Sowjanya Vallema,
Malayil Gopalan Sibib,
Rahul Patilb,
Vishakha Goyalbi,
A. Giridhar Babuc,
EA. Lohithd,
K. Keerthid,
Muhammad Umere,
N. V. V. Jyothid,
Matthias Vandichel
e,
Daniel Ioan Stroe
a,
Subhasmita Rayf,
Mani Balamurugan
*gb,
Aristides Bakandritsos
*bh,
Sada Venkateswarlu
*b,
Rajenahally V. Jagadeesh
*bi and
Radek Zboril
*bh
aDepartment of Energy, Aalborg University, Aalborg, 9220, Denmark
bNanotechnology Centre, Centre for Energy and Environmental Technologies, VSB-Technical University of Ostrava, 17. listopadu 2172/15, 708 00, Ostrava-Poruba, Czech Republic. E-mail: balubdu@gmail.com; a.bakandritsos@upol.cz; venkisada67@gmail.com; jagadeesh.rajenahally@catalysis.de; radek.zboril@upol.cz
cDepartment of Basic Sciences, SR University, Warangal 506371, Telangana, India
dDepartment of Chemistry, Sri Venkateswara University, Tirupati 517502, Andhra Pradesh, India
eSchool of Chemical Sciences and Chemical Engineering, Bernal Institute, University of Limerick, Limerick V94 T9PX, Republic of Ireland
fDepartment of Condensed Matter Physics, Faculty of Mathematics and Physics, Charles University, Ke Karlovu 3, Prague 12116, Czech Republic
gSOFT Foundry Institute, Seoul National University (SNU), Seoul 08826, Republic of Korea
hRegional Centre of Advanced Technologies and Materials, Czech Advanced Technology and Research Institute, Palacky University, Slechtitelu 27, 77900 Olomouc, Czech Republic
iLeibniz-Institut fur Katalyse e.V., Albert-Einstein-Str. 29a, Rostock, D-18059, Germany
First published on 20th August 2025
Amidst escalating global concerns over rising atmospheric CO2 levels, the capture and effective utilization of C1 and C2+ sources are crucial not only for advancing a sustainable society but also for economically viable chemical synthesis. CO2 valorisation as a chemical feedstock has garnered significant attention, driving the development of diverse catalytic systems and reaction pathways. Among them, single-atom catalysts (SACs) have emerged as a transformative class of materials owing to their maximized atom efficiency, well-defined active sites, and tunable electronic structures, enabling high catalytic selectivity and activity. When hosted on covalent-organic frameworks (COFs) and metal–organic frameworks (MOFs), SACs benefit from the structural regularity, high surface area, and chemical modularity of these porous crystalline scaffolds, further enhancing their catalytic performance and stability. This review provides an in-depth discussion of COF and MOF derived SACs for CO2 valorisation through electrochemical, photochemical, and thermochemical approaches. We have explored the key factors that influence the performance of the CO2 reduction reaction (CO2RR) to enhance both selectivity and efficiency. In addition to catalyst preparation and synthetic applications, we provide an in-depth analysis of the mechanistic aspects and theoretical simulations of the COF and MOF SACs based CO2 utilization. We explore the role of machine learning models in advancing SACs based CO2 valorisation. We also identify key challenges including SAC agglomeration, mechanistic ambiguity, selectivity control and limited long term operational stability, while discussing future perspectives (such as electronic structure tuning, multi atom site design, and machine learning-assisted catalyst discovery) in this field of broad scientific, technological and societal interest.
Broader contextAddressing the urgent global challenge of reducing atmospheric CO2 levels and mitigating climate change is an utmost priority. CO2 is also increasingly recognized as an abundant and renewable carbon feedstock that can be valorized into value-added chemicals and fuels. Efficient and selective conversion of CO2, requires the development of advanced catalytic systems. Among these, single-atom catalysts (SACs) have emerged as highly promising candidates due to their maximal atom utilization and exceptional tunability in activity, selectivity, stability, and reusability. The rational design of SACs relies on the appropriate choice and integration of precursors and supports. In this regard, covalent organic frameworks (COFs) and metal–organic frameworks (MOFs) are particularly attractive as both precursor materials and supporting architectures. These porous crystalline materials offer high surface areas, well-defined pore structures, and the capacity to anchor metal species at the atomic scale—features essential for optimal catalytic performance. SACs derived from COFs and MOFs exhibit great promise for the thermochemical, photocatalytic, and electrocatalytic reduction of CO2 into valuable C1 and C2+ products. Moreover, the integration of density functional theory (DFT) and machine learning (ML) approaches can provide mechanistic insights and accelerate the discovery of efficient catalysts. |
![]() | ||
Fig. 1 (a) and (b) Pie chart represents globally, each country produces different levels of CO2 emissions, and the top contributions from 2020 to 2023 (CO2 emission in 2023, International Energy Agency).5 |
Recent advancements in electrochemical conversion, photochemical catalytic reduction, and thermocatalysis have shown promise in converting CO2 into various chemicals and fuels.10–13 Each method operates under specific reaction conditions, with heterogenous catalysts playing a significant role in the CO2 reduction reaction (CO2RR), facilitating the production of C1 and C2+ hydrocarbons.14–17 In thermocatalysis, CO2 hydrogenation requires significant energy to break its strong CO bonds, necessitating high temperatures (200 °C to 500 °C) and pressure (10 to 100 bar) for CO2 valorisation. To enhance reaction rates and selectivity, researchers have developed low-energy barrier nanoparticle catalysts, particularly those based on noble metals.18,19 The electrochemical CO2RR initially utilized amalgamated metal compounds. Over time, metallic catalysts (Mo, Ru, Cu) and semiconductor nanoparticles (e.g., p-InS, n-/p-Ga–As) were developed for CO2 to methanol conversion, typically occurring at the cathode of an electrochemical cell with the appropriate electrolytes.20–23 However, a major challenge in CO2RR is the competing hydrogen evolution reaction, which reduces the efficiency and product purity. Another promising approach for CO2 conversion into fine chemicals and renewable fuels is photochemical catalysis, which harnesses solar energy with zero emissions. In photocatalysis, the surface charge state of the catalyst is crucial for ensuring efficient CO2 conversion. Researchers have developed various materials including Lanthanides, MXenes, MoS2, and noble metals (Ru, Rh, Pt) as well as bimetallic and core–shell-based Schottky heterojunctions, where co-catalysts play a key role in accelerating charge transfer and reducing overpotential.24–29 However, challenges such as efficiency, selectivity, and long-term stability remain major bottlenecks in nanoparticle-based photochemical CO2RR.30 To address these limitations across all conversion methods, researchers have recently developed a unique class of metal catalysts known as single-atom catalysts (SACs).31 Metal single atom (SA) based catalysts are highly desirable because these frontier catalytic materials (SACs) exhibit exceptional atomic efficiency, tunable activity, selectivity, stability, reusability, and quantum size effects.32–38 Moreover, these advanced materials bridge the gap between homogeneous and heterogeneous catalysis, making them highly preferable for challenging synthetic reactions.39–42 The design of potential SACs crucially depends on the choice of precursors, substrate supports, and reaction conditions. Among these, COF and MOF serve as excellent precursors and supports for the preparation of desired SAs.43–47 The timeline of development on COF and MOF based SACs, along with the number of publications on CO2 conversion, is shown in Fig. 2a and b.48–51
![]() | ||
Fig. 2 (a) Timeline showing the significant milestones of MOF/COF–SACs in valorisation of CO2 till date. (b) Number publications on MOF/COF–SACs in valorisation of CO2 (from web of science). |
Notably, COF and MOF based or derived catalysts offer exceptional specific surface area, unique morphology, mesoporous characteristics, and atomic dispersion of metal species, which are key parameters for SACs to function effectively.52–59 Compared to traditional nanoparticle-based catalysts, SACs derived from MOFs, which feature unique active sites and adaptable topologies, are emerging as promising materials for CO2 valorisation.60–63 Recently, MOF-derived/anchored SACs with specific site coordination arrangements have been designed for CO2 conversion into targeted products by tailoring MOF precursors and fabrication methods. Additionally, COFs possess huge porosity, and multiple heteroatom functional groups, making them an ideal substrate for SACs. The bottom-up synthesis of COFs using various building blocks allows for the customization of skeleton structures, pore sizes, and topological arrangements.64,65
Several reviews have been published on CO2 reduction reactions,66–69 primarily focusing on MOF or COF based SACs for electrochemical reduction of CO2. However, in this work we attempt to meet the urgent need for a comprehensive review CO2 valorisation using MOF and COF derived SACs through miscellaneous methods, such as thermochemical, photochemical, and electrochemical approaches (Scheme 1). In this review, we provide an in-depth analysis of this topic, covering not only the design and applications of MOF and COF SACs for CO2 valorisation, but also spectroscopic analyses (ex situ/in situ/operando), density functional theory (DFT) mechanistic studies, and machine learning (ML) models aspects. Additionally, we discuss the limitations and future perspectives on SACs and CO2 utilization.
![]() | ||
Scheme 1 Schematic overall representation of COF and MOF single atom catalysts for CO2 valorisation, along with their characterization and theoretical studies. |
Optimizing synthesis routes and atomic configuration is vital for enhancing the catalytic activity of SACs in CO2 conversion.75,76 Synthesis methods regulate atomic dispersion, coordination environments, and electronic structures, while atomic engineering strategies such as defect creation, coordination tuning, and ligand interactions further refine catalytic performance.70,77 These factors collectively influence reaction pathways and product selectivity, allowing precise control over catalyst efficiency in CO2 conversion.
Tailoring synthesis strategies is essential for optimizing SACs for targeted CO2 valorisation pathways, including electrocatalytic, photocatalytic, and thermocatalysis transformations. The integration of precise synthesis and atomic engineering enables fine tuning of active sites, electronic states, and metal–support interactions, which directly influence reaction kinetics and selectivity. For instance, defect rich MOF derived SACs enhance CO2 adsorption and activation, while ligand-modified COFs provide tunable coordination environments that steer product selectivity.78,79 The rational design of SACs through an integrated approach of controlled synthesis and atomic engineering plays a pivotal role in improving catalytic efficiency, stability, and economic feasibility in CO2 valorisation technologies. To achieve these characteristics, researchers have explored diverse synthesis strategies. This section primarily covers, pyrolysis, pre-synthetic metalation, and post-synthetic metalation which are mostly used for CO2RR catalyst design.
![]() | ||
Fig. 3 Schematic illustration of SAC synthesis strategies: (a) direct pyrolysis of COFs. Reproduced from ref. 80 with permission from Wiley (CC-BY 4.0), copyright 2022. (b) Sacrificial template-assisted pyrolysis of COFs. Reproduced from ref. 83 with permission from Wiley (CC-BY 4.0), copyright 2023. (c) Direct pyrolysis of MOFs. Reproduced from ref. 85 with permission from Wiley (CC-BY 4.0), copyright 2022. (d) Sacrificial template-assisted pyrolysis of MOFs for CO2RR. Reproduced from ref. 89 with permission from Wiley, copyright 2020. |
Whereas as, confined pyrolysis methods employ nanoconfinement to limit atomic migration and aggregation of metal atoms, thereby stabilizing atomically dispersed SACs. Confined pyrolysis is also a strategy to regulate the structural transformation of MOFs into porous carbon materials while preserving their framework integrity and optimizing catalytic properties. Unlike conventional pyrolysis, which often leads to collapse or shrinkage, confined pyrolysis employs hard templates such as silica-based or polymeric materials to maintain structural integrity, prevent metal nanoparticle agglomeration, and facilitate the formation of a robust porous network.90 A key example is the synthesis of Zn–N–HOPCPs, where ZIF-8 single crystals are embedded within a SiO2 template before pyrolysis, yielding a 3D-ordered macroporous carbon structure (Fig. 4a).91 This hierarchical architecture enhances molecular diffusion, maximizes active site exposure, and significantly improves CO2 conversion efficiency, particularly in cycloaddition with epoxides. Notably, the Zn–N–HOPCP catalyst, synthesized via confined pyrolysis, demonstrated outstanding performance in CO2 photoconversion to cyclic carbonates under mild conditions. Owing to its broad light absorption spectrum (230–800 nm), this catalyst achieved an impressive 95% catalytic yield at room temperature and low pressure when coupled with TBAB as a co-catalyst. The high surface area (1128 m2 g−1) and well dispersed single Zn sites played a crucial role in enhancing CO2 activation. Consequently, Zn–N–HOPCPs exhibited superior activity and stability, underscoring the effectiveness of confined pyrolysis in producing highly efficient catalysts for CO2 utilization in value-added chemical synthesis.
![]() | ||
Fig. 4 Representative examples of (a) the confinement pyrolysis approach for MOF derived SACs, highlighting the controlled conversion of MOFs into atomically dispersed catalysts. Reproduced from ref. 91 with permission from Elsevier, copyright 2020. (b) Multi atomic sites derived from MOF via pyrolysis. Reproduced from ref. 92 with permission from Elsevier, copyright 2024. |
Recently, MOF-derived dual atom catalysts (DACs) have demonstrated enhanced catalytic efficiency due to synergistic contacts between homo or hetero metal sites. Expanding on this concept, pyrolysis-derived multi-atom catalysts introduce an additional level of tunability by incorporating a third metal atom, which strengthens electronic interactions and optimizes intermediate adsorption and reaction kinetics. A notable example of this approach is the synthesis of Fe–Co–Ni@NC trimetallic catalysts from MOF precursors (Fig. 4b).92 This strategy ensures atomic-level dispersion of Fe, Co, and Ni on a porous carbon framework without requiring acid treatment. The synthesis of Fe–Co–Ni TACs anchored on hollow carbon structures follows a systematic multi-step approach to achieve atomic dispersion and enhanced catalytic performance. Initially, Fe(acac)3 is incorporated in situ during the self-assembly of ZIF-8, leveraging its molecular compatibility to be efficiently confined within the MOF cages. The precursor-loaded MOF undergoes carbonization, forming an Fe–N-doped porous carbon (Fe@NC) structure. The Fe–NC nanoparticles are then dispersed in a Tris buffer solution and coated with polydopamine through dopamine self-polymerization. The polydopamine layer, rich in hydroxyl and nitrogen functional groups, facilitates the coordination of Co and Ni cations via covalent bonding or electrostatic attraction. Subsequent annealing at 900 °C under a nitrogen atmosphere results in the formation of Fe–Co–Ni@NC, where atomically dispersed Fe, Co, and Ni sites are stabilized within a hierarchically porous carbon framework, optimizing catalytic activity for ORR/OER. This process effectively prevents metal aggregation during pyrolysis, ensuring homogeneous atomic dispersion. The resulting Fe–Co–Ni@NC structure provides well-defined Fe–N4, Co–N4, and Ni–N4 coordination sites, leading to superior electrocatalytic performance in ORR and OER. The synergy among the three metal centers augments the adsorption and activation of key intermediates, significantly enhancing catalytic efficiency compared to SACs and DACs. Although this system was designed for ORR/OER, its underlying synthesis strategy, and electronic tuning principles are highly relevant for CO2RR catalyst development. Given the pivotal role of nickel and its heteroatomic coordination in modulating CO2RR activity and selectivity, similar MOF derived pyrolysis strategies could enable the rational design of high-performance multi-atom catalysts for CO2RR.93–95 Future studies could explore tailored trimetallic configurations to further enhance C–C coupling, CO2 activation, and long-term catalytic stability.
Although pyrolysis-derived SACs enhance conductivity and stability by forming confined sites such as M–N–C, M–N4, and M–N4–O in carbon frameworks, alternative synthesis strategies enable atomic-scale metal incorporation while preserving the porous structure integrity of COFs and MOFs. Among these, pre-synthetic metalation integrates metal centers directly into the framework during assembly, ensuring precise atomic coordination and uniform dispersion. As discussed in the following sections, this approach provides significant advantages in maintaining framework stability and tailoring the metal coordination environment.
A notable example of this approach is the synthesis of CoPc-DSDS-COF and CoPc-DNDS-COF, where Co-hexadecafluorophthalocyanine (CoPc) serves as the metal node and organic linkers such as 1,2,4,5-benzenetetrathiol (DSDS) or 2,5-diaminobenzene-1,4-dithiol (DNDS) create an extended conductive network (Fig. 5a).96 By integrating metallophthalocyanines as an intrinsic part of the COF backbone, the resulting materials exhibit a well-defined atomic Co–N4 coordination environment without requiring post-synthetic modifications or high-temperature pyrolysis. This controlled metalation ensures that the Co centers are uniformly distributed throughout the framework, preventing aggregation and maximizing active site exposure. The pre-synthetic incorporation of CoPc into the COF structure enhanced its electrical conductivity and charge transport, which is crucial for efficient electrocatalysis. The electronic band structure of these COFs enabled favourable charge transfer and optimal binding of CO2 reduction intermediates, ultimately improving catalytic activity and selectivity. Experimentally, CoPc-DNDS-COF demonstrates superior CO2-to-CH4 conversion, whereas CoPc-DSDS-COF shows selective CO production, highlighting the tunability of these materials based on linker modifications. This pre-synthetic metalation approach thus offers significant advantages over post-synthetic methods by preserving framework integrity, ensuring precise metal coordination, and enabling systematic structure–property optimization for CO2RR.
![]() | ||
Fig. 5 Schematic of SAC synthesis strategies: (a) post-synthetic metallization for COF based SACs. Reproduced from ref. 96 with permission from American Chemical Society, copyright 2025. (b) Pre-synthetic electrodeposition for MOF based SACs. Reproduced from ref. 97 with permission from Royal Society of Chemistry, copyright 2024. (c) Surfactant-assisted MOF synthesis for atomic dispersion. Reproduced from ref. 98 with permission from Springer, copyright 2023. |
![]() | ||
Fig. 6 Schematic representation of (a) the post-synthesis incorporation of SACs into COFs through electrodeposition. Reproduced from ref. 99 with permission from Nature, copyright 2024. (b) The post-synthesis SAC reduction on MOF through photo irradiation. Reproduced from ref. 104 with permission from American Chemical Society, copyright 2020. (c) Illustrative example to show SACs implantation on post-synthesis MOF using ALD. Reproduced from ref. 109 with permission from American Chemical Society, copyright 2016. |
Among these strategies, photo-assisted methods offer a distinct advantage by enabling mild and controlled synthesis conditions while maintaining the porous framework of MOFs. A representative example is the photoinduction method used to construct Cu SAs on UiO-66-NH2, a MOF recognized for its superior CO2 adsorption capacity (Fig. 6b).104 In this approach, copper nitrate pentahydrate is introduced into a suspension of pre-synthesized UiO-66-NH2, where the –NH2 groups serve as coordination sites, capturing Cu precursors and ensuring their spatial confinement at the atomic level. Unlike conventional high-temperature pyrolysis methods that risk framework degradation and metal aggregation, visible-light irradiation is employed to anchor and stabilize Cu SAs, preserving the MOF structure while facilitating strong Cu–N coordination within the framework. This mild and controllable synthetic approach ensures better atomic dispersion, thereby optimizing the electronic environment of Cu SACs. The photoinduced interaction between Cu and N leads to a two-coordinate planar geometry, effectively stabilizing the Cu SAs. Moreover, this process modulates the electronic properties of the material, reducing the bandgap and enhancing charge separation efficiency, which is beneficial for CO2RR. The optimized Cu SAs/UiO-66-NH2 catalyst exhibits improved CO2 activation, endorsing the formation of COOH* intermediates and selectively converting CO2 into ethanol and methanol.
The photoinduction method provides a precise and energy-efficient approach for SAC fabrication by employing post-synthetic metalation combined with ligand-assisted stabilization. Its ability to maintain structural integrity while optimizing active sites makes it a promising strategy for developing high-performance MOF-SACs for CO2RR. This approach exemplifies how tailored post-synthetic methodologies expand the design possibilities of MOF based catalysts, paving the roadmap for more efficient and sustainable CO2 valorisation technologies.
A facile gas-migration strategy has been developed to transform bulk metals directly into single atoms on a support, offering great potential for large scale production of single-atom catalysts (SACs).110 In this approach, copper foam and ZIF-8 precursors are placed separately in a porcelain tube furnace. The ZIF-8 is pyrolyzed at 900 °C under an inert argon atmosphere, while ammonia gas is simultaneously introduced through the system via the copper foil. Under these conditions, the strong Lewis acid base interactions between Cu atoms and NH3 molecules lead to the formation of volatile Cu(NH3)x species, which are ejected from the copper foam. These volatile Cu species are subsequently trapped by the defect-rich nitrogen moieties in the carbon support, forming atomically dispersed Cu single atoms on N-doped carbon (Cu-SA/N–C). This gas migration strategy has also been successfully applied to graphene oxide materials possessing abundant defect sites. The key principle of this method lies in the role of ammonia as a carrier gas that extracts Cu atoms from the bulk metal and facilitates their chelation with nitrogen sites on the porous carbon. These nitrogen sites are generated in situ by the volatilization of Zn from ZIF-8 at high temperatures, leaving behind anchoring nodes for single-atom coordination. This innovative strategy opens new avenues for producing SACs directly from bulk metals, significantly advancing the scalability and industrial applicability of single-atom catalysts. Furthermore, the design methods, associated challenges, and future prospects of SACs are summarized in Table 1.
