to prepare the CoN4–G support. Finally, by adsorbing [PtCl6]2− on the CoN4–G support and annealing at 400 °C in a 5% H2/Ar atmosphere, the Pt@CoN4–G catalyst was obtained (Fig. 5d).113 The prepared Pt@CoN4–G still maintains the wrinkled nanosheet morphology and abundant nanopores, which can provide a large surface area, abundant defective active sites, and mass transfer channels, thereby promoting the catalytic reaction. Fig. 5e verifies that the Pt clusters are surrounded by dense bright spots with two different contrasts; the brighter spots and darker spots can be identified as isolated Pt atoms and Co atoms, respectively, further proving that the Pt species are anchored on graphene. Nevertheless, characterization results show the coexistence of Pt single atoms and Pt clusters in the system, indicating that the adsorption amount of Pt lacks clear quantitative regulation, which reflects the randomness of the current synthesis scheme at the atomic scale. Benefiting from the SMSI between the Pt species and the CoN4–G support and the synergistic effect of multiple active sites, it manifests fine HER activity (η10 = 39 mV) and stability under alkaline electrolytes (Fig. 5f). Similarly, Guo's group anchored triphenylphosphine (PPh3)-coordinated Pt6 clusters to isolated Cr–N4 sites (Cr–N–C) on mesoporous carbon substrates through a simple wet impregnation process, followed by heat treatment to remove surface ligands to synthesize Pt–AC/Cr–N–C catalysts (Fig. 5g).114 Transmission electron microscopy (TEM) images of the prepared Pt–AC/Cr–N–C reveal that it is a 2D ultrathin sheet structure without visible metal NPs. Instead, most of the sub-nanometer (<1 nm) sized Pt atomic clusters are dispersed on the carbon substrate (Fig. 5h and inset). Combined with DFT, it was demonstrated that the unique Pt–Cr quasi-covalent interaction formed at the interface between the Cr–N4 site and the Pt ACs effectively suppressed the migration and thermal vibration of Pt atoms, stabilized the Pt atomic clusters, and exhibited a very low overpotential of only 19 mV to reach 10 mA cm−2 (Fig. 5i). In addition, an AEMWE using this catalyst (with only 50 μgPt cm−2) can operate stably at an industrial-level current density of 500 mA cm−2 at 1.8 V for >100 h with a small degradation rate of 90 μV h−1. The previous studies have offered a constructive approach for the precise design of single Pt-based robust electrocatalysts with high HER activity and durability. However, this approach, despite its simplicity and scalability, has a lengthy reaction time, which may result in the agglomeration of Pt NPs, damage to the carbon matrix's pore structure, and a decrease in surface area.
![]() | ||
Fig. 5 (a) Schematic illustration of the synthetic procedure of Pt5/HMCS. Step I: carbonization and etching of SiO2 to form HMCS. Step II: load the Pt5(GS)10 clusters into HMCS to form Pt5(GS)10/HMCS. Step III: conversion of Pt5(GS)10/HMCS to Pt5/HMCS electrocatalyst through thermal treatment. (b) HAADF-STEM image. (c) LSV curves of Pt5/HMCS-5.08% and commercial Pt/C (5 and 20 wt%) catalysts based on the geometric area of the working electrode in 1.0 M KOH. Reproduced with permission.112 Copyright 2019, Wiley-VCH GmbH. (d) Synthetic illustration of the synthetic process for Pt@CoN4–G. (e) Aberration-corrected HAADF-STEM image. Single Co atoms are marked by yellow circles, and single Pt atoms are marked by red circles. (f) LSV curves of different electrocatalysts in 1.0 M KOH. Reproduced with permission.113 Copyright 2023, Wiley-VCH GmbH. (g) Schematic illustration of the preparation of Pt–AC/Cr–N–C. (h) HAADF-STEM image. Inset, the size distribution of Pt ACs. (i) LSV curves tested in 1 M KOH solution. Reproduced with permission.114 Copyright 2023, American Chemical Society. |
Beyond the thermally induced reduction of precious metal Pt, the intrinsic defects of carbon carriers can also be harnessed for the reduction process. More precisely, the inherent defects within the carbon carrier, including vacancies and lattice distortion, serve as active sites. These sites enable the reduction of precious metal Pt from its compounds, such as oxides and salts, to the metallic state via chemical or physical interactions.115 This approach leverages the electronic properties of defects or the local chemical environment, eliminating the need for strong reducing agents, characterized by its eco-friendliness, high efficiency, and exceptional controllability.116 Nonetheless, it is constrained by the difficulty of precisely controlling defect density and distribution, and the reduction efficiency is limited by the electron-supplying capacity of the carrier's intrinsic defects.117
For instance, Luo et al. synthesized porous carbon supports (TC) featuring varying levels of inherent defects employing a straightforward template-assisted methodology. Subsequently, they deposited Pt onto these carbon supports to fabricate a range of Pt/TC catalytic materials.118 The findings demonstrate that the plentiful inherent defects (such as vacancies and topological defects) in the carbon supports are beneficial for the dispersion and anchoring of Pt species. Meanwhile, it also facilitates the electron transfer between Pt atoms and the carbon supports and modulates the electronic state of Pt species. These properties enhance the HER performance of the Pt/TC catalysts, which display a lower overpotential of 58 mV at a current density of 10 mA cm−2. DFT calculations verify that intrinsic defects, including vacancies and topological defects, play a crucial role in stabilizing Pt atoms. Among them, the Pt–C3 coordination is regarded as the most advantageous structure for enhancing the hydrogen evolution performance of Pt. This research offers fresh perspectives on the significant contribution of the inherent defects within the porous carbon supports to the HER performance of the Pt/C catalysts. Such insights are of profound importance for the future design and fabrication of advanced carbon-supported catalysts as well as other carbon-based electrode materials. Besides, Zhang et al. used C60 as an electron-accepting carrier to anchor Pt atoms and synthesized a highly loaded and highly dispersed Pt catalyst (Pt/C60) at room temperature.119 The resultant catalyst has a Pt loading as high as 21 wt%, predominantly comprising Pt SAs with a minor fraction of Pt clusters. This composition endows it with excellent HER catalytic activity, showing an overpotential of merely 25 mV at 10 mA cm−2 in 1 M KOH solution.
In the construction system of carbon-supported Pt-based catalysts, the carbon reduction method primarily achieves the controllable loading and stable anchoring of active Pt species through two pathways: pyrolysis-assisted reduction and defect-induced reduction. The pyrolysis-assisted reduction strategy leverages high-temperature pyrolysis processes to drive the highly uniform dispersion of Pt species in the form of single atoms or sub-nanometer clusters, effectively enhancing the atomic utilization efficiency of precious metal Pt. Nevertheless, this method relies on high-temperature energy consumption, which not only increases the energy cost during the preparation process but also makes the ligand removal step prone to causing the collapse or reconstruction of the pore structure of the carbon support. This, in turn, disrupts the continuity and effectiveness of the mass transfer channels, potentially negatively impacting the catalytic performance. In contrast, the defect-induced reduction pathway delves into the intrinsic electronic properties of the inherent structural defects (including lattice vacancies, topological distortions, etc.) of the carbon support. It realizes an in situ reduction process through charge transfer between the defect sites and Pt precursors, avoiding the use of external chemical reducing agents and conforming to the concept of green synthesis. In this process, the structural defects on the surface of the carbon support act as active anchoring sites, which effectively suppress the migration and agglomeration behaviors of Pt species through interfacial electron interactions, thereby achieving uniform dispersion of ultra-small Pt NPs on the support surface. Nevertheless, this strategy faces two major technical challenges: first, there is a lack of precise control methods for the density and spatial distribution of carbon support defects, making it difficult to achieve a quantitative match between defect sites and Pt loading; second, the intrinsic electron-donating ability of carbon materials becomes a key factor limiting the reduction efficiency, resulting in the inability to optimize the reduction degree and loading of Pt species. Therefore, innovative material design and synthesis process improvements are urgently needed to overcome the existing technical bottlenecks.
In the field of preparing carbon-supported Pt-based catalysts, the chemical reduction method is primarily realized through two pathways: the first involves introducing non-noble metals and leveraging the redox potential difference to promote the formation of Pt particles, ultimately constructing PtM alloy catalysts; the second employs strong reducing agents such as NaBH4 to achieve the reduction and preparation of Pt particles.