Method | Challenges | Future prospects |
---|---|---|
Direct pyrolysis | Metal aggregation at high temperatures | Improved precursor design with atom-trapping ligands |
Limited atomic control | Use of additives to prevent sintering | |
Poor dispersion at high loadings | ||
Sacrificial and confined pyrolysis | Complex synthesis of templates | Scalable low-cost MOF/COF alternatives |
Scalability limited by precursor cost | Dual-template or hard–soft template combinations. Combination with 2D materials for enhanced exposure | |
Multi-step fabrication. Restricted diffusion of reactants during catalysis | ||
Pre-synthetic metalation | Low thermal stability of frameworks | Introduction of new and potential linker and MOF/COF design for enhanced thermal and chemical robustness |
Limited to compatible metal–ligand systems | Controlled dual-metal SACs | |
Post-synthetic metalation | Weak coordination may lead to leaching | Creating suitable coordination environment of the metal centre with heteroatoms (for example N or P) |
Incomplete metal incorporation | Application in liquid-phase catalysis | |
ALD (vapor phase) | High cost and low throughput | Integration with batch ALD systems |
Requires specialized equipment | High-value device-level SAC design (e.g., sensors, fuel cells) | |
Not easily scalable | ||
CVD (vapor phase) | Metal aggregation at high temperature | Tunable gas-phase synthesis for batch-scale SACs |
Complex precursor control | Strong metal–support interaction engineering | |
Limited substrate versatility | ||
Molten salt pyrolysis | Post-synthesis salt removal | Green salt systems (eutectics) |
Corrosive environment | High-throughput and template-free SAC synthesis | |
Limited support compatibility | ||
Gas Migration strategy | Reactor design complexity | Scalable production using bulk metals |
Control over metal migration path and trapping | Broader application to various defect-engineered supports | |
Ammonia safety concerns |
Building upon the discussion of synthesis methods, the strategic design of SACs and dual-atom catalysts (DACs) has significantly advanced CO2RR by optimizing active site utilization and enhancing selectivity.111–113 However, further improvements in catalytic performance require moving beyond a single variety of isolated metal centers. Multi-atom catalysts offer an advanced approach by incorporating atomically dispersed multimetallic sites or well-defined clusters, fostering synergistic interactions that can modulate electronic structures and reaction kinetics.114 A carefully selected combination of SACs can effectively address the inherent limitations of individual catalysts. For example, Co and Fe-based SACs exhibit strong CO binding, which hampers CO desorption, whereas Ni-based SACs facilitate CO release but encounter high energy barriers for COOH formation.75,115,116 To mitigate these challenges, DA and MA catalysts have been explored, leveraging cooperative effects between metal centers to enhance catalytic activity and efficiency. Although direct reports on multi-atom catalysts for CO2RR remain scarce, studies in other catalytic transformations suggest that controlled nucleation and confinement strategies, such as those employed in MOFs and COFs, could be adapted for their rational design in CO2RR.93,117 Integrating insights from SACs and DACs, future research could strategically engineer multi-metallic atomic active centers within porous frameworks to achieve optimal stability, electronic properties, and reactivity, further advancing the prospects of electrochemical CO2 conversion.
![]() | ||
Fig. 7 (a) Identification of SACs using AC-STEM (inset: schematic of SAC structure). (b) HAADF image of dual atom catalysts, (c) EELS mapping distinguishing Ni and In atoms, (d) corresponding EELS spectra. Reproduced from ref. 124 with permission from Wiley, copyright 2023. (e) HRTEM image of Fe–N4 site (inset: schematic of Fe–N4), (f) LT-STM image of Fe–N4, (g) dI/dV spectra along with line section shown in the inset. Reproduced from ref. 125 with permission from Science (CC-BY 4.0), copyright 2015. (h) EXAFS spectrum of Mo-COF, and (i) Fourier-transformed Mo K-edge EXAFS of Mo-COF compared to Mo foil and MoO3. Reproduced from ref. 128 with permission from Elsevier, copyright 2021. |
Spectroscopic techniques play a vital role in understanding the electronic structure and chemical interactions of SACs. X-ray photoelectron spectroscopy is commonly used to analyze surface chemical states and elemental composition.126 However, for SACs, X-ray absorption spectroscopy (XAS) has emerged as a more powerful tool due to its ability to probe both the electronic configuration and local coordination environment at subatomic resolution. This makes XAS particularly advantageous in unraveling the structural and chemical identity of isolated metal sites, even in complex matrices like MOFs, COFs, or carbon supports. XAS, comprising XANES and extended EXAFS, provides subatomic resolution of electronic configurations. While XANES helps in identifying oxidation states and coordination geometries, EXAFS gives details on bond distances and coordination numbers.127 For instance, in Mo-COF, XAS provided clear evidence confirming the atomic dispersion of Mo sites (Fig. 7h and i).128 The Mo K-edge XANES spectrum showed that the absorption edge of Mo–COF closely matched MoO3, indicating a +6 oxidation state (Fig. 7h). A slight shift in the pre-edge region indicated Mo coordination with nitrogen, rather than oxygen. Further, FT-EXAFS as shown in Fig. 7i revealed a primary peak at ∼1.28 Å (Mo–N/Mo–C bond), with no detectable Mo–Mo peak, directly confirming isolated Mo atoms without clustering. Wavelet transform (WT) analysis supported this finding, showing a scattering signal at ∼5.15 Å−1 (Mo–N/Mo–C), unlike Mo foil (Mo–Mo) or MoO3 (Mo–O). This combined spectral evidence underscores the strength of XAS in identifying and characterizing SACs in COFs.
![]() | ||
Fig. 8 (a) In situ TEM images illustrating temperature and time-dependent molecular transformations during the formation of MOF-derived SACs. Reproduced from ref. 132 with permission from Wiley, copyright 2022. (b) In situ XAS analysis of Cu–Fe SACs under operando conditions during CO2RR. (c) In situ Raman spectroscopy revealing the mechanistic evolution of the SAC environment during CO2RR. Reproduced from ref. 134 with permission from Nature (CC-BY 4.0), copyright 2022. (d) In situ FT-IR spectroscopy demonstrating the real-time mechanistic interplay within the SAC environment during CO2RR. Reproduced from ref. 128 with permission from Elsevier, copyright 2021. |
A combination of ex situ and in situ techniques is essential for characterizing MOF and COF derived SACs for CO2RR applications. Electron microscopy techniques provide atomic-level structural insights, while spectroscopic methods reveal electronic and chemical properties.137 In situ techniques further enhance understanding by tracking dynamic changes during catalytic reactions, offering valuable insights for designing efficient CO2RR catalysts.138 Future studies should focus on integrating multiple characterization techniques to leverage their complementarities and to achieve a more comprehensive understanding of SAC functionalities.139
Initially, MOFs were used to produce highly conductive N-doped carbon materials through high-temperature annealing, as MOF-derived carbon exhibits greater efficiency than the original MOF under harsh conditions. This approach also helps mitigate disadvantages related to stability and reactivity in such an environment.144 However, the presence of metal ions is more important for enhancing the catalytic rate. Since MOFs contain abundant N-ligands, they transform into N-doped carbon during pyrolysis, where the N-ligands act as an anchor, stabilizing metal atoms via metal–N coordination bonds. This stabilization leads to the formation of desired active sites for synthesizing various MOF derived SACs.144,145 The CO2 conversion efficiency of MOF and COF derived SACs depends on various intrinsic factors, such as the modulation of the active sites and morphology as well as extrinsic properties, including metal coordination environments, electronic structures, porosity, and reaction conditions.146
CO2 conversion by MOF derived SACs can be achieved through various methods, including thermochemical, electrochemical, photochemical, and other approaches, each with unique advantages and challenges. Thermochemical CO2 reduction involves high-temperature catalytic processes, such as the hydrogenation of CO2 to methanol or hydrocarbons via Fischer–Tropsch synthesis. This method often requires metal-based catalysts like Cu, Fe, or Ru to drive the reaction efficiently. Electrochemical CO2 reduction (eCO2R) utilizes renewable electricity and electrocatalysts to convert CO2 into valuable products such as carbon monoxide, formate, and hydrocarbons. Catalysts for eCO2R range from molecular complexes to nanostructured metals and single-atom catalysts.147 Photochemical CO2 reduction, on the other hand, mimics natural photosynthesis by using semiconductor photocatalysts (e.g., TiO2, perovskites, and MOF) to capture solar energy and drive CO2 conversion into chemicals like methane or methanol and so on. Each of these methods presents distinct challenges-thermochemical methods require high energy input, electrochemical approaches face selectivity and stability issues, photochemical systems suffer from low quantum efficiency, and biochemical strategies require careful control of biological activity. Future advancements in catalyst design, reactor engineering, and process integration are essential to improve efficiency and scalability for industrial CO2 utilization.
The eCO2RR is a complex multi-step process that involves the activation of CO2 and its subsequent conversion into various C1 products such as CO, formate, methane, and methanol, as well as C2 products such as ethylene, ethanol, and acetic acid (Fig. 9).147–153 The reaction occurs at the electrode–electrolyte interface, where CO2 molecules are adsorbed onto the catalyst surface and undergo electron transfer, protonation, and bond rearrangements. The efficiency and selectivity of CO2RR depend on several factors, comprising the choice of catalyst, applied potential, electrolyte composition, and reaction conditions. The mechanism generally begins with the activation of CO2, which is thermodynamically stable and requires an initial electron transfer to generate the CO2˙− radical anion. This intermediate can follow different pathways depending on the catalyst and reaction conditions. For instance, on transition metal catalysts like gold (Au) and silver (Ag), CO2˙− can be protonated to form adsorbed CO as a key intermediate. Depending on the intermediate binding energy, CO either desorbs from the catalyst surface, undergoes further reduction if the active sites have a sufficient binding affinity, or poisons the catalyst due to excessively strong binding. The electrochemical reduction of CO2 on single-atom catalysts happens through several proton-coupled electron transfer steps. Like on Au or Ag, the CO2 molecule attaches to the single metal atom, forming a bent M–CO2H intermediate that is stabilized by metal–carbon interaction. This activation step is needed for further reduction. With continuous electron and proton transfers, CO or other products are formed. The selectivity of SACs depends on how strongly the intermediates (especially *COOH and *CO) bind and on the metal atom's coordination environment. For example, Cu-based catalysts, which have intermediate binding affinity for CO, facilitate further proton-electron transfer steps, leading to C–C coupling and the formation of hydrocarbons for example ethylene and ethanol. The unique ability of Cu to stabilize CO and promote its dimerization is critical for the formation of C2+ products. Another important reaction pathway is the direct protonation of CO2˙− to formate, which is favoured in metals like Sn, Pb, and Bi. These catalysts stabilize the HCOO* intermediate, preventing its further reduction to CO or hydrocarbons. Similarly, the intermediate binding affinity of CO without dimerization leads to further reduced C1 products such as methane and methanol, which are typically dictated by the catalytic active site. Competing side reactions, particularly the hydrogen evolution reaction (HER), often reduce the efficiency of CO2RR. Suppressing HER is vital for improving the FE toward CO2 reduction products. Strategies such as electrolyte engineering, catalyst surface modification, and pH optimization help enhance CO2RR selectivity by tuning the reaction environment. Understanding the mechanistic pathways of CO2RR is essential for developing next-generation catalysts that can drive the reaction with higher efficiency, selectivity, and durability, paving the way for the sustainable usage of CO2 as a feedstock for fuels and chemicals. In the following sections, the synthetic methodology used to design various coordination structures and the effect of local structure on CO2RR will be discussed.
![]() | ||
Fig. 9 General mechanism and strategy for C1 products: (a) General mechanism for CO2RR to C1 products. (b) General mechanism for CO2RR to C2 products. |
By precisely tuning the electronic and geometric structure of the active site, catalytic performance can be optimized. Ligand field engineering modifies the d-band electronic states, affecting CO2 adsorption and activation pathways, while the introduction of heteroatoms near the metal site can regulate charge density, enhancing selectivity and reaction rates. Additionally, the dynamic restructuring of metal sites under reaction conditions can generate more active and resilient catalytic centers. Synergistic interactions between metal atoms and organic linkers further improve stability and prevent catalyst degradation. The coordination environment defines the immediate atomic arrangement surrounding each metal atom, specifying which atoms (e.g., nitrogen, carbon, oxygen, sulfur) are directly bonded to the metal and their spatial arrangement. This structure dictates the formation of different SAC active sites, such as M–N4, M–N2O2, or M–N3S, each exhibiting distinct catalytic properties. In the following sections, the key factors influencing CO2RR performance will be discussed, along with how synthetic strategies enable structural tunability to achieve high selectivity and efficiency (Table 2).
Electrocatalyst | Electrolyte | Potential (V) | Product | FE (%) | Ref. |
---|---|---|---|---|---|
M–N4 SAC | |||||
Ni SAs/N–C | 0.5 M KHCO3 | −0.89 | CO | 71.9 | 145 |
C-AFC©ZIF-8 (Fe–N) | 1.0 M KHCO3 | −0.43 | CO | 93 | 146 |
Fe–N–C | 0.5 M NaHCO3 | −0.6 | CO | 90 | 147 |
SE-Ni SAs@PNC | 0.5 M KHCO3 | −0.7 to 1.2 | CO | 90 | 148 |
Fe–N2+2–C8 | 0.5 M KHCO3 | −0.58 | CO | 93 | 149 |
Cu–N4C8 | 0.1 M KHCO3 | −0.8 | CO | 96 | 150 |
Co–N4 | 0.1 M KHCO3 | −0.8 | CO | 82 | 152 |
Ni SAs/NCNTs | 0.5 M KHCO3 | −0.9 | CO | 97 | 153 |
Ni–N–C | 0.5 M KHCO3 | −0.8 | CO | 96.8 | 153 |
Ni–N–C | 0.5 M KHCO3 | −0.66 to −0.96 | CO | 100 | 154 |
Fe–N4 | 0.1 M KHCO3 | −0.8 | CO | 90 | 155 |
Cu–N4–C/1100 | 0.1 M KHCO3 | −0.9 | CO | 98 | 156 |
Fe–N–C900 | 0.1 M KHCO3 | −1.2 V (V Ag/AgCl) | CO | 86.8 | 157 |
Fe–N–C-1000 | 0.5 M KHCO3 | −0.9 | CO | 100 | 158 |
Ni–N/C@900 | 0.1 M KHCO3 | −0.76 | CO | 90 | 159 |
M–Ni–N–C-2/CNTs | 0.1 M KHCO3 | −0.7 | CO | 98 | 160 |
Pyrrolic vs. Pyridinic | |||||
Ni SAC-1000 | 0.5 M KHCO3 | −0.8 | CO | 98.2 | 161 |
Cu-SA/NPC | 0.1 M KHCO3 | −0.36 | CH3COCH3 | 36.7 | 162 |
Fe3+–N–C | 0.5 M KHCO3 | −0.2 | CO | >80 | 163 |
Ni–N3–C | 0.5 M KHCO3 | −0.75 | CO | 99.37 | 164 |
Ni–NPyrrolic–C | 0.5 M KHCO3 | −0.85 V | CO | 92 | 165 |
Low-coordinate SACs | |||||
C–Zn1Ni4 ZIF-8 | 1 M KHCO3 | −0.83 | CO | 98 | 166 |
Co–N2 | 0.5 M KHCO3 | −0.68 | CO | 94 | 167 |
Cu–N–C-900 | 0.1 M KHCO3 | −1.6 | CH4 | 38.6 | 168 |
NiSA–N2–C | 0.5 M KHCO3 | −0.8 | CO | 98 | 82 |
Ni–N3–C | 0.5 M KHCO3 | −0.65 | CO | 95.6 | 169 |
MS-L-Ni–NC (Ni–N3–C) | 0.5 M KHCO3 | −0.8 | CO | 98.7 | 170 |
Fe/Ni–N–C | 0.5![]() |
−0.677 | CO | 92.9 | 171 |
Ni–NC–NS (Ni–N2–C) | 0.5![]() |
−1.0 | CO | 86 | 172 |
Ni–N3/NC | 0.1![]() |
−1.3 | CO | 94.6 | 173 |
Axially coordinated SACs | |||||
Fe-SA/ZIF (Fe–N5) | 0.1 M KHCO3 | −0.7 | CO | 98 | 174 |
Ni–N4–O/C | 0.5 M KHCO3 | −0.9 | CO | 100 | 175 |
Carbon coordinated SACs | |||||
Co–N2–C3 | 0.1 M KHCO3 | −0.8 | CO | 92 | 176 |
Co–C2N2 | 0.1 M KHCO3 | −0.8 | CO | 177 | |
Ni–N1–C3 | 0.5 M KHCO3 | −0.9 | CO | 97 | 178 |
Oxygen coordinated SACs | |||||
Fe1N2O2/NC | 0.1 M KHCO3 | −0.5 | CO | 99.7 | 179 |
FeN2O2/NC | 0.5 M KHCO3 | −0.7 | CO | 95.5 | 180 |
Sulphur coordinated SACs | |||||
Co–S1N3 | 0.5 M KHCO3 | −0.5 | CO | 98 | 181 |
MnN3S1 | 0.5 M KHCO3 | −0.45 | CO | 70 | 182 |
Ni–NSC | 0.5 M KHCO3 | −1.035 | CO | 98 | 183 |
Phosphorous coordinated SACs | |||||
Ni-SA/CN-P | 0.5 M KHCO3 | −0.8 | CO | 96.9 | 184 |
Ni–P1N3 | 0.5 M KHCO3 | −0.75 | CO | 98 | 185 |
Halogen coordinated SACs | |||||
Ni1–N–C (Cl) | 0.5 M KHCO3 | −0.7 | CO | 94.7 | 187 |
NiN4Cl–ClNC | 0.5 M KHCO3 | −0.7 | CO | 98.7 | 188 |
Ni–NBr–C | 0.5 M KHCO3 | −0.7 | CO | 97 | 189 |
FeN4Cl/NC | 0.5 M KHCO3 | −0.6 | CO | 90.5 | 190 |
P-block metal SACs | |||||
Inδ+–N4 | 0.5 M KHCO3 | −0.95 | HCOOH | 96 | 191 |
In–N–C | 0.5 M KHCO3 | −0.99 | HCOOH | 80 | 192 |
InA/NC | 0.5 M KHCO3 | −2.1 vs. Ag/Ag+ | CO | 97.2 | 193 |
In-SAC-1000 | 0.5 M KHCO3 | −0.6 | CO | 97 | 194 |
Bi SAs/NC | 0.1 M NaHCO3 | −0.5 | CO | 97 | 195 |
Al–NC | 0.1 M KHCO3 | −0.65 | CO | 98.76 | 196 |
SnN3O1 | 0.1 M KHCO3 | −0.7 | CO | 94 | 197 |
Hetero-metal SACs | |||||
NiCu-SACs/N–C | 0.5 M KHCO3 | −0.6 | CO | 92.2 | 198 |
Co0.5Ni0.5–N–C | 0.5 M KHCO3 | −0.5 to −1.1 | CO | 50 ± 5 | 199 |
Ni–Al NC | 0.1 M KHCO3 | −0.8 | CO | 98 | 200 |
Cu–In–NC | 0.1 M KHCO3 | −0.7 | CO | 96 | 201 |
Ni/Fe–N–C | 0.5 M KHCO3 | −0.7 | CO | 98 | 202 |
NiN3©CoN3–NC | 0.1 M KHCO3 | −1.1 | CO | 97.7 | 203 |
Ni/Cu–N6–C | 0.5 M KHCO3 | −0.6 | CO | 97.7 | 204 |
Ni–N3/Cu–N3 | 0.5 M KHCO3 | −1.1 | CO | 99.1 | 205 |
Cu–Fe–N6–C | 0.1 M KHCO3 | −0.7 | CO | 98 | 206 |
Fe/Cu–N–C | 0.1 M KHCO3 | −0.8 | CO | 99.2 | 207 |
O–Ni2–N6 | 1.0 M KHCO3 | −1.25 | CO | 94.3 | 208 |
Fe2–N6–C–O | 0.5 M KHCO3 | −0.8 | CO | 95.85 | 209 |
Fe2N6 | 0.1 M KHCO3 | −0.6 | CO | 96 | 210 |
CuNi-DSA/CNFs | 0.1 M KHCO3 | −0.98 | CO | 99.6 | 211 |
InNi DS/NC (O–In–N6–Ni) | 0.5 M KHCO3 | −0.7 | CO | 96.7 | 115 |
Fe1–Ni1–N–C | 0.5 M KHCO3 | −0.5 | CO | 96.2 | 212 |
Others | |||||
Ni–NPIC4 | 0.5 M KHCO3 | −0.65 | CO | 95.1 | 213 |
Ni1−N−C-50 | 0.5 M KHCO3 | −0.7 | CO | 96 | 214 |
Nix–N–C | 0.5 M KHCO3 | −0.7 | CO | 80 | 215 |
NiMn–N–C | 0.5 M KHCO3 | −0.72 | CO | 98.5 | 216 |
FeSAs/CNF-900 | 0.5 M KHCO3 | −0.47 | CO | 86.9 | 217 |
CHK-cOCTA | 0.1 M KHCO3 | −1.5 | CH4 | 54.8 | 218 |
Ni/HH | 0.5 M KHCO3 | −0.77 | CO | 97.9 | 219 |
Ni–NG–acid | 0.5 M KHCO3 | −0.9 | CO | 97 | 220 |
mesoNC–Fe | 0.1 M KHCO3 | −0.73 | CO | 85 | 221 |
Ni/NCTs | 0.5 M KHCO3 | −0.8 | CO | 100 | 223 |
Ni–NC3@Cu2O | 1.0 M KOH | −1.2 | C2H4 | 60 | 224 |
C2H5OH | |||||