For example, Xiao et al. used the difference in redox potential between metals to direct the replacement of ions to prepare ultrafine Pt confined in N-doped porous carbon fibers (Pt@NDPCF).122 Specifically, imidazole ligands can coordinate with Zn ions to form ordered porous MOF materials, and then the 0D MOF particles are aggregated into 1D fibers by electrospinning technology, followed by carbonization to obtain highly confined Zn (0) in N-doped hierarchical porous carbon fibers (Zn@NDPCF). Due to the large difference between the redox potentials of Pt(II)/Pt (+1.2 V) and Zn(II)/Zn (−0.76 V), Zn (0) in Zn@NDPCF can be easily replaced by Pt at room temperature to form Pt@NDPCF. This synthetic strategy can form highly dispersed ultrafine Pt–NPs with strong directional interactions between Pt and pyridinic N, which significantly enhances the heavy d–π effect and greatly improves the HER activity (η10 = 36 mV) and durability. Although this system inhibits agglomeration through nitrogen-doped carbon carriers, it has not completely solved the problem of intrinsic thermodynamic instability. Similarly, Zhang et al. loaded ultrafine Pt–Co alloy NPs (designated Pt–Co@NCNT) on the inner and outer shells of porous nitrogen-doped carbon nanotubes via a pyrolysis-replacement-reorganization strategy (Fig. 6a).123 That is, the Co confined in the N-doped CNTs (Co@NCNTs) undergoes a galvanic replacement reaction with a K2PtCl4 solution, and then the PtCl42− is replaced by Co NPs and forms Pt–Co NPs with additional Co NPs, confined in the porous N-doped carbon nanotubes (denoted as pre-Pt–Co@NCNT, Fig. 6b). The Pt–Co@NCNT electrocatalyst obtained by further thermal annealing reorganization can drive a current density of 10 mA cm−2 at an overpotential of only 20 mV in alkaline media (Fig. 6c), exhibiting over 100 h HER stability. Xu et al. proposed a simple pyrolysis-confinement strategy to synthesize Con–Pt1@NPC catalysts, where Co nano-metal-particles (Con) and Pt SA (Pt1) were co-confined in N-doped lantern-shaped hollow polyhedral carbon materials (Fig. 6d).124 First, ZIF-8@ZIF-67 (Co/Zn zeolitic imidazolate framework) with a core–shell structure was pyrolyzed as a precursor to generate hollow polyhedral porous carbon materials (Con@NPC) loaded with Co NPs as a carrier. Then, Con@NPC was immersed in H2PtCl6 solution, and driven by the difference in reduction potentials of Co and Pt ions, the Pt ions were reduced. Fig. 6e confirms that Co NPs coexist with Pt atoms, and Pt mainly falls on the carbon surface without destroying the Co lattice. This strategy can regulate the metal content and size of the NPs and has excellent HER activity in 1 M KOH solution (η10 = 20 mV, Fig. 6f). Whether it is Zn substitution in Zn@NDPCF or Co reduction in Con@NPC, an excess of sacrificial metal is consumed, resulting in a loss of loading efficiency of the precious metal Pt.
![]() | ||
Fig. 6 (a) Illustration of the synthesis of Pt–Co@NCNT. (b) TEM image of a pre-Pt3Co@NCNT. (c) LSV curves tested in 1 M KOH solution. Reproduced with permission.123 Copyright 2021, Wiley-VCH GmbH. (d) Synthesis of Con–Pt1@NPC via a pyrolysis-confinement strategy. (e) HAADF-STEM image of Con–Pt1@NPC. (f) Polarization curves of Pt–Co@NCNT, Pt1@NPC, Con@NPC and 20%Pt/C. Reproduced with permission.124 Copyright 2025, Wiley-VCH GmbH. (g) Schematic of the synthetic route of PtC60. (h) TEM image of PtC60. (i) LSV of Pt catalysts with the loading of 0.4 mg cm−2 and bare C60. Reproduced with permission.125 Copyright 2023, Nature. |
Besides, Chen et al. first prepared buckminsterfullerene (C60) bulk crystals by a modified liquid–liquid interface precipitation method, and then added NaBH4 to reduce H2PtCl6 (Fig. 6g).125 The formed Pt clusters induced the one-step exfoliation of C60 crystals to prepare Pt nanocluster catalysts on C60 nanosheets (Pt/C60). During the exfoliation process, the reduction of H2PtCl6 precursor with NaBH4 can successfully surpass the interlayer interactions existing in C60 bulk crystals along the [110] axis, facilitating the achievement of the targeted PtC60 nanosheets (Fig. 6h). As a result, the lattice distance of the fullerene nanosheets is unusually large (∼0.8 nm), while the size of the Pt clusters is extremely small (∼2 nm), resulting in strong confinement of the Pt clusters and a remarkable charge redistribution at the tight Pt/fullerene interface, attaining the current densities of 10 mA cm−2, 50 mA cm−2, and 150 mA cm−2 at overpotentials of 24.3 mV, 53.2 mV, and 110.0 mV, respectively (Fig. 6i). These studies provide new inspiration in relation to the design of Pt-based nanomaterials boasting high performance.
Relying on the electron transfer properties of chemical reducing agents or the mechanism of intermetallic displacement reactions, the chemical reduction method exhibits remarkable advantages such as strong operational adaptability and a well-established process system, enabling the controllable preparation of nanoscale Pt particles. However, this method still faces two major technical bottlenecks in practical applications. On the one hand, the issues of reducing agent residue and nanoparticle aggregation urgently need to be addressed. The former may introduce impurities that affect catalytic performance, while the latter leads to insufficient exposure of active sites. On the other hand, whether by introducing other metal atoms to create displacement reaction conditions or directly adding chemical reducing agents, raw material loss is inevitable. Compared with pyrolysis processes, Pt NPs prepared by the chemical reduction method generally have larger average particle sizes, which directly reduces the surface utilization rate of Pt atoms and thus limits their efficient application in the field of catalysis. To sum up, reduction reactions using reducing agents provide important ideas and methodological guidance for the efficient regulation of the morphology and structure of catalysts. However, it should not be overlooked that there is still room for optimization of this technology. Future research can focus on three core directions: the development of green reducing agents, the efficiency of catalyst preparation processes, and the precise regulation of active sites.
Wang's group used cathodic electrochemical deposition to synthesize PtRu bimetallic clusters with adjacent SnO2−x on the CNT (PtRu/CNT@SnO2−x) catalyst (Fig. 7a).130 As shown in Fig. 7b, the PtRu alloy was deposited on the carbon nanotubes. They investigated the effect of SnO2 loading on the conductivity and metal site deposition of CNT@SnO2−x and concluded that CNTs and SnO2−x multi-sites synergistically promoted alkaline HER kinetics, exhibiting an overpotential of only 75 mV at 100 mA cm−2 in 1 M KOH (Fig. 7c). However, this method uses non-conductive carriers and requires pre-modification of conductive layers (such as SnO2−x), which increases the process steps. Moreover, when the PtRu alloy is electrolyzed for a long time at 1 M KOH, the Ru component may dissolve out, resulting in overpotential drift. Li et al. synthesized Pt-doped Co on carbon cloth (CC) by co-electrodeposition of CoPt alloy, and then successfully obtained Pt–Co2P particles (Pt–Co2P/CC) grown on CC by phosphating (Fig. 7d).131 Scanning electron microscopy (SEM) confirmed that Pt–Co2P particles with a size distribution between 125 nm and 150 nm were uniformly grown on CC (Fig. 7e). By internally modifying Co2P via electrodeposition of Pt, the local chemical state and electronic structure of the catalyst can be tuned to enhance electron mobility and reduce the energy barriers for hydrogen adsorption and H–H bond formation, significantly improving the HER activity under alkaline conditions, with a much lower overpotential of 2 mV to achieve 10 mA cm−2 (Fig. 7f), which is superior to the performance of a great majority of current alkaline HER catalysts. But it cannot be ignored that during the co-electrodeposition of CoPt alloys, the excess Co2+ residue may form hydroxide precipitates, blocking the active sites.