CH3COOH | |||||
Ni SACs–Cu NPs | 1.0 M KOH | −0.7 | C2H4 | 80 | 225 |
C2H5OH | |||||
CH3COOH | |||||
P-NiSA/PCFM | 0.5 M KHCO3 | −0.7 | CO | 96 | 226 |
CuSAs/TCNFs | 0.1 M KHCO3 | −0.9 | CH3OH | 44 | 227 |
Ni–PCNFs | 0.1 M KHCO3 | −1.5 | CO | 98.6 | 228 |
Zn-SA/CNCl-1000 | 1.0 M KOH | −0.93 | CO | 97 | 229 |
Ni/Zn-6 | 0.1 M KHCO3 | −1.0 | CO | 94 | 230 |
CO2 + 2H+ + 2e− → CO + H2O −0.117 V vs. RHE | (1) |
![]() | ||
Fig. 10 (a) Pictorial representation of the formation of Ni SAs/N–C. (b) LSV of different Ni SAs. (c) Potential dependant CO FEs by Ni SAs/N–C. Reproduced from ref. 154 with permission from American Chemical Society, copyright 2017. (d) Free energy profile of M–N4–C10 and M–N2+2–C8 sites towards CO formation. Reproduced from ref. 158 with permission from American Chemical Society, copyright 2018. (e) CO partial current density as a function of Cu–N4C8 and Cu–N4C10 catalysts. (f) LSV curves of Cu–N4C8 and Cu–N4C10 catalysts. (g) and (h) FE of CO and H2 vs. applied potentials for (g) Cu–N4C8 and (h) Cu–N4C10 catalysts. Reproduced from ref. 159 with permission from Wiley (CC-BY 4.0), copyright 2024. (i) Scheme of the formation of Ni SAs/NCNTs. (j) LSV curves of Ni SAs/NCNTs. (k) Potential dependant CO FE by Ni SAs/NCNTs. Reproduced from ref. 161 with permission from Elsevier, copyright 2019. |
Yang et al. present a novel strategy for creating surface-enriched nickel single-atom catalysts (SAs) using thermal atomization of Ni NPs supported on NDC supports (SE–Ni SAS@PNC). This top-down approach facilitates high atom utilization on the catalyst surface. Through the annealing, the strong interaction between Ni nanoparticles and the N-doped carbon support causes the Ni nanoparticles to decompose, leading to the stabilization of atomically dispersed Ni atoms on the carbon surface. The progressive decomposition of Ni NPs ultimately results in their atomization. The SE–Ni SAS@PNC catalyst exhibits high FE over 90% in the potential range −0.6 to −1.0 V and at −0.9 V, the partial current density reaches to 16.42 mA cm−2.157 Unlike the physically absorbed iron in the pores of ZIF, the mixed metallic MOFs (ZIF-8) stabilize Fe dispersion in precursors through covalent bonding between Fe ions and 2-methylimidazole, where Fe ions are in the nodes. During annealing to produce M–N–C catalysts, Zn acts as a spacer, dispersing Fe to prevent the formation of metallic clusters. Using this strategy, Pan and co-workers achieved a high CO FE of 93% using single atomic Fe–N4 catalysts. Using DFT studies, they found that edge-hosted M–N2+2–C8 moieties bridging two adjacent armchair-like graphitic layers act as better catalysts than that of usually proposed bulk-hosted M–N4–C10 site implanted within a graphitic layer (Fig. 10d).158 Similarly, the precise control over the local atomic structure of Cu active sites was tested, specifically, by creating edge-hosted Cu–N4 sites within micropores between graphitic layers (Cu–N4C8) compared to bulk-hosted Cu–N4 sites (Cu–N4C10). The researchers synthesized Cu–N–C catalysts with varying Cu precursor amounts to precisely control the density of Cu–N4 sites and tune their local atomic arrangement and overall electronic structure (Fig. 10e–h). The edge-hosted Cu–N4C8 shows higher surface area and enhanced active sites that exhibited 96% FEco at −0.8 V vs. RHE with a current density of −8.97 mA cm−2. Whereas bulk-hosted Cu–N4C10 lacked the distinct micropore structure and performed far more poorly with a significantly narrower potential range and lower FEco. DFT calculations demonstrate that the unique micropore structure of the Cu–N4C8 catalysts modifies the Cu atoms d-orbital energy levels, shifting the d-band center upwards in the adsorbed *COOH intermediate. This modification decreases electron occupancy in the *COOH antibonding states, lowering the free-energy barrier for *COOH formation and reducing the activation energy, which accounts for increased reaction kinetics and rate enhancements. This correlates well with superior experimentally observed reaction performance characteristics for these edge-hosted materials compared to related materials exhibiting alternative Ni coordination.159
Similarly, by controlling the pyrolysis temperature of a zinc–cobalt bimetallic MOF, the researchers synthesized Co SAs with either four-coordinated Co with nitrogens (Co1–N4) or a mixture of nitrogens and carbons (Co1–N4−xCx) on a nitrogen-doped porous carbon (NDPC) support. However, the Co1–N4 catalyst exhibited superior performance (FE; 82%) with a current density of −15.8 mA cm−2 for CO production at −1.0 V vs. RHE. Mechanistic studies revealed that higher catalytic activity was attributed to the enhanced binding strength of CO2 and facilitated CO2 activation at the Co1–N4 active site.160 Although, the synthesis of Ni-SAC from ZIF is promising, scaling up is difficult due to the lower yield. To enhance the hetero atom doping and metal doping, additional nitrogen sources are required. For instance, Peilong Lu et al. utilized a facile method for the synthesis of single atomic nickel anchored N-doped carbon nanotubes (Ni SAs/NCNTs) by simply pyrolyzing the mixture of dicyandiamide and 2-methylimidazole with Zn/Ni salts (Fig. 10i–k). The Ni SAs/NCNTs exhibited high Ni loading as high as 6.63 wt% and achieved high FE >95% for CO from −0.7 to −1.0 V, with a peak FE of 97% at −0.9 V at an appreciable current density of 41.5 mA cm−2.161 To identify the intrinsic properties of SACs in eCO2R, several catalysts were synthesized using a similar protocol. However, creating catalysts with very similar microenvironment and support effects is challenging. The uniform catalyst structure allows for a direct comparison of different single-atom metals. This was verified by constructing isostructural multivariate metal–organic frameworks (MTV-MOFs) incorporating different metal porphyrin linkers (Fe, Co, Ni, Cu). The metal-porphyrin units acted as precursors, and ZrO2 acted as a template. The Ni1–N–C displayed exceptional CO2 reduction selectivity, up to 96.8% FE for CO, significantly exceeding those of Fe, Co, and Cu analogues.162 Similarly, precise control over the atomic environment of the single-atom metal site is always intriguing. Wang et al. report a systematic examination into the function of single atomic metal catalysts (Fe, Co, Ni) dispersed on NDC (M–N–C) catalysts. To achieve a metal-pyridine N (M–Nx) coordination structure, MOFs were used as precursors and double-stage pyrolysis was employed to create these catalysts. The first step involves the generation of C-ZIF by pyrolysis of ZIF-8 at 1000 °C followed by metalation, and the second pyrolysis step enhanced the generation of pyridine-like coordination with Ni-center. The selectivity for CO production followed the trend Ni > Co > Fe, while the activity towards CO production followed a different order: Co > Ni > Fe. The Ni–N–C catalyst exhibits outstanding selectivity with high FE close to 100%, for CO from −0.66 to −0.96 V.163
Likewise, the engineering of Fe–N4 active sites with controlled Fe–N bond lengths within a MOF-derived carbon matrix, offers precise control over the electronic and geometric environment of the catalytic centers. In this work, the catalytic performance was tuned by both intrinsic and extrinsic factors, and by systematically tuning the particle size, the Fe loading and Fe–N bond structures were tuned. The optimal Fe–N4 catalyst shows a CO FE exceeding 90% across a broad potential range, with a peak current density of 25 mA cm−2 at −0.8 V vs. RHE. Interestingly, spectroscopic analysis and DFT calculations validated the importance of Fe–N bond contraction in optimizing *COOH binding energies, reducing the energy barrier for CO2 reduction while maintaining high selectivity. This work not only provides fundamental insights into structure–activity relationships but also offers a rational design strategy for scalable and efficient CO2 electroreduction catalysts.164 Effects of annealing temperature on the stabilization of single atoms have been extensively explored in many previous studies, as higher annealing temperature enhances the graphitization of the carbon matrix. In this context, the authors achieve this by precise control of the catalyst's coordination environment to successfully synthesize atomically dispersed Cu–Nx (x = 3, 3.3, 3.8, 4) moieties anchored on a three-dimensional porous carbon support. The optimized Cu–N4 catalyst (Cu–N4–C/1100) displays significantly enhanced catalytic performance for CO2RR, achieving FE >90% for CO over a broad potential window (−0.6 V to −1.1 V vs. RHE) and a maximum of 98% FE at −0.9 V. This surpasses the CO selectivity of most of the previously reported copper-based single-atom catalysts. DFT calculations suggested the crucial role of the *COOH intermediate and its efficient interaction with the edge-hosted Cu–N4 sites in facilitating the desorption of *CO.165 Similarly, a series of Fe–N–C catalysts were synthesized via direct pyrolysis of iron metal organic framework (Fe-BTT) with high nitrogen content at 800, 900, and 1000 °C, followed by acid treatment. Among them, Fe–N–C900 catalyst showed exceptional electrochemical performance for CO2 reduction to CO, achieving a FE of 86.8% at −1.2 V vs. Ag/AgCl, with an overpotential of 496 mV and a current density of 1.26 mA cm−2. Advanced characterization techniques revealed a uniform cubic morphology and consistently distribution of pyridinic nitrogen and porphyrin-like Fe–Nx, which were significantly exposed after acid leaching. The Fe–N–C900 catalyst's superior performance was attributed to its high specific surface area, and mesoporous structure.166 In the same way, Xie et al. developed various metal nitrogen carbon (M–N–C) catalysts, where M represents Fe, Ni, Mn, Co, or Cu derived from MOFs for the electrochemical reduction of CO2 to CO. However, the Fe–N–C catalyst exhibited the best CO2RR catalytic performance, achieving nearly 100% FE for CO at −0.9 V vs. RHE even more than that of the Ni–N–C catalyst. Moreover, the Fe–N–C electrocatalyst annealed at 1000 °C displayed the highest catalytic activity and CO selectivity.167
High-temperature annealing processes always led to the agglomeration of metal ions to generate metal nanoparticles. MOFs are promising precursors due to their well-defined structures and ability to isolate metal centers. However, annealing the nitrogen-free MOF due to the absence of anchoring hetero atom usually ends up with metal nanoparticle decorated carbon. To overcome this issue spacers are used to separate the metal ions. The researchers synthesized Ni–N–C catalysts by the pyrolysis of bimetallic Ni/Mg-MOF-74 where Mg2+ acts as spacers and urea as a nitrogen source. The Ni–N–C@900 °C catalyst exhibited 90% FE for CO with a current density of 4.2 mA cm−2 at −0.76V vs. RHE. In contrast, the Mg-free Ni-MOF-74 mainly produced Ni nanoparticles resulting in decreased selectivity toward CO production.168 The sustainable and scalable strategy for producing efficient Ni–N–C catalysts for eCO2R using an environmentally friendly aqueous synthesis approach is demanding. Most of the current methods often use harmful organic solvents and produce catalysts with low metal loading and primarily microporous structures, hindering efficient mass transport. To overcome this issue, Yixin Zhang et al. used a surfactant-modified strategy to synthesize Ni–N–C catalysts from MOFs. Specifically, they used acetyltrimethylammonium bromide (CTAB) in an aqueous solution to modify the ZIF-8 MOF, facilitating better Ni2+ ion adsorption and creating mesopores in the resulting material. The ideal catalyst (M–Ni–N–C-2/CNTs) showed high catalytic activity (jCO; −19.8 mA cm−2) and selectivity for CO with a FE of 98% at −0.7 V vs. RHE at −1.0 V vs. RHE. Moreover, it was suggested that the pyrrolic nitrogen coordinated Ni–Nx sites contribute significantly to the enhanced performance of the catalyst.169
3CO2 + 16H+ + 16e− → CH3COCH3 + 5H2O −0.31 V vs. RHE | (2) |
![]() | ||
Fig. 11 (a) Schematic diagram of the formation of Ni SAC-1000. (b) CO FE of Ni SAC-1000. Potential dependent CO current density for Ni SAs/N–C. Reproduced from ref. 170 with permission from Royal Society of Chemistry (CC-BY 4.0), copyright 2024. (d) DFT optimized structure of Ni–Npyrrolic–C and Ni–Npyridinc–C. (e) LSV curves of Ni–Npyrrolic–C and Ni–Npyridinc–C. (f) CO FE of Ni–Npyrrolic–C and Ni–Npyridinc–C. Reproduced from ref. 174 with permission from Elsevier, copyright 2023. (g) FE of CO2 reduction products of Cu–pyrrolic–N4 catalyst. (h) Free energy diagram of acetone formation for Cu–pyrrolic–N4 and Cu–pyridinic–N4. Reproduced from ref. 171 with permission from Nature (CC-BY 4.0), copyright 2020. (i) Scheme of the formation of NiNCMx–y. (j) LSV curves of NiNCMx–y. (k) Potential vs. FE of CO for NiNCMx–y. Reproduced from ref. 173 with permission from Elsevier, copyright 2023. |
Annealing temperature always plays a pivotal role in achieving a new coordination motif. Prolonged annealing conditions can enhance the evaporation of the Zn as well as nitrogen and it may create defect sites in the carbon matrix. By employing the self-sacrificial template method, the researchers achieved a high Ni loading of 5.44 wt% while maintaining atomic Ni dispersion. The bimetallic ZIF precursor yielded the C–Zn1Ni4 ZIF-8 catalyst under controlled pyrolysis conditions and exhibited remarkable CO2RR performance high CO FE of 98% at −1.0 V vs. RHE. DFT calculations suggest that enhanced activity stems from the unique electronic properties of the coordinatively unsaturated atomically dispersed Ni–N2V2 structural motif as the most favourable active site (Fig. 12a–c).175 Similarly, high-temperature annealing processes were considered as a strategy to make less nitrogen-coordinated SACs at high-temperature enhance carbon coordination along with nitrogen ligands. By carefully managing the pyrolysis temperature during catalyst synthesis, the researchers selectively prepared Co catalysts featuring two-coordinate (Co–N2), three-coordinate (Co–N3), and four-coordinate (Co–N4) Cobalt centers. The results reveal that the Co–N2 catalyst achieves superior CO2RR performance compared to Co–N3 and Co–N4. The Co–N2 exhibited a remarkable CO FE of 94% at an overpotential of 520 mV along with a current density of 18.1 mA cm−2 (vs. RHE). The enhanced activity and selectivity are attributed to the lower coordination number in Co–N2, which facilitates the activation of CO2 to form the CO2˙− intermediate, a key step in CO2RR.176
![]() | ||
Fig. 12 (a) Graphical representation of the synthesis of C–ZnxNiy ZIF-8. (b) CO partial current density of C–ZnxNiy ZIF-8. (c) Applied potential dependence of CO FE of C–ZnxNiy ZIF-8. Reproduced from ref. 175 with permission from Royal Society of Chemistry, copyright 2018. (d) Fabrication of low-coordination single-atom Ni electrocatalysts Ni–N3–C. (e) LSV curves of Ni–N3–C. (f) FE of CO for Ni–N3–C. Reproduced from ref. 178 with permission from Wiley, copyright 2021. (g) Schematic of the synthesis method and proposed vacancy-manipulated Ni–Nx active site architecture of Ni–NC–NS. (h) LSV curves of Ni–NC–NS. (i) FE of CO Ni–NC–NS. Reproduced from ref. 181 with permission from Wiley (CC-BY 4.0), copyright 2024. |
Copper is mostly acting as a better catalyst for C2-product selectivity, and it was already established that nearby cooper center is crucial for C–C coupling in most catalysts. However, controlling the proximity of active sites in single-atom copper catalysts is always challenging. The author successfully controlled the distance between neighboring Cu–Nx sites and the metal loading by controlling the annealing temperature, which demonstrated the ability to control the selectivity towards methane or ethylene. At lower copper concentrations (<2.4 mol%), the Cu–Nx sites are sufficiently far apart to favour the formation of methane (CH4), a C1 product, as the dominant CO2 reduction pathway. At higher concentrations (4.9 mol%), the proximity of the copper sites enables C–C coupling to form ethylene (C2H4), a C2 product. This illustrates a critical relationship between catalyst structure (Cu site density) and the selectivity of the CO2RR. DFT calculations supported that the isolated Cu–N4 and the isolated Cu–N2 sites predominantly yield CH4 while two proximate Cu–N2 sites allow two CO intermediates to bind facilitating C–C coupling towards C2H4 formation.177 Precise control over the coordination environment of SACs is crucial for optimizing catalytic activity and selectivity, yet is difficult to achieve. Existing methods often rely on MOFs containing abundant nitrogen, limiting their generalizability. The researchers developed a “host–guest cooperative protection” approach. This involves the synthesis of a bimetallic MgNi-MOF-74 which acts as a host. Pyrrole monomers were then introduced into the MOF channels and polymerized in situ, acting as guest. The Mg2+ ions present in the host spatially separate the Ni2+ ions within the MOF structure, preventing Ni atom clustering during pyrolysis. Likewise, polypyrrole acts as both a nitrogen source (for stabilizing Ni) and a protective agent preventing aggregation. By controlling the pyrolysis temperature, three catalysts NiSA–N4–C, NiSA–N3–C, and NiSA–N2–C with differing Ni–N coordination numbers (4, 3, and 2 respectively) were successfully synthesized. Among them, NiSA–N2–C displayed significantly better catalytic activity with 98% FE for CO at −0.8 V (vs. RHE). DFT study proved that NiSA–N2–C favoured COOH* intermediate formation (rate-determining step), explaining the catalyst's high activity.89 Existing methods often lack precise control over the coordination number of the central atom. Therefore, post-synthetic metal substitution (PSMS) was employed as a strategy to achieve precise control over the coordination number. The researchers employed a two-step strategy, which involves the synthesis of zinc-based metal–organic framework (ZIF-8) and pyrolysis to generate a NDC material (Zn–N3–C) followed by etching and Ni substitution resulting in a new catalyst (Ni–N3–C). Ni–N3–C exhibited higher FE for CO production (up to 95.6%) compared to Ni–N4–C and the undoped N-doped carbon material (N–C) (Fig. 12d–f).178 The crucial innovation lies in utilizing a 2D bimetallic ZIF precursor and a molten salt method, which prevents the typical aggregation and structural collapse observed in high-temperature pyrolysis of such materials. This leads to a unique catalyst with a high surface area and density of well-defined, highly active Ni sites, ultimately resulting in superior catalytic performance. The molten salt method is pivotal in avoiding the typical structural degradation during pyrolysis that commonly occurs with zeolite-based precursors. By using molten NaCl/KCl, they create an ionic liquid medium during pyrolysis that encapsulates and stabilizes the precursor nanosheets preventing agglomeration and structural collapse and yielding ultrathin layered nanosheets after pyrolysis and acid-washing treatment.
The exceptional properties translate into outstanding electrocatalytic performance in the CO2RR. The MS–L–Ni–NC demonstrates high CO FE (FEco) of >95.9% over a broad potential window, achieving an ideal FEco of 98.7% at −0.8 V vs. RHE (reversible hydrogen electrode), substantially outperforming catalysts prepared via alternative methods (MS–Ni–NC, D–L–Ni–NC). The superior results are corroborated by unusually large partial current densities (20.6 mA cm−2 at −1.0 V vs. RHE). DFT calculations provide compelling mechanistic support for the observed activity and selectivity. The analysis confirms that the coordinated unsaturated pyridine Ni–N2 sites are responsible for high activity because the calculations of Gibbs free energy differences predict these are favourable for the formation of *COOH (a reaction intermediate) and CO desorption, both key steps in the CO2RR reaction.179
Traditional pyrolysis of MOF precursors at elevated temperature often results in the formation of inactive metal nanoparticles due to the aggregation of metallic atoms. This research cleverly avoids this using a molten salt (NaCl/KCl) bath during pyrolysis of the layered Fe/Ni-ZIF-8 structure that avoids agglomeration due to preventing nanoparticle collapse or shrinkage. This preserves the precursor's 2D layered morphology, preventing the loss of active sites. It presents a strategy to create highly efficient and durable bimetallic Fe/Ni–N–C catalysts for eCO2R by using a combination of Fe/Ni–ZIF-8 as the precursor, and pyrolysis in an ionic-liquid molten salt bath. The key innovation is the use of a 2D layered precursor which avoids agglomeration and enables the controlled generation of well-exposed and highly active sites. The subsequent Fe/Ni–N–C catalyst achieves a remarkable CO FE (FEco) of 92.9% and TOF of 10.77 × 103 h−1 in a typical H-cell configuration. It further demonstrates the synergistic effects of Fe and Ni in promoting superior charge transfer, making this novel material robust and stable.180 Utilizing a crystallographically engineered MOF nanosheet as a support material for atomically dispersed nickel (Ni) single atoms, where controlled vacancy formation maximizes the density of unsaturated NiN2–V2 sites and is directly responsible for enhanced activity. The catalyst, Ni–NC–NS, was synthesized via solvent-controlled growth of 2D MOF nanosheets (ZIF-8–NS), followed by nickel incorporation and a carefully controlled pyrolysis at 950 °C. This process removes Zn from the structure, precisely generating Ni–NC–NS in nanosheet morphology while preventing unwanted Ni particle formation and/or aggregation. The Ni–NC–NS catalyst exhibited almost 100% FE for CO (FEco) from −0.7 to −1.0 V versus RHE (Fig. 12g–i). The NiN2–V2 sites exhibit a significantly lower energy barrier for *COOH intermediate formation compared to materials containing different numbers of atomic sites.181 SACs, especially those with low-coordination configurations, present a potential solution because of their unique electronic structures and maximized atom utilization. A coordinately unsaturated Ni–N3 single-atom electrocatalyst supported on a MOF-derived N-doped carbon (NC) support was prepared. The MOF was used to generate an N–C framework with abundant exposed N sites, followed by the introduction of Ni atoms to generate an atomically precise, low-coordination structure, creating active Ni–N3 sites within a carbon support matrix. The Ni–N3/NC catalyst exhibited highly efficient CO2-to-CO conversion with a FE of 94.6% at a current density of 100 mA cm−2.182
![]() | ||
Fig. 13 (a) Synthetic scheme of Fe–SA/ZIF preparation. (b) Polarization curve of Fe–SA/ZIF with various materials. (c) FE of CO for Fe–SA/ZIF. Reproduced from ref. 183 with permission from Elsevier, copyright 2021. (d) Demonstration of the synthesis process of Ni–N4–O/C. (e) LSV of Ni–N4–O/C. (f) FEs of CO at various potentials for Ni–N4–O/C. (g) The differences in CO FEs of Ni–N4–PRO/C and Ni–N4–O/C. Reproduced from ref. 184 with permission from Wiley, copyright 2021. |
![]() | ||
Fig. 14 (a) Graphic demonstration of Ni–SAC preparation. (b) and (c) LSV curve and FE of CO for Ni N1–C3. Reproduced from ref. 187 with permission from Springer, copyright 2023. (d) EXAFS fitting curves of Co–N2C3 at R space with proposed model inserted. (e) Polarization curves of Co–N5−xCx and NC. (f) FEco for Co–N5−xCx. Reproduced from ref. 185 with permission from Elsevier, copyright 2022. (g) EXAFS fitting of Co NP–SNC at R space. (h) Polarization curves of Co–HNC and Co NP–SNC. (i) Dependence of productivity and FE of Co–HNC on the applied potential. Reproduced from ref. 186 with permission from Wiley, copyright 2018. |
However, the CO–N2C2 demonstrates poor CO FE of 35% at 1.0 V but 100% FE for syngas (Fig. 14g–i).186 Single–atom catalysts (SACs) have shown promise, with low-coordinated structures (e.g., M–Nx–Cy where x < 4) offering potentially superior activity compared to the more common and often less active M–N4 configurations. A novel Ni SACs (Ni–N1–C3) was synthesized via an in situ method involving the pyrolysis of a nickel-impregnated metal–organic framework (MET-6) precursor, followed by an acid washing step (Fig. 14a–c). The optimized Ni–N1–C3 catalyst demonstrated a high CO FE of 97% at −0.9 V vs. RHE. DFT calculations revealed that the unique electronic structure of Ni–N1–C3 featuring a low coordination number, coupled with an upward-shifted d-band center and rich charge density on Ni sites, resulted in favourable *COOH adsorption and decreased energy barrier for *COOH formation. The results demonstrate a feasible pathway to efficiently reduce CO2 by tuning electronic structures through control of coordination and catalyst support.187
![]() | ||
Fig. 15 (a) Graphically presentation of the synthesis of Fe1N2O2/NC. (b) LSV curves of Fe1N2O2/NC, Fen/C, and NC catalysts. (c) CO partial current densities Fe1N2O2/NC. Reproduced from ref. 188 with permission from Royal Society of Chemistry, copyright 2022. (d) Synthesis illustration of FeN2O2/NC. (e) Polarization curves of FeN2O2/NC. (f) CO partial current densities of FeN2O2/NC. Reproduced from ref. 189 with permission from Elsevier, copyright 2025. |
![]() | ||
Fig. 16 (a) Synthetic scheme of Co–SxN4−x SACs. (b) Polarization curves of Co–SxN4−x SACs. (c) FE of CO for Co–SxN4−x SACs. (d) Intermediate free energy changes for for CO2RR to CO on the Co–SxN4−x SACs. Reproduced from ref. 190 with permission from Nature (CC-BY 4.0), copyright 2025. (e) Graphical drawing of the preparations of MnSA/SNC. (f) LSV profiles of the MnSA/NC and MnSA/SNC. (g) CO2RR FE at various applied potentials of the MnSA/SNC. Reproduced from ref. 191 with permission from American Chemical Society, copyright 2021. (h) Synthetic scheme of Ni–NSC. (i) LSV curves of Ni–NSC. (j) FEco at various applied potentials for Ni–NSC. Reproduced from ref. 192 with permission from Elsevier, copyright 2024. |
The phosphorus doped Ni–SA/CN–P catalyst exhibit 91.8% of CO FE at −1.1 V and the CO current density in the flow cell reached to 91.2 mA cm−2, which is better than that of the sample Ni–SA/CN without phosphorous dopant. The introduced P atoms adjust the electronic environment around Ni (Fig. 17a–c). XPS and DFT analysis showed that the P dopants reduce the energy required to form the crucial *COOH intermediate (a key step in the CO2RR process), likely through modification of the d-band energy levels of the nickel active sites, causing shifts in charge distribution, thereby reducing the energetic barrier to the CO2RR process.193 While metal–nitrogen (M–Nx) sites in N-doped carbon materials are promising, introducing other heteroatoms, like phosphorus (P), can further tune the electronic structure and enhance catalytic performance. Nickel single-site catalysts (SSCs) with dual-coordinated phosphorus and nitrogen (Ni–PxNy, x = 1,2; y = 3,2) atoms were synthesized via an in situ phosphorization method using a 2-methylimidazole and triphenylphosphine (PPh3) containing precursor (Fig. 17d–f). Ni-P1N3 exhibited significantly superior CO2RR activity than Ni–N4. Specifically, Ni–P1N3 showed a CO FE in the range of 85.0%–98.0% over a wide potential range (−0.65 to −0.95 V vs. RHE) with a CO current density reaching 14.30 mA cm−2. DFT calculations showed that the asymmetric Ni–P1N3 structure was more energetically favourable for CO2 intermediate (COOH*) adsorption/desorption compared with the Ni–N4 configuration. These enhanced binding properties resulted in accelerated reaction kinetics, higher CO selectivity, and enhanced catalytic performance. This work provides an effective strategy for creating highly active and selective SSCs and demonstrates how controlling the coordination environment by utilizing multiple heteroatoms is crucial for maximizing CO2RR efficiency and targeting specific reaction products.194
![]() | ||
Fig. 17 (a) Synthetic scheme of Ni–SA/CN–P. (b) LSV curves of Ni–SA/CN–P. (c) Potentialdependent CO FE for Ni–SA/CN–P. Reproduced from ref. 193 with permission from Elsevier, copyright 2024. (d) Synthetic scheme of Ni–PxNy. (e) LSV graph of Ni–PxNy. (f) CO FE of the Ni-PxNy at different potentials. Reproduced from ref. 194 with permission from Springer, copyright 2023. |
![]() | ||
Fig. 18 (a) Synthetic scheme of the Ni–N–C(X). (b) LSV curves of Ni–N–C(X). (c) CO faradaic efficiencies of Ni–N–C(X). Reproduced from ref. 196 with permission from Springer, copyright 2022. (d) Schematic diagram to produce NiN4Cl–ClNC. (e) LSV curves of NiN4Cl–ClNC. (f) CO faradaic efficiencies of NiN4Cl–ClNC. Reproduced from ref. 197 with permission from Wiley (CC-BY 4.0), copyright 2023. (g) Schematic diagram of the synthetic route of Ni–NBr–C. (h) LSV curves of Ni–NBr–C. (i) FEco of Ni–NBr–C. Reproduced from ref. 198 with permission from Wiley, copyright 2025. (j) Pictorial demonstration of the synthesis of FeN4Cl/NC. (k) LSV curves of various FeN4Cl/NC. (l) Potential dependent CO FE of FeN4Cl/NC-7.5. Reproduced from ref. 199 with permission from Elsevier, copyright 2022. |
To address the challenges in CO2 electroreduction a high-performance membrane electrode assembly (MEA) was designed using a novel single atom catalyst. An asymmetrically coordinated Ni single-atom catalyst, Ni–NBr–C, featuring axial Br coordination at NiN4Br sites anchored onto hollow Br/N co-doped carbon nanocages, was synthesized using a NaBr-assisted confined-pyrolysis strategy (Fig. 18g–i). The Ni–NBr–C catalyst demonstrates superior performance in an MEA device, achieving a CO FE exceeding 97% over a current density range of 50–350 mA cm−2. This work strongly suggests this catalyst has significant potential for industrial-scale CO2 electroreduction applications.198 Introducing axial atoms into M–N–C SACs is proposed as a general strategy for designing advanced catalysts for various electrochemical applications. Thus, iron–nitrogen–carbon single-atom catalysts (Fe–N–C SACs), often exhibiting relatively low selectivity and activity due to strong CO* binding on Fe sites, were modified by introducing an axial chlorine (Cl) atom to create FeN4Cl/NC. This was achieved by pyrolyzing Fe-loaded 2D ZIF nanosheets followed by low-temperature incubation in HCl (Fig. 18j–l). X-ray absorption spectroscopy (XAS) confirmed the FeN4Cl structure with axial Cl coordination at 2.26 Å and four N at 2.02 Å. The resulting catalyst demonstrates a CO FE of 90.5%, and high current density (10.8 mA cm−2) at a low overpotential of 490 mV. The enhanced performance was attributed to the Cl-induced modulation of Fe electronic structure facilitating CO* desorption and inhibiting H* adsorption, as predicted and supported by DFT calculations.199
Most of the SACs described in the literature are based on d-block metals due to their diverse oxidation states, and readily tunable electronic properties. Similarly, in recent years p-block elements as single atoms enable precise control over the electronic and geometric structure of the active site, maximizing atom utilization efficiency and potentially leading to improved selectivity and activity. However, challenges exist because p-block SACs often exhibit lower stability compared to their d-block counterparts due to their weaker metal–ligand interactions. Furthermore, controlling the coordination environment and oxidation state of the isolated p-block atom presents a significant synthetic challenge. The In-SAs/NC catalyst was synthesized via a facile two–step method. First, indium(III) acetylacetonate [In(acac)3] was encapsulated within a ZIF-8 matrix. Subsequently, pyrolysis of the In(acac)3@ZIF-8 precursor under an inert atmosphere resulted to the formation of an In-SAs/NC catalyst. The In-SAs/NC catalyst demonstrated a high FE of 96% for formate at −0.65 V vs. RHE, with a current density of 8.87 mA cm−2.200 The single atomic In catalyst (In–N–C) was prepared by pyrolyzing In-doped ZIF-8 at 900 °C in an N2 atmosphere. The increasing ratio of In:Zn in In–N–C is because of the evaporation of Zn ions during pyrolysis, whereas In mostly remains in the catalyst. The In–N–C acts as an effective catalyst for CO2 to formic acid/formate in aqueous media with a FE of ∼80% at −0.99 V relative to RHE. The atomic In (Indium) has strong electronic interaction with neighboring nitrogen atoms on the carbon skeleton, enabling efficient conversion of CO2 to formate. DFT calculations suggested that the low energy barrier for the *OCHO intermediate formation improved the CO2 conversion to format.201
While indium exhibits format selectivity, exploiting local coordination environment design principles of single atom catalyst, possibly switches the dominant reaction pathway from formate production to CO generation. The researchers synthesized, InA/NC catalyst via pyrolysis of an indium-BTC metal–organic framework (MOF) and dicyandiamide (DCD). By the controlled pyrolysis (temperature ranges from 700–1000 °C) successful dispersion and coordination of indium at an atomic level was achieved. The InA/NC catalyst pyrolyzed at 1000 °C, demonstrates remarkable selectivity towards CO production in a mixed ionic liquid/acetonitrile (MeCN) electrolyte, in stark contrast to typical In-based catalysts that favour format production. The FE for CO reaches 97.2%, with a total current density of 39.4 mA cm−2.202 Similarly, to achieve the selectivity shift, the researchers synthesized In-SAs/NC catalysts using a two-step method involving the encapsulation of In(acac)3 within a MOF (MIL-68) derived from 1,4-benzenedicarboxylic acid, and pyrolysis at controlled temperatures (800 °C and 1000 °C, yielding In-SAs-800 and In-SAs-1000). In-SAs-1000 (prepared at 1000 °C), exhibited an exceptionally high FE for CO (up to 97% at −0.6 V vs. RHE), compared to In-SAs-800 which favoured formate (Fig. 19a–c). The selectivity shift was directly correlated with nitrogen content, which is controlled by pyrolysis temperature. DFT calculations suggest that In–N4 centers favour formate production, while, as nitrogen content decreased while increasing the pyrolysis temperature, neighbouring carbon atoms emerge as the catalytic centers responsible for preferential CO production.203
![]() | ||
Fig. 19 (a) Synthetic scheme of InSAC. (b) LSV curves of InSAC. (c) FEco of InSAC at various potentials. Reproduced from ref. 203 with permission from Springer, copyright 2022. (d) Representation of the conversion from Bi-MOF to single Bi atoms. (e) LSV curves of Bi SAs/NC. (f) FEco of Bi SAs/NC at various potentials. Reproduced from ref. 204 with permission from American Chemical Society, copyright 2019. (g) Schematic demonstration of Al–NC. (h) LSV curves of Al–NC synthesis. (i), FEco for Al–NC at various potentials. Reproduced from ref. 205 with permission from American Chemical Society, copyright 2024. (j) Schematic representation of activity difference between Sn–N4 vs. SnN3O. (k) Polarisation curves of Sn–NOC and Sn–Pc. (l) FE of CO and H2 catalyzed by Sn–NOC at various potentials. Reproduced from ref. 206 with permission from Wiley (CC-BY 4.0), copyright 2021. |
Erhuan Zhang et al. used similar approach in synthesis of Bi–N4 SACs by combined pyrolysis of Bi-MOF and dicyandiamide. By in situ environmental transmission electron microscopy (ETEM) the author confirmed the initial Bi nanoparticle formation followed by atomization, facilitated by the ammonia (NH3) released from the decomposition of dicyanamide during pyrolysis. Characterization confirms the presence of Bi SAs coordinated with four nitrogen atoms (Bi–N4) within the porous carbon support (Fig. 19d–f). The Bi SAs/NC catalyst exhibits higher activity for CO2RR to carbon monoxide (CO) compared to bismuth clusters or nanoparticles (Bi NPs/NC, Bi Cs/NC) attaining a high FE of 97% at a very low overpotential of 0.39 V vs. RHE.204 Likewise, being one of the most abundant elements in Earth's crust, the researchers synthesize a SAC with atomically dispersed aluminium (Al) atoms coordinated with four nitrogen atoms (Al–N4) on a NDC support. The Al–NC catalyst synthesized using complex-assisted pyrolysis strategy (Fig. 19g–i). A NDC support was first synthesized by pyrolyzing the ZIF-8 followed by the Al(phen)3 complex was successfully adsorbed onto the NC–(Al(phen)3@NC) and pyrolyzed again. The Al–N4 site cleverly manages the kinetic barrier problem found with other, simpler, traditional single-atom transition-metal catalysts. The Al–NC catalyst exhibits high FE for CO of 98.76% at −0.65 V vs. RHE and dramatically surpassing the performance of other traditional single atom nickel (Ni–NC) and iron (Fe–NC) catalysts. DFT calculations confirm the comparatively lower energy barrier for *COOH formation and *CO desorption when compared to those using alternative single atom transition metal-based catalysts studied.205 The synthetic control over both atomic-level catalyst site configurations and their influence on regulating and tuning reaction pathway energies to maximize yield towards different products is promising. The SACs with Sn–N4 configuration is known for its poor CO selectivity because it predominantly generates formic acid (as denoted in eqn (3)) and hydrogen (H2). To overcome this issue, the researchers synthesized SnN3O1 motif supported on a N-doped carbon matrix (NC) by a gas transport strategy. NC was obtained by simple pyrolysis of ZIF-8 precursor and used as substrate and SnO2 powder was used as the Sn and O sources. During annealing, volatile Sn species, from the evaporation of SnO2, transferred to the surface of NC and were coordinated by heteroatoms, forming the Sn–NOC electrocatalyst (Fig. 19j–l). Especially, the loading level of Sn atoms in Sn–NOC catalysts can be controlled by adjusting the pyrolysis temperature from 950 to 1050 °C. The Sn–NOC annealed at 1000 °C catalyst exhibits a maximum FE of 94% at −0.7 V versus RHE with a CO partial current density of 13.9 mA cm−2. DFT calculations support the finding that the unique SnN3O1 atomic arrangement reduces the activation energy barriers for *COOH formation, while significantly increasing the activation barrier for the competing HCOO intermediate formation.206
CO2 + 2H+ + 2e− → HCOOH −0.25 V vs. RHE | (3) |
![]() | ||
Fig. 20 (a) Synthetic scheme of NiCu–SACs/N–C catalyst. (b) Comparative LSV of different NiCu–SACs/N–C catalyst. (c) FE of CO for NiCu–SACs/N–C at various potentials. Reproduced from ref. 207 with permission from Elsevier, copyright 2024. (d) Pictorial representation of the synthesis of CoxNi1−x–N–C. (e) LSV curve of Co0.5Ni0.5–N–C. (f) FE of CO for Co0.5Ni0.5–N–C at various potentials. Reproduced from ref. 208 with permission from American Chemical Society, copyright 2023. (g) Schematic demonstration of the synthesis of Ni–Al NC catalyst. (h) LSVs of different Ni–Al NC. (i) FE of CO for Ni–Al NC at various applied potentials. Reproduced from ref. 209 with permission from Royal Society of Chemistry, copyright 2024. (j) Pictorial representation of the synthesis of Cu–In–NC. (k) LSV curve of Cu–In–NC. (l) FE of CO for Cu–In–NC at various potentials. Reproduced from ref. 210 with permission from Wiley, copyright 2022. |
The incorporation of dissimilar metal centers close to the catalytically active site correspondingly induces the electronic modulation of the catalysts, which leads to altering the intermediate binding energy and the intrinsic catalytic activity and stability as well. The researchers introduced diatomic Ni–Fe sites anchored on nitrogen doped carbon. This catalyst demonstrates superior performance compared to its single-atom counterparts. The Ni/Fe–N–C catalyst was synthesized via pyrolysis of a Fe-doped Ni–ZIF-8 precursor following an ion-exchange strategy. The Ni/Fe–N–C catalyst displays high CO FE >90% over a broad potential window (−0.5 to −0.9 V vs. RHE), reaching 98% at −0.7 V vs. RHE. It maintains exceptional selectivity (>99%) for over 30 hours of continuous operation (Fig. 21a–c). DFT calculations indicate that the Ni–Fe pair acts synergistically, lowering the reaction energy barriers associated with *COOH formation and CO desorption resulting in greatly improved activity compared to monometallic Ni and Fe centers.211 Coupling hetero-nuclear transition metal atoms with well-controlled synthesis methods, promotes the formation of well-defined active sites, thereby maximizing the surface areas conducive to high mass transfer and minimizing the undesired competing reactions. The researchers synthesized the NiN3@CoN3–NC catalyst via a multistep process involving the preparation of Zn/Co/Ni–zeolite imidazolate frameworks (ZIFs), followed by pyrolysis. This approach effectively creates isolated and well-defined Ni and Co dual-atom active sites. The coupled Ni–Co dual-atom catalyst displays significantly enhanced CO2 reduction activity than that of its single-atom equivalents (Ni SACs and Co SACs). The NiN3©CoN3–NC shows FE towards CO formation (FEco > 90% in the potential window from −0.6 V to −1.1 V) reaching a maximum FEco of 97.7% at −0.7 V with a partial current density of 14 mA cm−2 at −1.1 V vs. RHE (Fig. 21d–f).212 Similarly, a copper atom was introduced near to nickel site and the Ni/Cu–N–C was synthesized via a host–guest approach using a Cu–phenanthroline complex within a NiZn–ZIF–8 precursor. The Ni/Cu–N–C catalyst exhibits a record-high TOF of 20695 h−1 at −0.6 V vs. RHE and FE of 97.7% for CO. The incorporation of Cu near the Ni center leads to an upward shift of Ni 3d band energy, favourably influencing the interaction between Ni and the *COOH intermediate. This reduces the activation energy for the rate-determining step (*COOH formation). Operando XAS supports this, showing a reversible shift in the Ni oxidation state during catalysis indicating charge transfer processes (Fig. 21g-i).213 Similarly, Ni–Cu dual atom sites supported on hollow NDC were developed and demonstrated remarkable performance than that of single-metal catalysts and previously reported dual-atom catalysts. The atomically dispersed Ni–Cu catalyst exhibits near FE of 100% CO across all pH condictions, with CO partial current densities of 190 ± 11 mA cm−2 (acidic), 225 ± 10 mA cm−2 (neutral), and 489 ± 14 mA cm−2 (alkaline). DFT calculations suggest that Cu pushes the Ni d-band center toward the Fermi level. This modulates the binding energy of the COOH* intermediate and accelerates its formation, thereby contributing toward increased and overall, significantly boosting catalytic conversion to CO, while further reducing those reaction energies (Fig. 21j–l).214
![]() | ||
Fig. 21 (a) Schematic of Ni/Fe–N–C synthesis. (b) Linear sweep voltammograms (LSVs) of Ni/Fe–N–C catalysts. (c) CO faradaic efficiency (FE) for Ni/Fe–N–C at various potentials. Reproduced from ref. 211 with permission from Wiley, copyright 2019. (d) Schematic of NiN3@CON3–NC synthesis. (e) LSV of NiN3@CON3–NC catalyst. (f) CO FE for NiN3@CON3–NC at various potentials. Reproduced from ref. 212 with permission from Elsevier, copyright 2023. (g) Schematic of Ni/Cu–N–C synthesis. (h) LSVs of Ni/Cu-N–C catalysts. (i) CO FE for Ni/Cu–N–C at various potentials. Reproduced from ref. 213 with permission from American Chemical Society, copyright 2022. (j) Schematic of Cu/Ni–NC synthesis. (k) LSV of Cu/Ni–NC catalyst. (l) CO FE for Cu/Ni–NC at various potentials. Reproduced from ref. 214 with permission from Wiley, copyright 2023. |
Controlled structural modification with a synergistic relationship of single atom sites significantly impacts efficiency, stability, and selectivity. For instance, researchers developed Cu and Fe diatomic sites and demonstrated the synergistic relationship between them in CO2RR compared to monometallic Cu–N–C and Fe–N–C catalysts. Cu–Fe–N6–C was synthesized via pyrolysis of a PcCu–Fe–ZIF-8 precursor. The Cu–Fe–N6–C catalyst exhibits a high FE of 98% at −0.7 V vs. RHE, maintaining 98% of its initial FE after 10 hours of continuous electrolysis.
The superior activity is ascribed to the cooperative catalytic effect of the Cu and Fe sites, where the Cu–Fe diatomic structure enhances the adsorption of CO2 and lowers the activation energy barrier for CO2 reduction to CO. This directly promotes efficient catalytic processes and produces high yield of desired products toward maximization of efficiency (Fig. 22a–c).215 Precisely controlled Fe–Cu diatomic sites within a nitrogen-rich carbon matrix optimized to enhance CO2 reduction efficiency. Fe/Cu–N–C was synthesized through a MOF-assisted approach by incorporating Fe and Cu precursors into a ZIF-8 framework. Careful control of precursor placement, using Fe(acac)3 within the ZIF-8 cavities and Cu directly coordinated to the N atoms within the structure minimized aggregation, generating an optimal N4Fe–CuN3 diatomic structure. The Fe/Cu–N–C catalyst demonstrates a high CO FE (>95% over a broad potential window of −0.4 to −1.1 V vs. RHE, reaching 99.2% at −0.8 V vs. RHE), high TOF (5047 h−1 at −1.1 V vs. RHE), and low overpotential (50 mV), significantly surpassing most reported single-atom catalysts. DFT calculations suggest a cooperative effect between the Fe and Cu atoms, accelerating charge transfer and adjusting the d-band center, thus decreasing the activation energy barriers for *COOH formation and *CO desorption (Fig. 22d–f).216 The influence of precisely controlled metal–metal interactions is a crucial need for improved performance and selectivity for CO2 electroreduction. To illustrate, the Ni2(dppm)2Cl3 cluster was subsequently incorporated into a ZIF-8-derived NDC matrix and pyrolysis in an Ar atmosphere produced the Ni2/NC catalyst. This approach yielded a uniform catalyst featuring two Ni1–N4 moieties sharing two nitrogen atoms. The Ni2/NC demonstrates a CO FE of 94.3% at a current density of 150 mA cm−2, along with high stability. Operando synchrotron X-ray absorption fine structure (XAFS) and FTIR spectroscopy revealed a dynamically changing structure, including oxygen-bridge formation at negative potentials resulting in a stabilized and unusually highly active O–Ni2–N6 structure (Fig. 23a–c). The dynamically evolving site facilitates CO2RR efficiency under operating electrochemical conditions. The proposed reaction mechanism suggests that this specific, oxygen-bridged Ni2–N6 site structure strongly influences the kinetic rate-limiting step of *COOH formation.217 Similarly, it was anticipated that the incorporation of another iron can tune the electronic structure via orbital coupling and possibly enhance the activity and selectivity. The common limitation of single Fe–N–C catalysts is strong CO binding, which limits the activity and leads to high overpotentials.
![]() | ||
Fig. 22 (a) Schematic illustration of Cu–Fe–N6–C synthesis. (b) Linear sweep voltammogram (LSV) of Cu–Fe–N6–C. (c) CO faradaic efficiency (FE) for Cu–Fe–N6–C at various potentials. Reproduced from ref. 215 with permission from Wiley, copyright 2021. (d) Schematic illustration of Fe/Cu–N–C synthesis. (e) LSVs of Fe/Cu–N–C catalysts. (f) CO FE for Fe/Cu–N–C at various potentials. Reproduced from ref. 216 with permission from Royal Society of Chemistry, copyright 2021. |
![]() | ||
Fig. 23 (a) Schematic of Ni2/NC synthesis. (b) LSVs of different Ni2/NC catalysts. (c) CO FE for Ni2/NC at various potentials. Reproduced from ref. 217 with permission from American Chemical Society, copyright 2021. (d) Pictorial representation of the synthesis of Fe2–N6–C–O. (e) LSV curve and FE of CO for Fe2–N6–C–O at various potentials. Reproduced from ref. 218 with permission from American Chemical Society, copyright 2022. (f) Pictorial representation of the synthesis of InNi DS/NC. (g) LSV curve of InNi DS/NC catalyst. (h) CO FE for InNi DS/NC at various potentials. Reproduced from ref. 124 with permission from Wiley, copyright 2023. (i) Schematic of Fe1–Ni1–N–C synthesis. (j) LSVs of different Fe1–Ni1–N–C catalyst. (k) FE of CO for Fe1–Ni1–N–C at various potentials. Reproduced from ref. 221 with permission from American Chemical Society, copyright 2021. |
To validate, a series of electrocatalysts (Fe1–N4–C and Fe2–N6–C) with well-defined Fe single-atom and dual-atom sites were synthesized using a template-assisted approach by pyrolyzing at Ar/H2 atmospheres to control site structures. The Fe2–N6–C dual-atom catalyst exhibits a significantly higher TOF of 26637 h−1 and improved durability compared to the Fe1–N4–C SACs. The FE for CO production exceeds 80% over a wide potential range. Detailed experimental and theoretical analysis reveal that the orbital coupling between the two Fe atoms in the Fe2–N6–C catalyst reduces the energy gap between antibonding and bonding states in CO adsorption, thus promoting CO2RR activity and facilitating CO desorption (Fig. 23d–f).218 Likewise, precisely constructed homonuclear Fe2N6 diatomic sites anchored on NDPC were synthesized via pyrolysis of a Fe2(CO)9-containing ZIF-8 precursor. The Fe2N6 diatomic catalyst exhibits high CO FE of up to 96% at −0.6 V vs. RHE and a remarkably low Tafel slope of 60 mV dec−1. This surpasses the performance of single atom FeN4 catalysts which suffer from typically slow kinetics associated with sluggish CO desorption steps. DFT calculations show a significantly lower energy barrier for CO2 activation and CO desorption on Fe2N6 compared to FeN4 sites.219
Tuning the coordination motif always influences the CO2RR by exploiting the interatomic electronegativity offset within a dual atom catalyst (DAC). For instance, CuNi–DSA/CNF DSA catalyst, featuring atomically dispersed Cu and Ni sites supported on PVP electrospun carbon nanofibers were synthesized and compared the performance to monometallic counterparts. DFT calculations reveal that the electronegativity difference between Cu and Ni subtly modulates the electronic structure of the active sites. This was experimentally verified and observed high CO FE of 99.6% across a broad potential window (−0.78 to −1.18 V vs. RHE), a high TOF of 2870 h−1, and exceptional long-term stability. The findings establish the interatomic electronegativity offset as a powerful design principle for highly selective and efficient CO2RR electrocatalysts, offering valuable insights into catalyst design for sustainable energy conversion.220 Similarly, the rational designing of O-bridged bimetallic sites significantly enhances the performance of single-site catalysts for CO2 reduction. This approach provides key mechanistic insights and offers potential routes for high efficiency. For instance, an oxygen-bridged indium-nickel (In–O–Ni) dual-SACs supported on nitrogen doped carbon (InNi DS/NC) was successfully synthesized and In and Ni atoms decorated in a unique proximal arrangement, featuring an axial oxygen atom bridging the In and Ni centers (Fig. 23g–i). The oxygen bridge (In–O–Ni unit) was suggested as the crucial one to the catalyst's performance. The catalyst InNi DS/NC exhibits FE for CO production of >90% across a broad potential range (−0.5 to −0.8 V vs. RHE), reaching a maximum of 96.7% at −0.7 V vs. RHE with a high jCO of 317.2 mA cm−2 at −1.0 V vs. RHE (in a flow cell). DFT calculations disclose that the electronic modification arising from the In–O–Ni DSA optimizes the adsorption energy of *COOH intermediate, by lowering the activation energy barrier for CO production. This effect is further enhanced by a reduction in the *COOH formation energy and enhanced CO desorption relative to their monometallic analogues.124
The importance of neighboring single-atom interactions in electrocatalysis, offering a new design strategy for multi-metal SACs. Most SACs are designed with isolated metal centers, neglecting possible long-range interactions between neighboring single atoms that could modulate catalytic performance. Inspired by metalloenzymes, where adjacent metal sites cooperate in substrate activation, this study by Jiao et al. explores the non-bonding interactions between Fe and Ni single atoms in a NDC (Fe1–Ni1–N–C) catalyst for CO2 electroreduction. Fe1–Ni1–N–C was synthesized through a Zn-assisted atomization strategy, using the direct annealing of MOFs incorporating Fe and Ni-doped ZnO NPs (Fig. 23j–l). The Fe1–Ni1–N–C catalyst demonstrated a remarkably high CO FE of 96.2% at −0.5 V, significantly outperforming Fe1–N–C and Ni1–N–C catalysts composed of isolated Fe or Ni single atoms. The superior activity of Fe1–Ni1–N–C is attributed to a synergistic non-bonding interaction between the adjacent Fe and Ni single-atom pairs. DFT calculations reveal that this interaction significantly activates the Fe atoms to promote the formation of the COOH* intermediate.221
Tuning the morphology of the precursor is considered one of the most effective synthetic strategies for developing highly efficient and selective electrocatalysts. For instance, interconnected Ni–N-doped porous carbon (NiNPIC) materials were synthesized using Zn–Ni bimetallic MOFs via a simple and rapid ultrasonic-assisted synthesis method. The approach enabled precise control over the amount of Ni incorporated into the final catalyst. The NiNPIC catalysts demonstrated a high CO FE of 95.1% with CO current density of 10.2 mA cm−2.222 Traditional methods for synthesizing carbon-based SACs using electric heating are time-consuming, energy-intensive, and require inert gas protection. These methods also suffer from the potential aggregation of single atoms during the extended heating period. The researchers used a microwave-assisted rapid pyrolysis method to overcome the drawbacks of conventional electric heating. They used a Ni-doped ZIF (Ni–ZIF-8) as a precursor, and the ZnCl2/KCl mixture acted as a microwave absorber, dramatically reducing pyrolysis time to just 3 minutes while ensuring complete carbonization without needing inert gas protection. The KCl concentration influenced the catalyst's morphology, composition, and electrocatalytic performance. The optimized Ni1-N-C catalyst (Ni1–N–C–50, produced using a 50:
1 KCl/ZnCl2 ratio) displayed exceptional activity and selectivity for CO2 electroreduction with high FE for CO (up to 96%).223 The catalysts with high micropores and mesopores generated during pyrolysis enhance CO2 adsorption, and mass transport and prevent bubble formation in the electrolyte, both improving overall performance. The researchers designed a hierarchical nanoporous Ni–N–C catalyst obtained from a porphyrin-based zirconium MOF (PCN-222) synthesized from TCPPNi [TCPP = meso-tetra(4-carboxyphenyl)porphyrin] and ZrOCl2 (Fig. 24a and b). The Ni20–N–C catalyst delivered high FE for (∼99%) at −0.6 V vs. RHE and partial current densities for CO in 2 M KHCO3 (200 mA cm−2 at −0.30 V vs. RHE, highlighting the enhancement enabled by the improved mass transfer.224 The principles of differing diffusion rates are leveraged in targeted ways to modify the structure and improve performance in SACs. The Kirkendall effect is employed indirectly during the synthesis process by introducing a sacrificial template material. During high-temperature pyrolysis, the more mobile elements diffuse out, creating voids, porosity and potentially unique architectures within the solid framework and the less mobile element is left behind. The resultant material will have tailored porosity and offer improved catalyst dispersion that simultaneously influences reactant diffusion and mass transport. The researchers synthesized a multicomponent catalyst, NiMn–N–C, by introducing a trace amount of Mn into a Ni–N–C catalyst precursor derived from a MOF (Fig. 24f–h). The addition of Mn, even in trace amounts, triggers the Kirkendall effect during the pyrolysis step of catalyst synthesis and causes significant alterations in the catalyst's morphology and changes in surface chemical composition. The optimized NiMn–N–C catalyst exhibits a high CO FE of over 90% across a significant range of applied potentials, outperforming simpler, standard non-Mn catalysts.225 Although MOFs can be employed as an excellent precursor for the synthesis of SACs, most of the active sites are buried inside the carbon matrix, which hinders the accessibility of active sites. To overcome this issue, Xi Chen et al. introduced large mesopores into the nano frames to increase the availability of single-atom sites and also boost mass and charge transports. The mesoporous carbon nano frames embedded with atomically dispersed catalyst FeSAs/CNF-900 from Fe-doped MOFs demonstrated remarkable performance and exhibited a high CO FE of 86.9% at −0.47 V vs. RHE and a CO partial current density of 3.7 mA cm−2. Also, it was confirmed that the Fe centers predominantly exist in a porphyrinic Fe–N4 coordination environment, which optimally modulates electronic properties and enhances *COOH intermediate stabilization (Fig. 24c–e).226 Precise morphological control of MOF is crucial, offering potential avenues for the highly targeted design and fabrication of efficient electrocatalysts for CO2RR. For instance, by tuning the MOF precursor morphology (controlled through the addition of different salts during synthesis), the relationship between the morphology and composition of a multicomponent copper-based nanoreactor catalyst was explored by Wanfeng Xiong et al. Through the pyrolysis of MOF precursors with different structures (cuboctahedron, cube, truncated octahedron, and octahedron), the researchers synthesized copper-based nanoreactors with varied structures and compositions and evaluated their performance in electrocatalytic CO2 reduction. Characterization reveals that CHK–COCTA possesses a unique combination of Cu–N4 single-atom sites, Cu2O nanoparticles, and metallic copper (Cu). These multiple components act synergistically as tandem catalytic sites. It was proposed that CO2 is initially converted to CO at the Cu–N4 sites, and subsequently, CO is converted to deep reduction products as denoted in eqn 4 and 5 (primarily CH4, and lesser amounts of C2H4) at the Cu2O/Cu sites.227
CO2 + 8H+ + 8e− → CH4 + 2H2O 0.17 V vs. RHE | (4) |
2CO2 + 12H+ + 12e− → C2H4 + 4H2O 0.08 V vs. RHE | (5) |
![]() | ||
Fig. 24 (a) Illustration of Nix–N–C catalyst fabrication. (b) Total current density and CO faradaic efficiency (FE) vs. potential for Ni20–N–C, Ni100–N–C, and Ni–AB catalysts. Reproduced from ref. 224 with permission from American Chemical Society, copyright 2020. (c) Schematic illustration of FeSAs/CNF-900 formation. (d) Linear sweep voltammograms (LSVs) of FeSAs/CNF-900. (e) CO partial current densities for ZnxFey–C–T catalysts at various potentials. Reproduced from ref. 226 with permission from Elsevier, copyright 2020. (f) Schematic illustration of NiMn–N–C synthesis. (g) LSVs of NiMn–N–C. (h) CO FEs of NiMn–N–C. Reproduced from ref. 225 with permission from Royal Society of Chemistry, copyright 2025. (i) Preparation of Ni SACs with urchin-like hierarchical architecture (Ni/HH), nanosheets (Ni/NS), nanorods (Ni/NR), and solid cubes (Ni/SC). (j) CO partial current densities (jCO) of Ni SACs. (k) CO FEs of Ni SACs. Reproduced from ref. 228 with permission from Wiley, copyright 2023. |
![]() | ||
Fig. 25 (a) Schematic of mesoNC-Fe synthesis from ZIF-8-Fe@SiO2. (b) CO partial current density for mesoNC–Fe. (c) CO faradaic efficiency (FE) for mesoNC–Fe at various potentials. Reproduced from ref. 230 with permission from Elsevier, copyright 2019. (d) Schematic of ZnO@ZIF-NiZn core–shell nanorod synthesis and Ni/NCNT fabrication. (e) Linear sweep voltammograms (LSVs) of Ni/NCNTs. (f) CO FE for Ni/NCNTs at various potentials. Reproduced from ref. 232 with permission from Elsevier, copyright 2020. (g) Schematic of NiNC3@Cu2O fabrication. (h) LSVs of NiNC3@Cu2O. (i) CO FE for NiNC3@Cu2O at various potentials. Reproduced from ref. 233 with permission from Royal Society of Chemistry, copyright 2025. |
The strategic combination and engineering of distinct catalytic centers can significantly enhance CO2RR performance, enabling the production of highly valuable multi-carbon products. The researchers synthesized a Ni–NC3@Cu2O core–shell nanoreactor where single-atom Ni–NC3 is partially or fully encapsulated by electrodeposited Cu2O. Ni–NC3 was synthesized through the carbonization of a metal-azolate framework (MAF), followed by the deposition of a Cu2O nanoparticle shell using a simple, controllable electrochemical deposition process (Fig. 25g–i). Ni–NC3@Cu2O-10 catalyst exhibited an impressively high FE for C2 products, achieving a FE for ethanol of 55.5% at −1.0 V vs. RHE and reaching as high as 60.6% under a slightly stronger driving force (−1.4 V vs. RHE), surpassing many previous reports. The Ni–NC3 core component provides CO intermediates which then undergo further reaction over Cu2O where C–C coupling takes place.233 The successful creation of a novel tandem catalytic system demonstrates exceptionally high efficiency and an impressive degree of control over product selectivity to maximize specific outputs. The author synthesized a tandem catalyst by incorporating Ni SAs into ZIF-8 through an ion exchange method, resulting in a highly uniform distribution of single-atom nickel active centers throughout the carbon matrix. Subsequently, Cu NPs and Ni@ZIF-8 materials were physically mixed and subjected to pyrolysis at 900 °C under an inert Ar atmosphere, followed by acid-washing. This process produced a high density of catalytic centers within this advanced composite material. The catalyst, Ni SACs–Cu NPs, consists of Ni SACs encapsulated within copper nanoparticles (Cu NPs) predominantly exhibiting the (111) crystal facet. This unique design enables efficient and selective CO2 reduction through cooperative catalysis, where Ni SACs generate CO, which subsequently undergoes C–C coupling reactions on the Cu NPs, leading to the production of acetate (60%) and/or syngas with highly tunable CO/H2 ratio.234
CO2 + 6H+ + 6e− → CH3OH + H2O −0.39 V vs. NHE | (6) |
![]() | ||
Fig. 26 (a) Schematic of NiSA/PCFM catalyst synthesis via electrospinning. (b) CO FE for NiSA/PCFM. (c) Partial current densities for NiSA/PCFM at various potentials. Reproduced from ref. 235 with permission from Nature (CC-BY 4.0), copyright 2020. (d) Synthesis and images of flexible CuSAs/TCNFs membranes. (e) Linear sweep voltammograms (LSVs) of CuSAs/TCNFs membranes. (f) CO FE for CuSAs/TCNFs membranes at various potentials. Reproduced from ref. 236 with permission from American Chemical Society, copyright 2019. |
While the role of cations, particularly K+, in influencing CO2RR activity and selectivity is widely acknowledged, their precise mechanistic influence remains debated. Previous studies using Au-based catalysts yielded ambiguous results because the rate-determining step (RDS) in CO2RR on Au is electron transfer, which is unlikely to be directly impacted by K+. This research utilizes a highly symmetrical Ni–N4 structure, largely inert in both CO2RR and HER, allowing the isolation of K+ effects and providing a clear platform for investigating K+ precise mechanistic role (Fig. 27a–c). In the absence of K+, the Ni–N4 catalyst exhibited little to no CO2RR activity, with HER being the dominant reaction. However, upon adding K+, the CO2RR activity increased dramatically. A significant shift in the RDS occurred: from a concerted proton–electron transfer (CPET) in the absence of K+ to a predominantly independent proton transfer when K+ is present, lowering the overall reaction barrier and achieving near 100% CO FE at high current densities. In situ XANES measurements revealed that K+ interacts non-covalently with chemisorbed CO2, changing the bond geometry of CO2 and its vibrational behaviour. These in situ findings directly illustrate how K+ stabilizes chemisorbed CO2. Grand canonical potential kinetics (GCP-K) simulations supported the experimental results showing a lower overall energy barrier for the K+ influenced reaction pathway and a much lower onset potential (−0.524 V) compared to reported results using other Ni catalysts and far outperforming the HER. This study offers definitive experimental and theoretical evidence that K+ facilitates the reaction by forming a non-covalent interaction with chemisorbed CO2, stabilizing the chemisorbed intermediate, lowering the activation energy, and thus achieving highly efficient and selective CO2 reduction. The findings directly reconcile experimental and theoretical studies on CO2RR and address long-standing questions concerning the mechanistic role of cations in enhancing electrocatalytic activity.241 Similarly, the effect of strong concentration on Ni–N–C electrocatalysts for CO2 reduction in acidic media in strongly acidic conditions was explored. Hyewon Yun et al. demonstrated high concentrations of cations (specifically K+) synergistically enhance the CO2RR activity of Ni–N–C catalysts in acidic media (Fig. 27d–f). This effect surpasses simple additive contributions, showcasing a true synergistic interaction between the cations and protons. A shift from H2O to H3O+ as the primary proton source accelerates CO2RR kinetics at strongly acidic pH levels (<2). This is significant because acidic environments inherently favour protonation processes but hinder CO2 reduction due to enhanced HER competition and low CO2 solubility. It was demonstrated that leveraging the accelerated H3O+ protonation in high CO activity in acidic MEA resulted in nearly 95% FE for over 50 h at −100 mA cm−2. The enhanced activity of IL–Ni–NC under acidic conditions was significantly influenced by efficient control over the transport of protons and cations, resulting in overall superior catalytic performance. The importance of controlled proton delivery and reduced HER were demonstrated. In situ ATR-SEIRAS data reveals increased CO adsorption at higher K+ concentrations, especially at pH 1.7, which confirms the strong influence of cations in improving catalytic kinetics and shows distinctly varied interfacial water behaviour and local pH under different acidic conditions, explaining the role of H3O+ in accelerating reaction rates and stability in an acidic MEA under specific conditions. Furthermore, DFT calculations exposed lower free energy barriers for COOH formation under higher K+ concentrations due to significant H3O+ effects and further synergistic effects between cations and protons.242
![]() | ||
Fig. 27 (a) Schematic of Ni–N4–HM synthesis. (b) Linear sweep voltammograms of Ni–N4–HM in electrolytes of pH 3. (c) Proposed CO2RR mechanism showing the effect of K+ concentration on energy barriers for CO2RR and HER. Reproduced from ref. 241 with permission from Nature, copyright 2024. (d) Pictorial representation of K+ concentration effects in CO2RR using Ni–N–C catalyst. (e) CO current density of IL–Ni–NC and Ag/C catalysts with and without added K+. (f) CO faradaic efficiency (FE) for IL–Ni–NC and Ag/C at various potentials. Reproduced from ref. 242 with permission from Elsevier, copyright 2025. |
This comprehensive understanding of multifaceted factors will pave the way for the rational design of next-generation catalysts for sustainable CO2 conversion technologies. In summary, coordination tuning in MOF and COF based catalysts plays a pivotal role in optimizing the eCO2RR by precisely modulating the electronic structure, local charge distribution, and active site environment. By strategically adjusting metal ligand interactions, heteroatom doping, and substrate functionalization, these porous frameworks enhance catalytic activity, selectivity, and stability while facilitating efficient electron transfer and intermediate stabilization. Such tunability not only deepens the mechanistic understanding of CO2 electro reduction but also offers a rational design approach for next-generation catalysts, contributing to the development of sustainable carbon-neutral technologies.
![]() | ||
Fig. 28 (a) Schematic of artificial photosynthesis system. (b) Schematic illustration of the reaction mechanism. (c) Relative energy levels involved in semiconductor photocatalytic CO2 reduction. Reproduced from ref. 243 with permission from American Chemical Society, copyright 2017. |
Significant progress has been made in photocatalytic CO2 reduction research, with various semiconductor materials (TiO2, ZnO, CdS, etc.), MOFs, and single-atom catalysts being explored. However, improving the efficiency, selectivity, and long-term durability of these photocatalysts remains a key challenge in the field. Continued efforts in materials design, synthesis, characterization, and mechanistic understanding, along with advanced modeling, are expected to accelerate advancements and pave the way toward cost-effective, scalable technologies for solar fuel production and carbon mitigation.245,246 Coordination tuning plays a crucial role in optimizing the photocatalytic CO2 reduction performance of MOF and COF based catalysts by precisely modulating the electronic structure, charge distribution, and local environment of active sites. By adjusting metal–ligand coordination, heteroatom doping, and linker functionalization, these porous frameworks can enhance light absorption, facilitate charge separation, and stabilize key reaction intermediates. Such tunability not only improves catalytic efficiency and selectivity but also provides a platform for rational catalyst design aimed at sustainable CO2 conversion. Several MOF and COF-based materials have been explored as photocatalysts for CO2RR. For instance, 2,2′-bipyridine-based COF containing single Ni sites (Ni–TpBpy) was investigated as a photocatalyst for CO2RR (Fig. 29a–c). Ni–TpBpy efficiently enhanced CO selectivity in aqueous media, achieving a yield of 4057 μmol g−1 over a 5-hour reaction with 96% selectivity.246 Similarly, a CoN4Cl2 single site was fabricated by loading Co species into a 2,2′-bipyridine and triazine-containing COF. The TPy-COF-Co displayed a high CO production rate (426 mmol g−1 h−1) turnover number (2095) and TOF (1607 h−1). Theoretical calculation suggested the enhanced charge transfer from triazine-based COF framework to CoN4Cl2 sites facilitated CO2 activation by lowering the energy barrier of *COOH generation (Fig. 29d–f).247 In contrast, tuning the coordination number altered the product distribution significantly. For example, a 2,2′-bipyridine-based COF (TpBpy) with MoN2 sites (Mo-COF) efficiently catalyzed CO2 photoreduction to C2H4 with a selectivity of 32.92% along with CH4. This coordination tuning strategy in other catalytic systems opens new possibilities for photocatalytic CO2 reduction to higher carbon products (Fig. 29j–l).128 Similarly, single-atomic COF composed of different types of coordinated nitrogen atoms, such as imine, pyridine, and pyrazine N, were constructed to study their effects on electronic structure and photocatalytic CO2 reduction. The imine and pyridine N atoms coordinated single-atomic Co COF exhibited a CO evolution rate of 85.5 μmol h−1 with 96.4% selectivity in the presence of Ru photosensitizer, outperforming pyrazine N atoms coordinated Co sites (24.3 μmol h−1, 65.5% selectivity) (Fig. 29g–i).248 Based on the polar coordination microenvironment, an effective Ni-SAs anchored on N, O sites of COF was synthesized using an atomic deposition method, as shown in (Fig. 30a), and applied for efficient photocatalytic CO2RR.249 To understand the band structure of the photocatalyst, ultraviolet photoelectron spectroscopy (UPS) analysis was performed on 1.0% Ni-tp-COF (Fig. 30b). The valence band (VB) value was determined as 1.00 V, calculated from the cutoff energies (Ecutoff) of 16.82 eV and onset energies (Eonset) of 1.04 eV. The VB potential was calculated using the formula EVB = 21.22 − (Ecutoff − Eonset), resulting in a VB potential of 5.44 eV, which was then referenced to the reversible hydrogen electrode VB potential −4.44 eV. Furthermore, the band gap energy (Eg) for1.0% Ni-tp-COF was found to be 1.99 eV, closely matching that of pristine tp-COF 2.01 eV (Fig. 30c). Moreover, 1.0% Ni-tp-COF catalyst exhibited a positive slope in the Mott-Schottky plots, confirming their n-type semiconductor characteristics (Fig. 30d). For n-type semiconductors, the conduction band (CB) potential can be approximated to the flat band potential. The CB potentials, measured against Ag/AgCl, were found to be −1.02, and −1.24 V, corresponding to −0.82, −1.04 V versus the normal hydrogen electrode (NHE), respectively. Based on these values and the bandgap energies, the valence band (VB) potentials were calculated using the relation EVB = ECB + Eg yielding VB positions of 1.19, 0.95 0.86 V versus NHE (Fig. 30e). Notably, the CB potentials of all samples are more negative than the reduction potential of CO2 to CO, indicating a strong thermodynamic driving force for the CO2 reduction reaction (CO2RR) and demonstrating the excellent reductive capability of these photocatalysts. The 1.0% Ni-tp-COF catalyst produce with a yield of 15.0 μmol g−1 h−1 and highly selective conversion of CO2 to CO (>99%) (Fig. 30f). Under visible light irradiation, photogenerated electrons and holes are produced within the photocatalyst. These charge carriers drive the adsorption of CO2 molecules onto the surface of catalytic active sites, forming adsorbed *CO2 intermediates. Afterward, the *CO2 species undergo a series of reduction steps: a one-electron transfer leads to the formation of *COOH intermediates, followed by a two-electron transfer resulting in the generation of *CO species. Finally, the *CO desorbs from the catalyst surface, leading to the production of CO gas (Fig. 30g). This sequence of adsorption, stepwise reduction, and desorption constitutes the fundamental mechanism underlying the photocatalytic valorisation of CO2. Beyond the intrinsic catalytic sites in metal nodes and organic linkers, MOFs have also been extensively utilized as hosts for confining or supporting guest catalytic sites, leveraging their high surface area and well-defined porous structures. Additionally, MOFs can act as tailored microenvironments, influencing the catalytic behaviour and reactivity of encapsulated species.
![]() | ||
Fig. 29 (a) Schematic of CO2 photocatalytic reduction using Ni–TpBpy. (b) and (c) CO and H2 evolution by Ni–TpBpy under 1 and 0.1 atm CO2. (d) CO2 reduction yields (2 h reaction) using TpBpy with different metal ions. Reproduced from ref. 246 with permission from American Chemical Society, copyright 2019. (d) Synthesis of TPy-COF and BPy-COF for the fabrication of TPy-COF-Co and BPy-COF-Co. (e) Time-resolved generation of Gas (CO and H2) over TPyCOF-Co. (f) Gas production rates in the CO2 photoreduction over TPy-COF-Co, BPy-COF-Co, and TBp-COF-Co. Reproduced from ref. 247 with permission from Wiley, copyright 2025. (g) Synthesis of COFs with different N species (COF-Ph, COF-Py, COF-Pz) and their Co-coordinated counterparts (COF-Ph-Co, COF-Py-Co, COF-Pz-Co). (h) Photocatalytic CO2 reduction activities (3 h) over various catalysts. (i) Time-dependent CO and H2 evolution during photocatalytic CO2 reduction by [Ru]-photosensitized COF-Py-Co. Reproduced from ref. 248 with permission from Elsevier, copyright 2025. (j) Pictorial representation of the synthesis of Mo-COF. (k) Photocatalytic evolutions of CO, C2H4 and CH4 by Mo-COF. (l) Time-resolved generation of Gas (CO, C2H4, and CH4) over Mo-COF. Reproduced from ref. 128 with permission from Elsevier, copyright 2021. |
![]() | ||
Fig. 30 (a) Schematic presentation of synthesis of Ni anchored COF. (b) UPS spectra of 1.0% Ni-tp-COF. (c) Tauc plots of tp-COF and Ni-tp-COF catalysts. (d) Mott-Schottky plot of 1.0% Ni-tp-COF. (e) Band structures tp-COF and Ni-tp-COF catalysts. (f) Photocatalytic CO2 reduction, CO production, and the selectivity performance of the photocatalysts. (g) Mechanism of photocatalytic CO2 reduction using 1.0% Ni-tp-COF under visible light. Reproduced from ref. 249 with permission from Elsevier, copyright 2025. |
MOF-derived materials, owing to their structural versatility and compositional tunability, have garnered significant attention as electrocatalysts for CO2 reduction. Notably, heteroatom-doped carbon materials derived from MOFs provide a straightforward and effective strategy for designing high-performance CO2RR catalysts. A MOF assisted synthetic approach for generating high-performance atomic pair photocatalysts for efficient CO2 photoreduction to CO is considered highly promising. In particular, the synergistic effects of carefully tailored dual atomic catalytic sites enhance performance by modifying reaction kinetics while significantly boosting the CO2RR selectivity. For instance, Co1Cu1/NC was synthesized using a MOF-assisted strategy involving the pyrolysis of a ZIF-8 precursor, in which copper nitrate ions were incorporated into the cavities, and cobalt-tetrabenzoporphyrin (CoTBPP) was anchored onto the external surface of ZIF-8. Notably, characterization data divulge the 4N-coordinated metal atom (N2–Co–N2–Cu–N2) in Co1Cu1/NC (Fig. 31a–c). The Co1Cu1/NC photocatalyst, exhibits significantly enhanced CO2 reduction activity (22.46 mmol g−1) in conjunction with [Ru(bpy)3]2+ as a photosensitizer, and a CO selectivity of 83.4% after 2 hours of visible-light irradiation. The excellent photocatalytic activity of Co1Cu1/NC is attributed to the formation of the *COOH intermediate, facilitated by a mechanism in which CO2 activation and reduction are enabled through synergistic effects between neighbouring Co–N4 and Cu–N4.250 Understanding the impact of coordination number in SACs photocatalytic CO2RR is particularly intriguing, as it can influence the binding affinity of reactive intermediates and active sites at molecular level. For example, a series of Co single atom photocatalysts, CoSA–Nx/C, with varying nitrogen coordination numbers, were synthesized, and their CO2 photoreduction properties were investigated. The CoSA–N2/C photocatalyst exhibited superior CO2 photoconversion activity and CO selectivity (82.6%) compared to CoSA–N3/C and CoSA–N4/C. Additionally, it demonstrated a high CO yield rate of 10110 μmol g−1 h−1 with outstanding stability.
![]() | ||
Fig. 31 (a)–(c) Synthesis, CO/H2 yields (2 h visible light irradiation, λ > 420 nm), and CO selectivity of Co1Cu1/NC with varying dual-metal site coordination environments. Reproduced from ref. 250 with permission from American Chemical Society, copyright 2023. (d)–(f) CO2 photoreduction activity (CO/H2 yield rates) for CoSA–Nx/C under various conditions, compared to other cobalt catalysts (2 h visible light irradiation). Reproduced from ref. 251 with permission from Elsevier, copyright 2024. (g)–(i) Fe−N−O/NC synthesis, CO2 photoreduction activity with different Fe loadings, and turnover numbers (TONs) (2 h visible light irradiation). Reproduced from ref. 252 with permission from American Chemical Society, copyright 2020. (j) and (k) Ni−SACs@BNC synthesis and photocatalytic CO2R activity (3 h irradiation under varying conditions). Reproduced from ref. 195 with permission from American Chemical Society, copyright 2025. |
The activity enhancement of CoSA–N2/C photocatalyst was ascribed to increased CO2 adsorption due to the unoccupied Co 3d electronic orbitals and a reduction in *COOH intermediate energy barrier (Fig. 31d–f). Thus, the regulation of local coordination structure and its effect in the CO2 photo reduction was emphasized.251 Similarly, the breaking symmetry of SACs could boost performance and tune selectivity. For instance, Ni–SACs@BNC was synthesized and the Ni–SACs@BNC catalyst exhibited high photocatalytic CO2 reduction activity and selectivity, achieving a CO evolution rate exceeding that of Ni–SACs@NC and the analogous control catalyst (N–C), reaching as high as 93.2% under optimized reaction conditions for 3 hours under visible light (Fig. 31j–k).195 Similarly, SAC symmetry was further tuned using an axial coordination strategy. For example, atomically precise Fe–N4O active sites were generated via a facile, low-temperature (500 °C) top-down strategy, which involved high-temperature pyrolysis (500 °C) under Ar to achieve in situ direct atomization of Fe particles in the presence of NH4Cl, significantly increasing the CO content on the catalyst support. A clear positive correlation was found between C
O concentration, catalyst activity, and reaction selectivity. CO2 photoreduction experiments using a Ru(bpy)3Cl2 photosensitizer demonstrated remarkably enhanced performance parameters and established both catalyst selectivity and activity. Fe–NO/NC achieved a high TON of 1494 in 1 hour for CO generation with 86.7% selectivity, demonstrating remarkable stability across multiple test cycles with consistently high performance (Fig. 31g–i).252 The development of photocatalysts with hierarchical pores could alter the mass transport properties and electrical surface characteristics, potentially enhancing the efficiency overall photocatalytic efficiency. For instance, a hierarchical porous single atomic cobalt photocatalyst containing highly distributed, isolated Co single atoms on micro and mesoporous NDC (0.8-Co–ISAS/MMNC-900) was developed and used as a catalyst for the photoconversion of CO2 to CO. The catalyst exhibited a high CO formation rate (7261 μmol g−1 h−1) with 90.1% selectivity. DFT calculation revealed that the porous carbon structure encouraged the development of an exterior local electric field, which enhanced charge separation and enabled the accumulation of photo-excited electrons for effective CO2 adsorption and *COOH intermediate formation. This unique strategy facilitates improved photocatalytic CO2RR via multiscale morphology engineering.253 Moreover, to understand the cooperative effect between SAs and NPs on the photochemical valorisation of CO2, a simple solvothermal method followed by impregnation was used to design a Co-SAs/Al-bpydc/Ni-NPs catalyst (Fig. 32a).254 SEM, HAADF–STEM, and elemental mapping images (Fig. 32b–g) clearly illustrated the presence of Co-SAs and Ni-NPs. In this photocatalytic system under visible light (λ ≥ 420 nm), the enormously dispersed Co SAs effectively adsorb and activate CO2 molecules, while the Ni NPs facilitate protons transfer to Co SAs, leading to a highly selective conversion of CO2 to CO (91%) with a yield of 12.8 mmol g−1 h−1 (Fig. 32i). The Co-SAs/Al-bpydc/Ni-NPs catalyst exhibits a higher photocurrent response than its counterparts, attributed to the excellent migration photogenerated electron–hole pairs (Fig. 32j). In addition to the experimental results, DFT calculations revels that the synergistic effect between Co SAs and Ni NPs lowers the energy barrier for CO2 RR and enhances CO selectivity compared to individual Co SAs and Ni NPs (Fig. 32k), in the presence of a photosensitiser (Ru complex) and sacrificial agent (triethanolamine). In summary, coordination tuning in MOF and COF based catalysts serves as a powerful strategy to enhance the efficiency and selectivity of photocatalytic CO2 reduction.255–260 By precisely modulating metal–ligand interactions, electronic structures, and surface properties, these materials facilitate optimal charge separation, stabilize key intermediates, and promote desired reaction pathways. The tunability of these frameworks not only advances the fundamental understanding of photocatalysis but also paves the way for the rational design of next-generation catalysts for sustainable CO2 conversion.
![]() | ||
Fig. 32 (a) Synthesis of the Co-SAs/Al-bpydc/Ni-NPs catalyst. (b)–(g) SEM, HAADF STEM, and elemental mapping images of Co-SAs/Al-bpydc/Ni-NPs. (h) CO selectivity as a function of Co SAs loading. (i) CO production rate and selectivity of the Co-SAs/Al-bpydc/Ni-NPs catalyst compared to its counterparts. (j) Photocurrent response of the Co-SAs/Al-bpydc/Ni-NPs catalyst and its counterparts. (k) Schematic representation of Ni-NPs/Al-bypdc, Co-SAs/Al-bpydc/Ni-NPs, and Co-SAs/Al-bypdc photocatalytic CO2 RR (orange and green spheres Co, Ni). Reproduced from ref. 254 with permission from American Chemical Society, copyright 2024. |
Investigating possible synergies between atomic-level active sites during design optimization to enhance specific reaction steps without compromising other catalytic activities presents an exciting avenue for further research in catalytic design. Consequently, Hu et al. synthesized an effective catalyst (CuZnNx@C–N) made up of Zn single atoms and Cu clusters stabilized by nitrogen sites on a carbon skeleton using ZIF-8 as the precursor material.267 The synthesis method is illustrated in Fig. 33a. Upon high-temperature calcination, the MOF structure decomposes, leading to the high dispersion of Zn, which interacts with the N atoms and functions as a single atom site, confirming its atomic dispersion (Fig. 33b). However, the EXAFS spectrum of the CuZnNx@C–N catalyst exhibits a stronger signal intensity at the Cu–N bond position, while the Cu–Cu peak intensity is also pronounced, (Fig. 33c), indicating that the Cu species in the calcined sample exist as nanoparticles comprising Cu0 and CuNx sites, which are important for hydrogen dissociation.
![]() | ||
Fig. 33 (a) Schematic representation for the synthesis of CuZnNx@C–N catalyst. (b) EXAFS of CuOx@ZIF-8, CuZnNx@C–N (700 °C), Zn foil and ZnO. (c) EXAFS of CuOx@ZIF-8, CuZnNx@C–N(700 °C), Cu foil and CuO. (d) Hydrophobic angle measurement. (e) Influence of calcination temperature on RWGS performance. (f) The stability test of CuZnNx@C–N catalyst in RWGS reaction. (g)–(i) The catalytic performance of the CuZnNx@C–N catalyst calcined at 700 °C. Reproduced from ref. 267 with permission from Wiley, copyright 2023. |
Initially, CuOx nanoparticles were encapsulated in ZIF-8 as guest particles. Upon calcination, Cu atoms underwent redistribution due to their coordination with N atoms at high temperatures. Notably, since pyridinic N and pyrrolic N more effectively anchor SAs than graphitic N due to geometric effects, pyridinic N is the predominant N species in the catalyst, alongside some pyrrolic N and a minor fraction of graphitic N. The M–N4 active site, coordinated by four N atoms, is believed to facilitate electron transport and promote the formation of key intermediates in CO2 reduction. In this context, the four-coordinated Zn–N4/C and Zn free N4/C structures exhibit stronger adsorption capacities and more active sites for *CO2 than pyridinic N alone (N/C). The N4/C structure, particularly the Zn containing variant, provides numerous adsorption sites for *COOH, favouring CO generation. Beyond, serving as anchoring sites, N atoms also induce Lewis basicity in neighbouring C atoms via electron pair induction in pyridinic N, making pyridinic N atoms a significant active site for CO2 reduction.268 Concurrently, CuNx is considered a principal active site for CO2 hydrogenation, as Cu sites efficiently adsorb and activate H2 molecules. Fig. 33d indicates the formation of ZIF-8 after the addition of CuOx, with the synthesized ZIF-8 maintains its original hydrophobicity. The study reports that the bifunctional catalyst, incorporating the Zn–N4 structure and the ultrasmall CuNx clusters, exhibits outstanding catalytic performance in CO2 hydrogenation, facilitating highly efficient CO generation under atmospheric pressure. Catalytic testing showed that the catalyst calcined at 500 °C exhibited limited activity for CO2 reduction due to insufficiently coordinated active sites at lower calcination temperatures. However, increasing the calcination temperature enhanced catalytic activity and selectivity Fig. 33e. In particular, the catalyst calcined at 700 °C achieved the highest catalytic activity, approaching the thermodynamic equilibrium conversion rate (49.9%) of CO2 under the given reaction conditions while maintaining long-term catalyst stability Fig. 33f. Conversely, the catalyst calcined at 800 °C exhibited reduced CO2 conversion due to a significant decline in specific surface area. Notably, the CO2 conversion rate of the CuZnNx@C–N catalyst calcined at 700 °C improved substantially at reaction temperatures exceeding 400 °C, approaching the thermodynamic equilibrium conversion rate Fig. 33g. Moreover, the CO selectivity of CuZnNx@C–N remained consistently above 99% across various reaction temperatures Fig. 31h. Additionally, the space-time of CO for all catalysts increased significantly with rising reaction temperature Fig. 33i, attributed to favourable reaction kinetics. These findings establish CuZnNx@C–N as one of the high-performance catalysts reported for CO2-to-CO thermochemical conversion.
Formic acid is an organic acid recognized as one of the most promising modern fuel sources for low-temperature proton–exchange membrane fuel cells, as well as a renewable hydrogen carrier. The current commercial process, which produces 800 kT (kilo ton) of formic acid annually, involves the use of methanol, toxic carbon monoxide, and corrosive sodium methoxide.269 Therefore, formic acid produced via CO2 hydrogenation represents an important biodegradable and sustainable feedstock for the future. Due to the high kinetic and thermodynamic stability of CO2, noble metal-based homogeneous catalysts are typically employed for formic acid synthesis. Recently, Fellenberg et al. developed ruthenium (Ru) single-atom sites densely packed on COF for CO2 hydrogenation into formic acid.270 They synthesized COF using various organic linkers to tune nitrogen density within the framework, thereby tailoring the Ru single atoms shown in Fig. 34a. Experimental data indicate that, at the same Ru content, the COF chemical composition plays a more significant role in determining the fraction of isolated SA Ru species. COFs enable ruthenium catalysts with extremely high densities of single-atom sites, leading to an improvement in TOF, which contributes significantly to the design of highly active catalysts. The hydrogenation reaction was conducted under relatively mild conditions, at a total pressure of 40 bar and temperatures of 90 °C and 120 °C, with H2/CO2 ratios of 1 and 3 in aqueous medium. Among the tested catalysts, 10% ruthenium anchored on a COF synthesized with benzene-1,3,5-tri carboxaldehyde (TFB) and hydrazine (Hz) (Ru10% TFB Hz) exhibited the highest TOF (243 h−1) at both 90 °C and 120 °C. In Ru10% COF TFB o-Tol, the amounts of Ru–N complexes and Ru hydroxide species increased under the reaction conditions, while the concentration of metallic Ru phases decreased. This suggests the redispersion of Ru species and an enhanced interaction between Ru complexes and N species in the COF during the reaction Fig. 34b. Conversely, a progressive increase in metallic ruthenium species was observed in Ru10% TFB-TAB (Fig. 34c and d), indicating the oxidation of Ru SA sites due to the absence of oxygen in the COF framework. This suggests that intrinsic catalytic activity (TOF) is strongly influenced by the nitrogen content in the catalysts. As the N/Ru ratio increased, overall intrinsic activity also increased Fig. 34e. However, while Ru10% TFB-TAB, and TFB-Hz were initially the most active catalysts, they lost activity in the second and third reaction cycles (Fig. 34f). In contrast, Ru10% TFBo-Tol maintained it's catalytic, which correlated with a decrease in the concentration of Ru SA site activity. The thermocatalytic hydrogenation of CO2 to alcohols has garnered significant interest, as both methanol and ethanol offer ease of transport and substantial value in chemical synthesis. Traditional CO2 hydrogenation using metal oxides requires harsh reaction conditions (>300°C, 1–5 MPa), leading to excessive energy consumption.271 Designing, a functionalized material that can simultaneously anchor metal SAs for hydrogen dissociation and enhance CO2 adsorption is an efficient strategy for thermocatalytic conversion. Recently, Rosado et al., rationally designed and synthesized a functionalized NU-1000 MOF that serves as a platform for the simultaneous coordination stabilization of Cu SAs and CO2 adsorption, enabling its effective conversion to methanol.272 By modifying the ligand before the MOF synthesis, NU-1000 was functionalized with primary amino groups (–NH2), which have a high affinity for CO2 adsorption, and free thiol functionalities (–SH), which exhibit strong interactions with Cu SAs. These functional groups were incorporated into the framework using post-synthetic solvent-assisted ligand incorporation. Under working conditions, the NU-1000-based Cu SACs exhibited a high methanol space-time yield (STY), surpassing that of the industrially important CuZnO/Al2O3 catalyst. Notably, the catalytic of the MOF-based catalyst increased with temperature, whereas the commercial catalyst showed reduced performance beyond 260 °C due to nanoparticles sintering. Moreover, the MOF catalyst demonstrated significant activity at temperatures as low as 180 °C, achieving an STY of 35 mg MeOH gcat−1 h−1, which steadily increased with temperature, reaching 103 mg MeOH gcat−1 h−1 at 280 °C. This enhancement was attributed to the favoured adsorption of CO2 at the strategically incorporated –NH2 sites and the high concentration of active copper sites dispersed as individual atoms along the MOF channels. While the thermocatalytic hydrogenation of CO2 to ethanol and higher alcohols is of great interest, large-scale production remains a significant challenge. Recently, Liu et al. developed a state-of-the-art catalyst for CO2 hydrogenation to ethanol by precisely depositing single-atom Ir species onto phosphorous cluster islands on In2O3 nanosheets. This catalyst achieved an impressive ethanol yield of 3.33 mmol g−1 h−1 and TOF of 914 h−1 under 1.0 MPa (H2/CO2 = 3
:
1) at 180 °C.262 Therefore, catalysts with high surface area and precisely controlled active sites for CO2 adsorption and molecular hydrogen dissociation are highly attractive. Single-atom decorated MOFs or COFs are promising candidates for thermocatalytic CO2 reduction in the near future. The top papers published in recent years on electrochemical, photochemical, and thermochemical CO2 valorisation are summarized in Table 3.
![]() | ||
Fig. 34 (a) COF synthesized with various molecular linkers with visual aspects. (b) and (c) Evolution of catalyst phase composition during the CO2 hydrogenation, Ru10% TFB o-Tol and Ru10% TFB TAB. (d) Evolution of in situ XAS spectra during the CO2 hydrogenation, Ru10% TFB-TAB. (e) TOFs measured on the Ru/COF catalysts at 90 °C and 120 °C as functions of N/Ru experimental molar ratio. (f) Stability of various catalysts. Reproduced from ref. 270 with permission from Elsevier, copyright 2024. |
Electrocatalyst | Electrolyte | Product | FE (%) | Potential (V) | Turn over frequencies (TOF-h−1) | Current density (mA cm−1) | Ref. |
---|---|---|---|---|---|---|---|
A–Ni–NSG | 0.5 M KHCO3 | CO | 97% | 0.61 | 14![]() |
22.0 | 128 |
Ni–N–C | 0.5 M KHCO3 | CO | 96.8 | −0.80 | 11![]() |
27.0 | 153 |
Fe3+–N–C | 0.5 M KHCO3 | CO | >80 | −0.20 | ∼1100 | 94.0 | 163 |
C–Zn1Ni4 ZIF-8 | 1 M KHCO3 | CO | 98 | −0.83 | 10![]() |
44.1 | 166 |
Ni–NBr–C | 0.5 M KHCO3 | CO | 97 | −0.70 | 35![]() |
350 | 189 |
InA/NC | 0.5 M KHCO3 | CO | 97.2 | −2.10 | ∼40![]() |
39.4 | 193 |
Al–NC | 0.1 M KHCO3 | CO | 98.76 | −0.65 | 12![]() |
330 | 196 |
Ni/Cu–N6–C | 0.5 M KHCO3 | CO | 97.7 | −0.60 | 20![]() |
>100 | 204 |
Ni–N3/Cu–N3 | 0.5 M KHCO3 | CO | 99.1 | −1.10 | 22![]() |
88.0 | 205 |
Zn-SA/CNCl-1000 | 1.0 M KOH | CO | 97 | −0.93 | 29![]() |
271.7 | 229 |
Photocatalytic | Condition | Light | Product | Selectivity (%) | TOF or TON | Yield | Ref. |
---|---|---|---|---|---|---|---|
Ni–TpBpy | (Acetonitrile, pure water, triethanolamine) | 300 W Xe lamp | CO | 96 (5 h) | 13.62 (5 h) | 4057 μmol g−1 | 237 |
Tpy–COF–Co | (Acetonitrile, deionized water, triethanolamine) | 300 W Xe lamp | CO | — | (TOF) 1607 h−1 and TON 2095 | 426 mmol g−1 h−1 | 238 |
Co1Cu1/NC | (Acetonitrile, pure water, triisopropanolamine) | 300 W Xe lamp | CO | 83.40 (2 h) | 59 | 22.46 mmol g−1 | 241 |
CoSA–Nx/C | (Acetonitrile, pure water, triisopropanolamine) | 300 W Xe lamp | CO | 82.60 | 98 (2 h) | 10![]() |
242 |
Fe–NO/NC | (Acetonitrile, deionized water, triethanolamine) | 300 W Xe lamp | CO | 86.70 | 1494 (1 h) | 81.8 μmol | 243 |
Co-COF | (Acetonitrile, [Ru(bpy)3Cl2]·6H2O, TEOA) | 300 W Xe lamp | CO | 95.70 | (TOF) 111.8 h−1 | 18![]() |
255 |
Fe SAS@Tr-COF | (Acetonitrile, water, [Ru(bpy)3Cl2]·6H2O, TEOA) | 300 W Xe lamp | CO | 96.40 | 2.89 | 980.3 μmol g−1 h−1 | 46 |
Cu-SA/CTF | (Triethanolamine, water) | CH4 | 98.31 | 24.05 (4 h) | 32.56 μmol g−1 h−1 | 256 | |
TCM-Bpy-COF-CoAC | (acetonitrile, water, [Ru(bpy)3Cl2]·6H2O, TEOA) | 5 W LED (λ = 400–800 nm) | CO | 81.80 | — | 26![]() |
257 |
Pt-SA/CTF-1 | (Triethanolamine, water) | 300 W Xe lamp | CH4 | 76.60 | — | — | 258 |
Fe@MIL-OV-300 | (Triethylamine, acetonitrile, water) | 300 W Xe lamp | CH3OH | — | (TOF) 16.03 h−1 | 15.85 mmol g−1 at 4 h | 259 |
NiSAs@NPs/TC | (Deionized water, CO2 gas) | 300 WXe lamp | CO, CH4 | — | — | 35.60 and 3.41 μmol g−1 h−1 | 260 |
Regime | Key methodological considerations | Main mechanistic descriptors | Typical DFT protocols/models | Ref. |
---|---|---|---|---|
Electrocatalysis | Electrode–electrolyte interface (implicit/explicit solvation) | Free energy profiles of intermediates | GGA, VASP/Gaussian/CP2K, VASPsol, CHE, Bader, NEB | 281 and 287–289 |
Applied potential corrections (CHE model) | Overpotential *COOH, *HCOO, *CO adsorption energies | |||
Charge transfer analysis (e.g., Bader charge, charge difference mapping) | Charge redistribution pathways | |||
Photocatalysis | Band structure & alignment (relative to redox levels) | Band gap, band edge positions | GGA, HSE06 (for better band gaps), periodic DFT, slab models | 283,284 |
Static charge separation (potential mapping) | Projected DOS | |||
Band edge calculations (vacuum alignment) | Charge localization/static separation | |||
(Carrier dynamics often beyond ground-state DFT: TDDFT, GW, BSE) | Surface reaction energetics | |||
Thermocatalysis | Transition state location (climbing NEB, dimer) Adsorption/desorption energy profiles | Reaction energy barriers | GGA, periodic DFT, TS searches, frequency calculations | 285,286 |
Entropic and zero-point energy corrections (vibrational analysis) | Turnover frequency (TOF, via microkinetics) | |||
Reaction barriers | Adsorbate binding energies | |||
Activation entropy & free energies |
DFT has become an indispensable computational tool for uncovering the mechanistic intricacies of CO2RR on SACs, particularly those derived from MOFs and COFs. By simulating the electronic structure, adsorption behavior, and reaction energetics at the atomic scale, DFT complements in situ techniques, bridging critical knowledge gaps regarding transition states, binding preferences, and energy barriers along the catalytic pathway. For instance, DFT studies have systematically evaluated the free energy profiles of key intermediates such as *COOH, *CO, and *CHO, offering valuable predictions regarding the preferred reaction pathways and product selectivity.290 Moreover, DFT helps elucidate the role of heteroatom coordination (e.g., N, B, or S) in tuning the local electronic environment of metal centers, thereby enhancing catalytic efficiency. For example, in the case of boron- and nitrogen-coordinated Sb SACs, DFT calculations have shown that pre-adsorbed hydrogen on boron sites suppresses the competitive hydrogen evolution, improving CO2 selectivity.291 Advanced DFT approaches, including ab initio molecular dynamics (AIMD), also provide further provide insights into catalyst stability under realistic electrochemical conditions, which is crucial for ensuring long-term performance in CO2RR. These computational insights help rationalize experimentally observed trends and guide the design of next-generation SACs by predicting optimal metal centers, coordination motifs, and electronic descriptors for enhanced catalytic selectivity and activity.
A compelling example illustrating the power of DFT for COF based SACs come from the investigation comparing CO2RR pathways on two metal-active centers, Ni and Ti, embedded in the aldehyde-amine COF (TAPT-Tp) framework.295 Structural models were constructed to simulate the adsorption of CO2 and the stepwise reduction to CO, with Gibbs free energy changes computed for each elementary reaction step (Fig. 35a and b). The DFT calculations revealed that Ni centers consistently exhibited lower energy barriers for critical intermediates (*COOH and *CO) than Ti sites (Fig. 35c). Notably, Ni-TAPT-Tp required only −0.07 eV for the *COOH formation step, compared to +0.61 eV for Ti-TAPT-Tp, indicating a more favourable proton–coupled electron transfer process on Ni. The structural coordination environment also played a decisive role. The four-coordinated, planar geometry of Ni-TAPT-Tp resulted in a more delocalized π-conjugated environment, which enhances light harvesting and charge mobility. The less sterically hindered coordination environment further facilitates CO2 adsorption and enables faster ligand exchange for CO desorption, thereby reducing the likelihood of active site poisoning. In contrast, the six-coordinated octahedral environment of Ti-TAPT-Tp restricts adsorption flexibility and increases desorption barriers, limiting catalytic efficiency. Such comprehensive modelling, which links geometric, electronic, and thermodynamic factors, provides a complete mechanistic map for designing optimal COF-SACs.
![]() | ||
Fig. 35 (a) Free energy diagrams for CO2RR on Ni-TAPT-Tp and Ti-TAPT-Tp COFs, highlighting key intermediates and energy barriers for each step. (b) Proposed CO2 reductionmechanism based on DFT insights. Reproduced from ref. 295 with permission from Wiley, copyright 2025. (c) Partial density of states showing Fe 3d orbital contributions in MOF-derived Fe–N–C catalysts. (d) Charge density difference plots illustrating electron redistribution before and after *CO adsorption. (e) pDOS of Fe-3d orbitals and adsorbed CO orbitals, indicating orbital coupling during CO2 reduction. (f) Schematic representation of Fe–CO orbital interactions comparing single- and dual-Fe sites, emphasizing their effect on *CO binding and desorption. (g) Comparison of free energy profiles highlighting the potential-determining steps for CO2 reduction at single versus dual-Fe atom active sites, (h) Theoretical comparison of limiting potentials for CO2RR and HER across various Fe–N–C configurations. Reproduced from ref. 296 with permission from American Chemical Society, copyright 2022. |
For MOF-derived SACs, DFT calculations have also provided transformative insights into the evolution from single-atom to dual-atom active sites, highlighting synergistic electronic interactions. In a study of Fe–N–C catalysts derived from Fe-doped MOFs, the partial density of states calculations were used to compare the electronic structures of Fe–N4–C (single-Fe) and Fe2–N6–C (dual-Fe) configurations (Fig. 35c).296 The results showed significant peak splitting and delocalization of Fe-3d orbitals in the dual-atom configuration, indicating strong electronic coupling between adjacent Fe atoms (Fig. 35c and d). This electronic delocalization shifted the d-band center of Fe to a more negative energy relative to the Fermi level (from −2.05 eV in Fe–N4–C to −3.00 eV in Fe2–N6–C–O), making *CO adsorption less favourable and thereby promoting its desorption, which is a common bottleneck in CO2RR (Fig. 35e and f). Charge density difference plots in Fig. 33d further illustrate enhanced electron transfer from Fe to adjacent N atoms, stabilizing the catalytic site and modulating intermediate binding. DFT-based crystal orbital Hamilton population analysis confirms that the Fe-CO is less covalent in Fe2–N6–C–O, resulting in a lower *CO desorption barrier relative to Fe–N4–C sites (Fig. 35g and h). The computed free energy diagrams (Fig. 35g) revealed that the potential-limiting step shifted from *CO desorption in Fe–N4–C (0.70 eV) to *COOH formation in Fe2–N6–C–O (0.35 eV), demonstrating how dual-metal synergy optimizes both activity and selectivity. Further, DFT calculations were employed to evaluate the CO2RR performance of complex COF@MOF-SACs (Fig. 36a), comparing catalytic pathways at dual metal sites (Co and Zn sites).83 The results showed that the formation of *COOH was the rate-determining step, with Co sites exhibiting a much lower free energy barrier (0.25 eV) than Zn sites (0.58 eV in COF@MOF800-Co and 1.32 eV in COF@MOF800) (Fig. 36b and d). Additionally, the competing HER reaction was analyzed, demonstrating that Co sites effectively suppressed HER in favour of CO2RR (UL(CO2)–UL(H2) = −0.11 eV), while Zn sites exhibited greater activity toward HER under similar conditions (Fig. 36c). This example highlights the critical role of heterometallic synergy in COF@MOF-SACs for tuning reaction selectivity and optimizing CO2RR efficiency (Fig. 36d).
![]() | ||
Fig. 36 (a) Geometric optimization of the COF@MOF–Co complex with key CO2RR intermediates, obtained through DFT simulations. Comparative energy profiles of COF@MOF–Co, Co, and Zn sites for (b) CO2RR, (c) HER, and (d) overpotentials at the rate-limiting step of CO2RR catalytic sites. Reproduced from ref. 83 with permission from Wiley (CC-BY 4.0), copyright 2023. |
A representative application of DFT in photocatalytic CO2 reduction is exemplified by the mechanistic investigation of a Pt SACs anchored on a triazine-based COF.303 In this system, density functional theory provided comprehensive insights into how atomic-level Pt incorporation modulates electronic structure and catalytic behavior under light-driven conditions. Geometry optimizations and binding energy calculations identified the most stable configuration as a Pt atom coordinated to one nitrogen and two carbon atoms within the COF matrix. This N–Pt–C2 site exhibited a binding energy of −2.42 eV and bond distances consistent with established SA coordination geometries, indicating strong structural anchoring. Bader charge analysis and charge density difference plots revealed that Pt donates approximately 0.46 electrons to adjacent framework atoms, confirming its partial oxidation and effective electronic coupling with the COF lattice. Band structure and DOS calculations further showed that Pt loading reduces the band gap from 2.33 to 1.73 eV, promoting enhanced visible-light absorption. The observed hybridization between Pt d-orbitals and the COF π-system introduces new frontier states, which enhance the redox capacity and facilitate charge separation under illumination. Work function analysis indicated a decrease from 5.88 to 5.56 eV upon Pt incorporation, accompanied by an upward shift in the Fermi level. This electronic modification favours photoinduced electron transfer from the COF backbone toward adsorbed reactants, thus enhancing surface reactivity. From an electronic perspective, the incorporation of Pt atoms into the COF framework enhances CO2 adsorption by modulating the local charge distribution at the active site, as revealed by charge density difference analysis (Fig. 37a and b), and by narrowing the band gap, as shown by the calculated band structure and density of states (Fig. 37c and d). This dual modulation of electronic structure and charge localization contributes to improved photocatalytic CO2 conversion efficiency and product selectivity in the Pt-COF system. DFT-derived reaction energy profiles (Fig. 37e and f) further demonstrate that the initial hydrogenation of CO2 to the *COOH intermediate proceeds with a significantly lower energy barrier on the Pt-COF surface compared to the SAC-free COF. Subsequent *CO formation and desorption are also thermodynamically favourable, steering the reaction pathway towards selective CO evolution while suppressing deeper hydrogenation to hydrocarbons. These findings demonstrate how DFT elucidates structure–function relationships at the atomic level and guides the rational design of high-performance COF-based SAC photocatalysts for solar-driven CO2 reduction.
![]() | ||
Fig. 37 DFT analysis of the Pt-COF photocatalyst for CO2 reduction. (a) and (b) Charge density difference and Bader charge results showing electron transfer from Pt to the COF. (c) and (d) Band structure and density of states illustrating band gap narrowing and Pt-COF orbital hybridization. (e) and (f) Free energy diagram and reaction pathway showing reduced energy barrier for *COOH formation and enhanced CO selectivity. Reproduced from ref. 303 with permission from Elsevier, copyright 2023. |
While the discussed Pt-COF study provides valuable mechanistic insights based on conventional Kohn–Sham DFT with generalized gradient approximation (GGA) functionals, further refinement of electronic structure predictions in photocatalysis can benefit from the application of advanced computational approaches. For instance, hybrid functionals such as HSE06 or range-separated methods offer improved accuracy in predicting band gaps and band edge positions, which are critical for evaluating light absorption and redox alignment.304,305 Time-dependent DFT (TDDFT) and many-body perturbation theories like GW and Bethe–Salpeter equation (BSE) methods can provide deeper insight into photoexcited states and exciton dynamics, while periodic DFT using slab models allows for more realistic simulation of surfaces and interfaces,306–308 particularly in heterojunction or MOF/COF composite photocatalysts. Incorporating these advanced tools into DFT workflows enhances the reliability of descriptors such as band bending, surface dipoles, and charge carrier mobility, thereby enabling more predictive modelling of photocatalytic CO2 reduction performance under solar irradiation.
Standard DFT workflows in this domain include transition-state identification via the climbing-image nudged elastic band (CI-NEB) method, vibrational frequency analysis for entropic and zero-point corrections, and Bader charge analysis to track electron transfer across intermediates.313,314 Energetic span theory is frequently applied to link computed barriers to turnover frequencies (TOFs), offering predictive insights into catalyst activity trends. Importantly, thermodynamic descriptors derived from DFT calculations are often coupled with microkinetic modeling to resolve complex product networks and competing mechanistic routes.
A representative example by Liu et al. evaluated five transition metals (Cu, Fe, Ni, Pd, Pt) anchored on Zr nodes in MOF-808 using periodic DFT.315 Binding energy analysis revealed that Pt was the most strongly anchored species, whereas Cu exhibited the most favourable catalytic reactivity (Fig. 38a). Two full reaction mechanisms for CO2 hydrogenation were examined, the *HCOO (formate) and *COOH (carboxyl) pathways. For Cu-MOF-808, the formate route was both thermodynamically and kinetically preferred, with the rate-limiting step, *H2COOH → *H2CO, exhibiting a barrier of 10.4 kcal mol−1 (0.45 eV) (Fig. 38b). Energetic span analysis revealed that Cu had the highest predicted TOF of 6.63 × 10−15 among the five metals, aligning with experimental findings. Mulliken charge analysis indicated electron-rich Cu sites facilitated stronger *CO2 activation, while frequency analysis confirmed thermal accessibility of key steps under realistic reaction conditions.
![]() | ||
Fig. 38 DFT-based mechanistic insights into thermocatalytic CO2 hydrogenation; (a) optimized geometries of MOF-808 supported single-metal sites (M = Cu, Ni, Pd, Pt, Fe), showing stable coordination environments. (b) Free energy profile for CO2 hydrogenation to formic acid (HCOOH) over Cu(II)-MOF-808. Reproduced from ref. 315 with permission from American Chemical Society, copyright 2025. (c) Energy diagram for CO2 hydrogenation to CH3OH via the formate pathway over Cu–N4 SAC and (d) Energy diagram for CO2-to-CO conversion via the RWGS pathway over Cu–N3 single-atom catalyst. Reproduced from ref. 316 with permission from Nature (CC-BY 4.0), copyright 2021. (e) Thermodynamic validation approach from Lin et al., illustrating how reversibility analysis supports or challenges DFT-predicted mechanisms. Reproduced from ref. 286 with permission from American Chemical Society, copyright 2022. |
Similarly, DFT-guided mechanistic investigation in thermocatalytic CO2 hydrogenation involves Cu SACs supported on buckled C3N4, where the coordination environment of the Cu site (Cu–N3 vs. Cu–N4) was shown to govern product selectivity and pathway preference.316 Geometry optimizations confirmed stable Cu anchoring at both tri and tetra coordinated nitrogen sites, with higher binding energy observed for Cu–N4 (2.98 eV), indicating stronger thermodynamic stabilization. Ab initio molecular dynamics simulations at 450 K affirmed structural integrity of both configurations, while Bader charge and PDOS analyses revealed greater electron transfer and d–p orbital coupling in the Cu–N3 model, consistent with a higher Cu valence state. Subsequent DFT calculations mapped out distinct reaction pathways: CO2 hydrogenation over Cu–N4 favoured the formate route toward CH3OH, with energetically accessible *HCOO and *H2COOH intermediates and a free energy drop of 0.36 eV in the final C–O bond cleavage (Fig. 38c). In contrast, Cu–N3 supported CO formation via the RWGS pathway through a *HOCO intermediate, with lower CO desorption energy (−0.36 eV), facilitating rapid turnover (Fig. 38d). These results demonstrate how fine-tuning the coordination environment of SACs can selectively direct CO2 conversion toward CH3OH or CO under identical thermal conditions, reinforcing DFT power in correlating active site structure with catalytic function in SACs.
While DFT derived energy profiles allow detailed mapping of elementary steps, Lin et al. emphasized that reaction feasibility must also satisfy global thermodynamic and kinetic criteria. To this end, they developed a theoretical framework based on De Donder's inequality and effective reversibility analysis to assess the thermodynamic self-consistency of proposed multi-step pathways, such as CO2 → CO → CH3OH. Mechanistic schemes that appear kinetically feasible at the step level may nevertheless exhibit thermodynamic inconsistency when evaluated over the full reaction network, particularly if reversibility constraints are not satisfied (Fig. 38e). To strengthen mechanistic discrimination, Lin et al. further integrated DFT calculations with experimental delplot analysis and time-resolved yield measurements, enabling the distinction between primary product formation and secondary transformation routes. This combined approach establishes a direct link between atomistic-level modelling and macroscopic kinetic observables, providing a more robust foundation for interpreting DFT-derived selectivity trends in thermocatalytic systems.
These DFT case studies exemplify the wide applicability of theoretical modeling in understanding and designing SACs for CO2RR. For COF-SACs, DFT clarifies how framework topology, metal coordination, and orbital distributions collectively influence catalytic pathways. For MOF-SACs, DFT elucidates how neighboring metal centers create electronic synergy, which fundamentally alters binding energies, electronic structure, and rate-determining steps. Such insights are crucial when interpreting experimental data, designing new catalysts, and guiding machine-learning models for predictive catalyst discovery. The combined theoretical-experimental approach ensures that DFT not only validates existing observations but also uncovers hidden correlations that might be experimentally inaccessible. As shown in these examples, DFT is essential for dissecting how electronic structures evolve within the dynamic microenvironments of MOFs and COFs, paving the way for the rational design of next-generation SACs tailored for efficient and selective CO2 reduction.
The identification of efficient catalysts relies on determining optimal active sites, coordination environments, and electronic structures. However, the extensive number of possible SAC configurations in MOF- and COF-derived systems makes traditional theoretical screening resource-intensive. High-throughput screening leverages DFT-based automated workflows to compute key catalytic descriptors such as adsorption energies, reaction-free energies, and rate-determining steps, across numerous SAC candidates.322,323 These workflows generate predictive performance maps, ranking materials based on expected catalytic efficiency.324 However, comprehensive DFT studies demand substantial computational resources, particularly when assessing multimetallic SACs or catalysts embedded in MOF/COF matrices, where periodic boundary conditions and solvent effects must be considered.
To mitigate computational costs, ML techniques have been applied to accelerate materials discovery. By training predictive models on extensive DFT datasets and validated experiments, ML algorithms rapidly estimate key catalytic descriptors for materials that have not been directly computed, maintaining high accuracy while significantly reducing computational time.325 Additionally, ML models uncover fundamental structure–property relationships,326 which is particularly advantageous for SACs within MOF/COF frameworks due to their intricate coordination environments. Graph neural networks (GNNs) have demonstrated particular effectiveness in MOF- and COF-derived catalysts, capturing both atomic-level metal coordination and long-range framework interactions, which are crucial for accurate catalytic performance predictions.327 Beyond simple property prediction, ML-driven inverse design has gained prominence.328 This approach allows researchers to define target catalytic properties such as optimal binding energies for intermediates while the model recommends suitable metal centers, coordination geometries, and linker environments, streamlining the transition from computational design to experimental realization. Different ML models are selected depending on dataset size. Linear models (e.g., LASSO, ENR) are suitable for small datasets, kernel regression models (e.g., SVR, GPR, KRR) for moderate-sized datasets, decision trees (e.g., KNR, RFR, ETR) and neural networks (e.g., FNN) for larger datasets, etc.329,330 GNNs have proven particularly effective for various SAC systems by encoding both the local metal site and supported coordination.327
For MOF-derived and COF-based SACs, DACs, and MACs, ML-assisted screening approaches remain largely unexplored. However, they represent an adaptable and promising methodology for catalytic applications. For instance, an ML-accelerated workflow analyzed M–N4–C catalysts based on graphdiyne for CO2RR, predicting catalytic performance with accuracy comparable to DFT while reducing computational cost.331,332 Similarly, a multi-objective ML framework integrating DFT adsorption energy data with heuristic stability metrics has been used to identify high-performance active sites in COF-supported SACs.319 Beyond screening, ML has been instrumental in inverse design, where target CO2 intermediate binding energies guide the selection of optimal SAC configurations, predicting ideal metal coordination environments and linker modifications for enhanced selectivity and stability. The dependence correlation mechanism establishes a direct relationship between reaction energy barriers, elemental composition, and active site configurations, providing crucial insights for electrocatalyst screening.331 In particular, the double-dependence correlation model refines this approach by identifying strong preferences for both specific elements and active sites simultaneously, thereby narrowing down promising catalyst candidates to a highly selective range.333 This methodology enhances the predictive accuracy of electrocatalyst discovery by prioritizing key reaction steps that dictate catalytic performance. As illustrated in Fig. 37, this framework is integrated into machine learning-assisted screening, where high-throughput DFT calculations first establish selectivity trends (Fig. 39a), followed by the mapping of comprehensive reaction pathways for CO2 valorisation (Fig. 39b). The double-dependence correlation mechanism (Fig. 39c) is particularly effective in refining catalyst selection by limiting viable candidates to those exhibiting optimal electronic and structural properties. Reported SACs for CO2 valorisation (Fig. 39d) demonstrate the applicability of this strategy, while the overall workflow of first-principles machine learning models (Fig. 39e) highlights its role in accelerating catalyst design. While SACs have been the primary focus of ML-driven screening, these methods have also been extended to DACs and MACs for improved catalytic synergy.331,333,334 A comprehensive computational screening study reported the favourability of Fe–Ni dual atoms on defective graphene for CO2RR.287 Additionally, prediction studies on DACs and MACs on metal surfaces have been well-practiced and reported. For instance, Karmodak et al. identified 11 stable Fe–M systems on an Au surface for CO2RR, demonstrating their potential to suppress the competitive HER.287 Similarly, a recent high-throughput ML study screened DACs for CO2RR by analyzing electronic descriptors and coordination-dependent adsorption energies, revealing that Cr-Zr DAC on Cu surface exhibited superior C2 selectivity compared to monometallic SACs due to enhanced charge redistribution at active sites.335 Overall, integrating high-throughput screening, ML, and experimental validation is transforming SAC development for CO2 valorisation. These data-driven strategies complement traditional experimental and mechanistic insights from DFT and in situ studies, establishing a rational and efficient platform for catalyst design. Further advancements in high-quality materials databases, predictive algorithms, and multi-source experimental–computational integration will enhance ML-driven catalyst design, ensuring broader adoption and practical implementation in CO2 conversion technologies.
![]() | ||
Fig. 39 (a) DFT-based selectivity analysis of SACs for CO2RR. Reproduced from ref. 290 with permission from Elsevier, copyright 2023. (b) Comprehensive reaction pathways for CO2 valorisation into value-added products. (c) Schematic representation of the double-dependence correlation model. Reproduced from ref. 332 with permission from American Chemical Society, copyright 2022. (d) Overview of various SACs reported for CO2 valorisation into different products. Reproduced from ref. 336 with permission from Royal Society of Chemistry (CC-BY 4.0), copyright 2023. (e) Workflow of the first-principles machine learning strategy for catalyst discovery. Reproduced from ref. 337 with permission from Wiley, copyright 2023. |
Likewise, efficient reactant (CO2) transport towards- and product removal from the catalyst surface is crucial for scaling-up. Lab-scale experiments often employ conditions facilitating excellent mass transport. However, industrial systems face more complex dynamic issues and mass transport limitations, hampering the effectiveness of highly active electrocatalysts. Scaling up to larger sizes requires addressing limitations in transferring material which greatly impacts cost and overall productivity. Integrating innovative material transportation systems for ensuring effective flow, improves the material transport to improve large-scale commercial applications that meet industrial production requirements. Also, electrochemical reactors designed for laboratory setups do not translate readily to industrial scales. Industrial-scale reactors require cost-effective, high-volume manufacturing with reliable materials. The capital expenditure and operation costs can be reduced with significant engineering design changes. Optimizing the flow design along with improving transport of reactant through appropriate materials are significant engineering hurdles needing innovation to achieve efficient and high production processes.
Similarly, achieving high product selectivity at industrial scale is challenging. Laboratory studies frequently report high selectivity towards certain products. However, this often comes from simplified experimental parameters at reduced scales where mass transport, chemical concentration or byproduct contamination effects were not accounted during optimization. Achieving high selectivity and maintaining quality consistently at large scale production remains more challenging due to increased occurrence of cross-contamination from diverse reactants, which increase challenges in efficient and cost-effective purification. The separation of desired products from mixtures becomes an energy intensive process and substantial source that greatly contributes to the increased production cost and requires advanced post processing which adds significantly higher cost for purification that will influence economic viability and make commercial success far more difficult to achieve. Advanced membrane technologies, advanced processing methodologies and exploring novel configurations along with refining processing methods and streamlining and implementing cost-effective procedures would enhance potential toward wider applicability and facilitate large-scale deployment in industrial setting while maintaining higher profits by minimizing processing step cost and maximizing overall energy efficiency in purification processes along with producing consistent and adequate purity requirements demanded. A comprehensive techno-economic analysis needs detailed cost assessment of the multiple stages for effectively capturing and transporting materials for deployment.
The electrochemical reduction of CO2 (CO2RR) requires a carefully designed electrolysis system that balances catalytic efficiency, mass transport, energy consumption, and long-term operational stability. Electrolysis setups for CO2RR can be broadly categorized into three main configurations:
H-type cells, flow cells with gas diffusion electrodes (GDEs), and membrane electrode assembly (MEA) systems. In contrast, flow cells enable higher performance by continuously supplying CO2 gas to a gas diffusion layer. These systems incorporate anion exchange membranes (AEMs), cation exchange membranes (CEMs), or bipolar membranes (BPMs) to facilitate ion transport and separate the electrode compartments (Fig. 40a). Flow cells are capable of operating at current densities >100–500 mA cm−2 and support continuous product removal and electrolyte control. The membrane electrode assembly (MEA) configuration integrates the catalyst layers directly onto the membrane, eliminating the liquid electrolyte in the reaction zone. This design improves system compactness and reduces ohmic resistance, although challenges include water management and gas sealing. As a starting point towards achieving industrial applications the eCO2RR usually investigated using flow-cell or membrane electrode assembly (Fig. 40b). Although not many MOF based materials were studied for high current density analysis, other CO2 reduction materials were tested for long-time operation with high current density. For instance, various copper-based materials have been tested for C2 products with very high current density with long-time operational stability (Fig. 40c and d). Operational parameters such as current density, cell voltage, CO2 flow rate, temperature, and electrolyte composition all influence performance. Industrial targets include current densities >200 mA cm−2, faradaic efficiencies above 80%, and cell voltages under 3 V. Challenges such as carbonate formation, product crossover, electrode flooding, and long-term catalyst degradation remain critical bottlenecks. Addressing these issues through membrane innovation, water management strategies, and robust catalyst development is essential for commercial viability.347
![]() | ||
Fig. 40 (a) Configuration of various electrochemical cells (AEM, PEM, BPM). Reproduced from ref. 346 with permission from Elsevier (CC-BY 4.0), copyright 2024. (b) Typical membrane electrode assembly. (c) FEs under various applied potentials catalyzed by CuONPs in 1 M KCl. (d) Partial current density of eCO2 reduction products catalyzed by CuONPs in 1 M KCl. Reproduced from ref. 347 with permission from Royal Society of Chemistry (CC-BY 4.0), copyright 2023. |
![]() | ||
Fig. 41 (a) Schematic flow diagram illustrating the operation of an electrolyser-based facility designed to produce 100 tonnes of carbon monoxide. (b) Comparison of energy costs associated with conventional, electrolysis-based, and plasma-based CO production methods, based on their respective baseline scenarios. Reproduced from ref. 350 with permission from Royal Society of Chemistry (CC-BY 4.0), copyright 2024. |
In addition to electricity, balance of plant costs including gas purification, product separation, water management, and stack replacement contribute significantly to the overall cost. Innovations such as Gas diffusion electrodes (GDEs) for better mass transport, zero-gap or membrane electrode assembly (MEA) reactors for reduced ohmic loss, Anodic organic oxidation (replacing OER) to generate co-products and lower energy input and Pressurized operation to improve CO2 solubility and system compactness are being explored to reduce total energy consumption and improve economic returns. Moreover, policy incentives, carbon pricing, and green premium markets for synthetic fuels and chemicals can further tilt the economic balance in favour of CO2 electroreduction technologies. For example, with a carbon credit of $50–100 per ton CO2, many systems could become economically favourable even at slightly higher electricity prices or lower product yields. Recent TEAs have shown that products like CO and formic acid are closest to commercial viability, while multi-carbon products (C2+) such as ethanol and ethylene require significant advancements in performance and system integration to reduce production costs Table 5.351
Techno-economics | Products | ||||
---|---|---|---|---|---|
Carbon monoxide | Formate | Methanol | Ethanol | Ethylene | |
a Single-atom catalysts.b For half-cell studies, the cell voltage is estimated from the assumption of 1.6–1.8 V versus a reversible hydrogen electrode for the oxygen evolution as the counter electrode reaction.351 | |||||
2019 US market prices ($ kg−1) | 0.15 | 0.50 | 0.26 | 0.48 | 0.58 |
Published production costs ($ kg−1) Average ± standard deviation [minimum, maximum] | 0.39 ± 0.19 [0.18, 0.64] | 0.96 ± 0.78 [0.10, 2.63] | 1.40 ± 1.03 [0.54, 2.64] | 3.92 ± 3.96 [0.37, 11.27] | 2.48 ± 1.83 [0.65, 4.92] |
Figures of merit | |||||
Catalysts | Ni SACsa | Bi, Sn | Co–Pc | N–C/Cu | Cu |
Faradaic efficiency (%) | 90–99 | 80–90 | ∼30 | ∼50 | >70 |
Cell voltageb (V) | 2.2–2.5 | 3–4 | >2 | 2.5–3.0 | 2.4 |
Current density (mA cm−2) | >400 | >100 | <100 | >100 | >200 |
![]() | ||
Fig. 42 (a) Pictorial representation of the advantage of co-electrolysis of CO2 with organic/biomass oxidation. Reproduced from ref. 341 with permission from Nature (CC-BY 4.0), copyright 2019. (b) Paired-electrolysis of CO2 with hydroxymethylfurfural. Reproduced from ref. 352 with permission from American Chemical Society, copyright 2020. (c) Co-electrolysis of CO2 and glycerol using bipolar membrane. Reproduced from ref. 353 with permission from Elsevier, copyright 2022. (d) Schematic of the coupled eCO2RR-HMFOR system in the MEA setup. Reproduced from ref. 354 with permission from Royal Society of Chemistry, copyright 2023. (e) Schematic representation of integrated electrolysis flow cell coupling CO2RR with HMFOR. Reproduced from ref. 355 with permission from Royal Society of Chemistry (CC-BY 4.0), copyright 2023. |
For Instance, an efficient CO2 electrolysis system was demonstrated by replacing the oxygen evolution reaction with biomass oxidation using 5 nm nickel oxide (NiO) nanoparticles as anodic catalysts. A two-compartment electrolysis cell was constructed to couple 5-hydroxymethylfurfural (HMF) oxidation using NiO nanoparticles with CO2 reduction to formate on BiOx electrocatalysts. Operating in CO2-saturated 0.5 M KHCO3, the system achieved a stable 2.5 V cell voltage at 2 mA cm−2 over 3 hours, yielding 81% faradaic efficiency (FE) for formate and 36% FE for biomass oxidation (Fig. 42b).352
Similarly, a co-electrolysis system combining CO2 reduction to ethylene (C2H4) at a copper cathode with glycerol oxidation to glycolic acid at a gold nano-dendrite anode was developed using a bipolar membrane to prevent carbonate crossover. The system operated at 175–225 mA cm−2 with 50% faradaic efficiencies for both C2H4 and GA, reducing cell voltage by 0.8 V compared to conventional setups. Techno-economic analysis showed the process could achieve a competitive ethylene price of ∼$1.1 per kg, highlighting its industrial potential over traditional eCO2R routes (Fig. 42c).353 A paired electrolysis system was developed for economical formate production and biomass valorisation by coupling CO2 reduction (eCO2RR) with 5-hydroxymethylfurfural oxidation (HMFOR). Single-atom Cu-doped Bi (Cu1Bi) served as the cathode, achieving formate current densities >1 A cm−2 and long-term stability, while NiCo layered double hydroxides (NiCo LDHs) catalyzed HMF oxidation to 2,5-furandicarboxylic acid with >95% faradaic efficiency. This coupling reduced energy consumption to ∼3493 kWh per ton of formate 22.9% lower than conventional systems highlighting a promising strategy for efficient CO2 utilization and biomass upgrading (Fig. 42d).354 To enhance product value and energy efficiency, Cu foam was modified into CuO nanoflowers (CuO–NF@Cu), enabling dual-function electrocatalysis for CO2 reduction (CO2RR) and 5-hydroxymethylfurfural oxidation (HMFOR). During CO2RR, CuO converted into Cu2O/Cu nanoflowers, delivering 70% faradaic efficiency (FE) for ethylene at 104.5 mA cm−2 over 45 h. For HMFOR, CuO–NF@Cu achieved 99.3% FE for FDCA at 1.62 V. In a paired system, using Cu2O/Cu–NF@GDL (cathode) and CuO–NF@Cu (anode), 188.8 mA cm−2 was reached at 2.75 V with 74.5%/96.6% FEs for C2H4 and FDCA, respectively. This demonstrates a practical route to co-produce high-value chemicals from CO2 and biomass.355
Successful implementation of paired electrolysis relies on the rational design of compatible COF and MOF based SACs, membrane architectures, and operating conditions to ensure synchronized kinetics and selective conversion at both electrodes. As a sustainable and modular platform, paired electrolysis offers a powerful route for integrating carbon capture with biomass vaporization, enabling circular chemical production powered by renewable electricity. This area of research opens a wide window for the development of COF and MOF based SACs.
(2) COF and MOF materials offer tunable pore environments that enable precise control over CO2 diffusion, adsorption, and single-atom catalyst (SAC) exposure crucial factors in enhancing CO2RR activity and selectivity. Looking forward, the design of functionalized ligands with electron-donating groups can strengthen metal–ligand interactions, stabilize active sites, and modulate local electronic structures for optimal CO2 activation. Post-synthetic modifications (PSM) further expand the toolbox for tailoring coordination environments and introducing catalytic specificity. These framework-based strategies hold significant promise for advancing SAC-based CO2RR toward higher efficiency, product selectivity, and long-term operational stability.
(3) Unraveling catalytic mechanisms at the atomic level remains a cornerstone for the rational design and optimization of single-atom catalysts (SACs). Traditional ex situ characterization offers only a static snapshot of the catalyst, often missing critical insights into the dynamic behavior of active sites under operating conditions. Therefore, integrating in situ and operando spectroscopic techniques such as X-ray absorption spectroscopy (XAS), Raman, infrared (IR), and ambient pressure X-ray photoelectron spectroscopy (AP-XPS) is imperative. These tools enable real-time tracking of the coordination environment, electronic structure, and intermediate species during catalysis, thereby shedding light on the true nature of active sites and their evolution.
(4) SACs face challenges in cost-effectiveness and scalability due to expensive precursors, complex synthesis, and material stability, including costly noble metals, precise methods, and stability of catalysts. Therefore, more research must be focused on reducing these drawbacks. Scalable synthesis via spray pyrolysis, electrospinning, or self-assembly strategies can reduce fabrication expenses.
(5) MOF/COF-SAC hybrids using carbon-based supports such as graphene, and CNT can improve durability and reduce the required metal content. Bimetallic and alloy-based SACs may improve efficiency, reducing metal consumption.
(6) DFT studies can be utilized to optimize metal–ligand interactions, electronic structures, and reaction pathways, reducing experimental trial-and-error. Molecular dynamics simulations assess catalyst stability under real-world conditions, and multiscale modeling for reaction kinetics, including Kinetic Monte Carlo and microkinetic modeling which can provide accurate predictions of catalytic activity.
(7) Emerging COF and MOF based SACs architectures are revolutionizing tandem synthesis workflows, particularly in the integration of CO2 reduction reaction (CO2RR) with the oxidation and valorisation of organic compounds. These organic substrates (ethanol, methanol, glycerol, aldehydes, amino acids, furans, glucose, and fructose etc.), which are more easily oxidized than water, provide a promising pathway for coupling CO2RR with organic transformations, leading to the simultaneous production of valuable chemicals. This approach not only maximizes resource utilization but also enhances process efficiency by harnessing the synergy between CO2 reduction and organic compound oxidation. Future advancements will focus on optimizing catalyst selectivity, stability, and reaction conditions to enable seamless integration of these processes, thereby advancing the sustainability and economic viability of CO2 valorisation strategies.
This journal is © The Royal Society of Chemistry 2025 |