![]() | ||
Fig. 7 (a) Schematic diagram of the preparation process of the M/CNT@SnO2−x catalyst. (b) HRTEM image of PtRu/CNT@SnO2−x. The yellow marked area belongs to the SnO2−x nanoparticle, and the pink area represents the electrodeposited Pt and Ru. (c) HER polarization curves for commercial Pt/C, Pt/CNT@SnO2−x–K2PtCl4, Pt/CNT@SnO2−x–PtCl4, and PtRu/CNT@SnO2−x in 1 M KOH solution at a scan rate of 10 mV s−1. Reproduced with permission.130 Copyright 2024, American Chemical Society. (d) Schematic illustration of the fabrication process of Pt–Co2P. (e) SEM image of Pt–Co2P particles grown on CC. The inset is the particle size distribution. Scale bar: 2 μm. (f) LSV curves of the catalysts tested at 5 mV s−1. Reproduced with permission.131 Copyright 2020, The Royal Society of Chemistry. (g) Schematic illustration of the preparation of Pt@DG. (h) HAADF-STEM image of Pt@DG. (i) Polarization curves of Pt@G, Pt@DG, and Pt/C with a scan rate of 10 mV s−1. Reproduced with permission.132 Copyright 2022, American Chemical Society. |
Similarly, photoreduction, as a green synthesis strategy, involves a metal precursor reduction process driven by photoinduced electron transfer, which mainly includes three kinetic regulation stages: controllable nucleation, directional growth, and final conversion into stable metal NPs.111 This method precisely anchors highly dispersed metal NPs on the surface of carbon carriers through photochemical reduction. Moreover, by reducing the reaction nucleation rate, uniformly dispersed metal clusters or single atoms can be prepared. At the same time, the use of low-temperature light sources such as ultraviolet rays can effectively prevent the aggregation of metal atoms.133 For example, Yao's group prepared Pt@DG catalysts with rich Pt–C3 coordination structures by photoreducing H2PtCl6 on defective graphene (DG).132 As demonstrated in Fig. 7g, the mixture of melamine and graphene was first pyrolyzed at 700 °C for 2 h to obtain N-doped graphene (NG), and DG was formed by annealing NG at 1150 °C to completely remove N species. Finally, Pt4+ was uniformly dispersed on DG and then photoreduced to obtain Pt@DG. Many bright dots are observed to distribute evenly on the graphene surface from high-magnification HAADF-STEM (Fig. 7h), indicating the successful formation of atomic Pt. Experimental and theoretical results confirm that the optimal Pt–C3 coordination structure has a stronger electron capture ability and lower ΔG, which promotes the reduction of adsorbed H+ and the desorption rate of H2, thus exhibiting excellent HER activity (Fig. 7i). Furthermore, Ji et al. used photoreduction to synthesize both Pt single atom/cluster anchored functional multi-walled CNTs (Pt/f-MWCNTs), which exhibit the overpotential of 43.9 mV at a current density of 10 mA cm−2.134 Photoreduction technology is limited by the penetration depth of ultraviolet light less than 100 μm and is only suitable for thin-layer materials such as defective graphene. It is prone to cause uneven reduction on thick substrates such as bulk carbon carriers.
Electroreduction and photoreduction technologies have provided unique technical pathways and innovative strategies for the controllable synthesis and performance optimization of catalysts. By driving electron transfer through electrical or light energy, precise regulation of material structures and active sites can be achieved. Although the above technologies have demonstrated remarkable advantages in basic research, their industrial applications still face multiple bottlenecks. The electroreduction process is constrained by its high energy consumption, and electrode materials are prone to structural collapse and active component loss during long-term cycling, leading to catalytic performance degradation. Additionally, in the preparation of large-area catalysts, the non-uniform distribution of electric fields inevitably causes spatial differences in active site distribution, affecting the overall performance consistency of the catalytic system. The practical application of photoreduction technology is limited by the theoretical upper limit of light absorption efficiency: existing photocatalysts have a narrow response range to the solar spectrum, and the rapid recombination of photogenerated carriers seriously reduces energy conversion efficiency. Meanwhile, the complex photophysical and photochemical reaction processes in photoreduction systems, as well as the interaction mechanisms between photocatalysts and reaction media, further increase the difficulty of regulating the reaction process. For future development, there is an urgent need to develop novel catalytic systems driven by the synergistic effect of electrical and light energy. By constructing a photoelectro-coupled reaction interface, the complementary advantages of the two energy forms can be realized. This strategy is expected not only to break through the performance bottlenecks of single-energy input modes but also to provide new ideas for catalyst structure design and performance regulation, thereby promoting the large-scale application of electroreduction and photoreduction technologies in fields such as energy catalysis and environmental remediation.
To sum up, the interaction between carbon supports and Pt significantly enhances the activity and stability of carbon-supported Pt-based catalysts for alkaline HER, as demonstrated in three aspects: (1) the specific surface area, porous architecture, and defect sites of carbon supports provide dispersion sites for Pt, effectively inhibiting agglomeration. (2) Surface functionalization and heteroatom doping (e.g., N/P/S) modulate the electronic structure of supports, strengthening interfacial coupling with Pt and optimizing H* adsorption kinetics as well as H2O dissociation dynamics. (3) The mechanical and chemical stability of 3D carbon materials suppresses Pt dissolution and agglomeration in alkaline media, extending catalyst lifespan. Such interactions synergistically boost catalytic activity through structural matching, electronic modulation, and interfacial reinforcement. Meanwhile, heteroatom doping and 3D structural design enhance stability via chemical bonding and skeletal support, paving the way for the development of low-Pt-loading catalysts.
Footnote |
† These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |