Sawsan Abo Talas†
a,
Pewee D Kolubah†b,
Rushana Khairovab,
Manal Alqahtanic,
Soliman I. El-Houtd,
Faisal M. Alissa
e,
Jehad K. El-Demellawief,
Pedro Castaño
*bg and
Hend Omar Mohamed*b
aDepartment of Chemical Engineering, Faculty of Engineering, Minia University, Minia, 61111, Egypt
bMultiscale Reaction Engineering (MuRE), King Abdullah University of Science and Technology (KAUST), Thuwal, 23955-6900, Saudi Arabia. E-mail: pedro.castano@kaust.edu.sa; hend.mohamed@kaust.edu.sa
cBiological and Environmental Science and Engineering (BESE) Division, Water Desalination and Reuse Center, King Abdullah University of Science and Technology, Thuwal 23955-6900, Saudi Arabia
dNanostructured Materials and Nanotechnology Department, Advanced Materials Institute, Central Metallurgical Research and Development Institute, CMRDI, P.O. Box 87, Helwan, 11421, Cairo, Egypt
eSaudi Aramco, EXPEC Advanced Research Center, P.O. Box 5000, Dhahran, 31311, Saudi Arabia
fCenter of Excellence for Renewable Energy and Storage Technologies (CREST), King Abdullah University of Science and Technology (KAUST), Thuwal, 23955-6900, Saudi Arabia
gChemical Engineering Program, Physical Science and Engineering (PSE) Division, King Abdullah University of Science and Technology (KAUST), Thuwal, 23955-6900, Saudi Arabia
First published on 22nd July 2025
The electrochemical reduction of carbon dioxide (CO2) is a crucial step toward a sustainable carbon economy, enabling the conversion of greenhouse gases into valuable fuels and chemicals. Among the emerging materials for this transformation, two-dimensional (2D) MXenes comprising transition-metal carbides, nitrides, and carbonitrides are notable due to their tunable surface chemistry and high conductivity. This review comprehensively analyzes recent advancements in MXene-based electrocatalysis for the CO2 reduction reaction (RR) and explores the unique electronic properties of MXenes that drive their catalytic performance. Composition, surface terminations, defect engineering, and interfacial dynamics dictate activity and selectivity and are analyzed to contextualize the structure–function correlations. This work discusses state-of-the-art strategies to enhance the performance of MXene-based electrocatalysts, including compositional modifications, heteroatom doping, and heterostructure integration. Mechanistic insight into the CO2RR is examined to pinpoint the advantages and challenges of MXenes in the overall reaction network. Finally, this work presents a forward-looking perspective, outlining challenges and emerging opportunities for MXenes in driving sustainable CO2 electrocatalytic conversion technology.
Wider impactThis review discusses key advances in the application of MXene-based materials as electrocatalysts for CO2 reduction, emphasizing their tunable surface chemistry, high conductivity, and structural robustness. This field has witnessed notable progress in understanding the role of composition, defect engineering, surface terminations, and heterostructure integration in tailoring catalytic performance and selectivity. These developments are of broad significance, as CO2 electroreduction presents a direct link between greenhouse gas mitigation and renewable energy utilization, enabling the production of value-added chemicals and fuels under mild conditions. The study of MXenes intersects materials science, electrochemistry, and environmental engineering, making it of compelling interest for both fundamental research and industrial applications. As global energy and climate goals intensify, the demand for efficient, scalable, and sustainable catalytic platforms is set to rise. Insights from this review—especially those concerning the molecular-level mechanisms and synthetic strategies—will help guide the rational design of next-generation 2D catalysts with enhanced activity and stability. Ultimately, these contributions will influence the development of modular and deployable CO2 utilization systems, shaping the future of materials science toward low-carbon technologies and circular economy models. |
Global energy consumption is projected to increase by about 80% by 2030, potentially driving CO2 levels and associated warming even higher without proactive mitigation.6 According to the Intergovernmental Panel on Climate Change (IPCC), anthropogenic greenhouse gas emissions raised average surface temperatures by 1.1 °C above the pre-industrial levels from 2011 to 2020, with more severe impacts to follow without substantial emission reduction.7 As shown in Fig. 1A, global temperatures have increased and are expected to continue rising throughout the lifespans of three representative generations born in 1950, 1980, and 2020. Projections indicate that, without proactive mitigation measures, global surface temperatures could rise by up to 4 °C by 2100, underscoring the urgent need for collective action across governments, industry, and society.
![]() | ||
Fig. 1 Circular economy: (A) observed (1900–2020) and projected (2021–2100) changes in global surface temperature (relative to 1850–1900), which are linked to changes in climate conditions and effects, illustrating how the climate has already changed and will change along the lifespan of three representative generations (born in 1950, 1980 and 2020). Reproduced from ref. 7 with permission from IPCC, Copyright [2023]. (B) Schematic of CO2 capture, storage, and utilization for fuel and chemical production. |
CO2 capture, storage, and utilization (CCSU) strategies aim to mitigate these emissions by capturing CO2 from major point sources (power plants, gas-processing facilities, and industrial sites) and purifying, compressing, and injecting it into deep geological formations (depleted reservoirs, saline aquifers, and coal seams) for long-term sequestration. The choice of capture method (pre-combustion, post-combustion, or oxy-fuel combustion) depends on the CO2 concentration, required purity, and pressure of the emission stream. Those strategies are considered vital tools to reduce global CO2 emissions by up to 32% by 2050 (Fig. 1B).8–15 The CCS provides the CO2 feedstock for utilization pathways, enhanced oil recovery, construction materials, chemical and fuel synthesis, and agricultural applications, creating value and supporting a circular-carbon economy.16 However, the thermodynamic stability of CO2 requires high energy input for the chemical activation and conversion of the OC
O molecule into hydrocarbons, alcohols, or oxygenates.
Four primary catalytic approaches have been explored for CO2 conversion: thermal, electrochemical, photocatalytic, and photothermal catalysis. In thermal catalysis, CO2 is converted through high-temperature reactions such as hydrogenation and dry reforming with methane (CH4), typically conducted at elevated pressures and under above-atmospheric conditions.17 Photocatalysis mimics natural photosynthesis, using solar energy to generate electron–hole pairs that migrate to the catalyst surface and drive redox reactions with adsorbed CO2 species.18 Photothermal catalysis combines photochemical and thermochemical pathways to enhance reaction rates. This approach operates at lower temperatures than conventional thermal methods by harnessing the synergistic effects of semiconductor excitation and localized heating induced by plasmonic or nonplasmonic nanostructures.19,20
Among these approaches, electrocatalysis offers a uniquely advantageous route for CO2 utilization by directly coupling with renewable electricity. Unlike thermal processes that depend on externally produced H2 (e.g., via water electrolysis), the electrochemical CO2 reduction reaction (CO2RR) proceeds via a proton-coupled electron transfer (PCET) mechanism. This enables CO2 conversion under mild operating conditions, typically at room or moderately elevated temperatures (<100 °C) and ambient pressure, making it energy-efficient and scalable.21 The CO2RR can generate a wide range of high-value products, including carbon monoxide (CO), formic acid (HCOOH), ethylene (C2H4), ethanol (C2H6O), and propanol (C3H8O).22 However, the competing hydrogen evolution reaction (HER), which reduces product selectivity by diverting electrons toward H2 production, remains a key challenge.
Most of the CO2RR products possess commercial relevance, and their selectivity can be tailored through rational design of the reaction environment, which involves optimizing electrolyte composition and reactor configuration, and, most critically, catalyst engineering.23,24 By tuning catalyst morphology, elemental composition, exposed crystal facets, and defect structures, researchers aim to enhance catalytic activity, boost selectivity for specific products, and improve long-term durability.25
Despite rapid progress in the electrocatalytic CO2RR, many conventional catalyst systems, such as transition metals, metal oxides, and carbides, suffer from low selectivity, poor product formation rates, and rapid deactivation.26 As a result, significant research has focused on engineering more effective catalytic architectures, including alloy catalysts, metal–support hybrids, and heterostructures, that enhance CO2 activation, suppress the competing HER, and improve stability.27–29
Recent years have seen a few breakthrough studies that define the current performance frontier for the CO2RR. For instance, Wang et al.30 developed a fluorine-modified copper (Cu) catalyst integrated into a flow-cell reactor, achieving an unprecedented current density of 1.6 A cm−2, >80% C2+ faradaic efficiency (FE), and 16.5% single-pass yield under ambient conditions. The fluorine surface functionalization was shown to enhance water activation and stabilize key CHO intermediates, thus promoting C–C coupling and boosting multicarbon product selectivity. Sargent et al.31 demonstrated that pairing a strong-acid electrolyte with an atomically sputtered planar Cu catalyst yields > 90% C2+ FE, 78% single-pass CO2 utilization, and 30% energy efficiency for C2+ products, with an exceptionally low energy cost of 249 GJ t−1 for ethanol production. While promising, the synthesis approach is complex and cost-intensive, posing scalability challenges. Yang et al.32 achieved 97% CO selectivity at −0.5 V vs. RHE and a specific current of 350 A g−1 using nickel (Ni) single atoms supported on N, S-co-doped graphene. A 100 h durability test confirmed long-term operational stability, and DFT calculations revealed that the non-centrosymmetric ligand environment around Ni(I) significantly enhanced adsorption strength for CO2 and key intermediates. These breakthroughs illustrate how combining precise atomic-level design, reactor integration, and advanced surface engineering can collectively overcome long-standing CO2RR challenges. Yet despite such advances, broader challenges remain in scalability, cost, and long-term performance, motivating exploration of new catalyst platforms.
Among the most promising emerging materials are two-dimensional (2D) systems, including graphene, transition-metal dichalcogenides (TMDs), layered double hydroxides (LDHs),23 and the newest class—MXenes. MXenes are a family of 2D transition-metal carbides, nitrides, or carbonitrides, known for their high electrical conductivity, tunable surface chemistry, excellent mechanical integrity, and abundant redox-active sites, making them particularly attractive for CO2RR applications.25,33–40 For instance, their metallic conductivity facilitates electron transport for complex proton-coupled electron transfer (PCET) steps. Surface terminations (e.g., –O, –OH, and –F) can be engineered to steer intermediate adsorption and reaction pathways, while their robustness supports long-term operation under electrochemical conditions. However, despite these promising features, most experimental MXene-based CO2RR studies focus on Ti3C2Tx, due to the ease of synthesis and well-established chemistry.41 This field remains in its early stages: over 80% of reported studies are theoretical, relying on quantum mechanical calculations to evaluate CO2 binding, activation energies, and selectivity trends.42 Experimental progress lags due to synthesis challenges, such as controlling etching, delamination, defect density, and surface terminations, affecting catalytic performance and reproducibility.
This review provides a comprehensive overview of recent progress in MXene-based catalysts for the CO2RR. A brief overview of the CO2RR covers general reaction mechanisms, pathways for various product groups (e.g., hydrocarbons, alcohol, and oxygenates), critical intermediates governing product formation, and the internal and external factors influencing the overall system architecture. Next, this work explores the distinctive chemical and structural properties of MXenes that enhance their catalytic potential for the CO2RR, highlighting how they differ from conventional CO2RR catalysts.
A critical assessment of recent developments in catalyst design and modification strategies follows, particularly on progress in operational parameters that significantly affect catalytic activity and selectivity. Then, this work explores the electrocatalytic mechanisms of the CO2RR using MXene-based catalysts, highlighting how these mechanisms can be tailored to enhance CO2 conversion efficiency and steer product selectivity, underscoring the novelty and promise of MXene-based catalysts. Additionally, critical factors affecting CO2RR performance, including electronic properties, surface chemistry, and electrolyte interactions, are reviewed. Finally, this work provides a forward-looking perspective, emphasizing the necessity of a systematic approach to developing cost effective MXene-based catalysts for practical applications.
![]() | ||
Fig. 2 (A) Schematic of the electrochemical CO2 reduction reaction system. Reproduced from ref. 47, with permission from John Wiley and Sons, copyright [2017]. (B) Reaction mechanism, product distribution from CO2 reduction (C1–C3). Reproduced from ref. 48, with permission from Royal Society of Chemistry, Copyright [2021]. (C) Left: volcano plot of the partial current density for the CO2RR at 0.8 V vs. the CO binding strength. Right: two onset potentials plotted vs. CO binding energy, the overall CO2RR, and the conversion of CO2 to methane or methanol (Reprinted with permission from Am. Chem. Soc. 2014, 136, 40, 14107–14113. Copyright [2014] American Chemical Society). |
Although the proton-assisted process enables the formation of a wide range of products, achieving the selective production of the desired compounds remains a significant challenge due to the similar redox potentials of competing reaction pathways. Selectivity limitations are exacerbated in aqueous electrolyte-based CO2 electrolyzers, where the HER occurs at a comparable potential (eqn (13)), directly competing with the CO2RR:
O2(g) + 4H+ + 4e− → 2H2O E° = 0.82 V | (1) |
CO2 + e− → CO2˙− E° = −1.9 V | (2) |
CO2 + 2H+ + 2e− → HCOOH(aq) E° = −0.61 V | (3) |
CO2 + 2H+ + 2e− → CO(g) + 2H2O E° = −0.53 V | (4) |
CO2 + 4H+ + 4e− → HCHO(g) + H2O E° = −0.48 V | (5) |
CO2 + 4H+ + 24 → C + 2H2O E° = −0.2 V | (6) |
CO2 + 6H+ + 6e− → CH3OH(aq) + H2O E° = −0.38 V | (7) |
CO2 + 8H+ + 8e− → CH4(g) + 2H2O E° = −0.24 V | (8) |
CO2 + 12H+ + 12e− → C2H4 + 4H2O E° = −0.34 V | (9) |
CO2 + 12H+ + 12e− → C2H5 OH + 3H2O E° = −0.33 V | (10) |
CO2 + 14H+ + 14e− → C2H6 + 4H2O E° = −0.27 V | (11) |
CO2 + 2H+ + 2e− → C3H7OH + 5H2O E° = −0.32 V | (12) |
2H+(aq) + 2e− → H2(g) E° = 0 V | (13) |
Depending on the available catalyst surface during the reaction, O2 is activated and converted into intermediates. These intermediates transform into final products via distinct mechanistic pathways (Fig. 2B).49 The paths are classified into C1 products (CO, HCOOH, methanol (CH3OH), and CH4), C2 products (C2H4 and C2H6O), and a C3 product (C3H8O).46,50,51 For instance, HCOOH formation includes the creation of an *OCHO intermediate via the activation and reduction of CO2.52 This step requires catalysts with high O affinity, such as p-block metals (e.g., tin [Sn], bismuth [Bi], indium [In], and lead [Pb]), facilitating the reduction of *OCHO to form HCOOH.53–57
In contrast, producing CO involves two proton–electron steps that generate a *COOH intermediate. Catalysts (e.g., gold [Au], silver [Ag], and zinc [Zn]) known for their lower affinities for O and H reduce *COOH to CO.46 The selectivity of the overhead product in the aqueous CO2RR can be explained by the ability of various catalyst surfaces to bind reaction intermediates, in which catalyst surface properties play a critical role in determining the reaction pathway and product selectivity (Fig. 2C).58 Moreover, CO serves as a crucial intermediate for higher hydrocarbons and oxygenates, including formaldehyde (HCHO), CH3OH, and CH4, involving four, six, and eight electron transfer, respectively.46 In addition, CO is widely considered an intermediate for forming C2H4 and C2H5OH. Table 1 shows the classification of value-added products and significant intermediates.
Products | Critical intermediates | |
---|---|---|
Two-electron products | Formate (HCOO−), formic acid (HCOOH) | *OCHO/HCOO* |
Carbon monoxide (CO) | *COOH | |
Deeply reduced C1 products | Methane (CH4) | CH3O*, *OH |
Methanol (CH3OH) | CH3O*, *OH | |
Multicarbon products | Ethylene (C2H4) | CH2CHO* |
Acetaldehyde (CH3CHO) | CH2CHO* | |
Ethanol (C2H5OH) | CH2CHO*, CH3CHO*, CH3CH2O* | |
Ethane (C2H6) | CH3CH2O* | |
n-Propanol (n-C3H7OH) | CH3CHO*, CH3CH2O* |
The formation of these C2+ products involves the dimerization of two CO molecules or the combination of CO with *CHO. Copper (Cu)-based catalysts are effective in these reactions because they facilitate C–C coupling, achieving high selectivity and activity for C2+ products.59 In addition, Cu-based materials are among the most promising catalysts for CO2 reduction because they exhibit relatively low activity toward the HER while demonstrating unique intrinsic catalytic activity for the CO2RR, forming a wide range of hydrocarbons, alcohols, and oxygenates. Although the elemental composition of the CO2RR catalyst is crucial in determining reaction pathways, the overall reaction mechanism is considerably influenced by several factors, including operating conditions, such as electrolyte pH and concentration, cation/anion size, applied potential, cell configuration, and catalyst surface characteristics (e.g., facets, defects, structure, morphology, and surface adsorbates).60–63 These physical and chemical parameters strongly influence the thermodynamic adsorption energies of critical intermediates and the kinetic barriers of the reactions, leading to alternative reaction pathways. The following section examines the crucial factors influencing the CO2RR mechanism in greater detail.
![]() | ||
Fig. 3 (A) Factors affecting the CO2 electroreduction process. (B) Schematic of the electrode–electrolyte interface for the CO2RR, accounting for the electric double layer and bulk electrolyte. |
The EDL comprises charged species and oriented dipoles, organized into three distinct layers: the inner Helmholtz plane (IHP), the outer Helmholtz plane, and the diffuse layer.68 The IHP is closest to the electrode surface where the electrochemical reaction occurs, whereas cations gather in the outer Helmholtz plane region in response to the applied potential. According to the classical Gouy–Chapman–Stern model, the EDL and bulk electrolytes have a distinctly different composition. Therefore, they are considered two distinct phases separated by the diffuse layer.70 Ions form the diffuse layer due to electrostatic repulsion and thermodynamic diffusion. The concentration of the diffuse layer decreases progressively outward toward the bulk electrolyte. In the electrochemical system, the species in the EDL are widely accepted to be in equilibrium with the bulk electrolyte in the absence of any faradaic processes.67
The CO2RR is an inner-sphere process in the IHP involving adsorption and bond rearrangement of CO2 and intermediates. When the CO2 molecule adsorbs onto the electrode surface, it binds through chemisorption, a process involving electron redistribution and chemical bond formation.71 The CO2 binding process at the surface is influenced by the highest occupied and lowest unoccupied molecular orbitals. The orientation of these orbitals relative to the surface determines the nature of the binding. Such a CO2-binding process on MXenes is expected to be effective due to the tunable surface chemistry. According to the computational investigations, their surface chemistry influences the diffusion and adsorption of species and metal ions on MXene surfaces.72 Therefore, understanding the molecular-level interaction between CO2 and MXenes at the E/E interface is crucial to evaluating intermediates' adsorption/desorption process, reaction kinetics, and dynamic behavior and designing highly active, selective, and durable catalysts.
Harris et al.73 employed multinuclear magnetic resonance experiments to elucidate the strong H-bonded water molecule to the terminal hydroxyl (–OH) group on the MXene surface, providing key insights for estimating Gibbs free energy profiles and the reaction-limiting potential (UL) of spontaneous electrochemical reactions.
Later, Cheng et al.74 employed the density functional theory (DFT) to model EDL properties and proposed an intermediate electronic structure and reaction pathways of the CO2RR on modified Nb2CO2 surfaces while considering the effect of water. The results indicated that pure Nb2CO2 is unsuitable as a catalyst for the CO2RR; nonmetal-doped MXenes can lower the UL of the CO2RR and does not significantly change the reaction products. In contrast, MXene surface modification with TMs reduced the UL of the CO2RR and altered the reaction products. The V-modified Nb2CO2 system was identified as the most effective CO2RR catalyst, favoring HCOOH as the primary product with a UL of −0.11 V (Fig. 4A). The presence of d-levels in TMs provides a diverse range of electronic configurations, which can participate in bonding with CO2 molecules, enhancing the electron transfer processes essential for the CO2RR. The interaction between the d-levels of the metal and the π* orbitals of CO2 can lead to a more efficient electron transfer, which is crucial for breaking strong CO bonds in CO2.75,76
![]() | ||
Fig. 4 (A) Potential limiting the UL of NbCO2 and modified Nb2CO2. Reproduced from ref. 74, with permission from Elsevier, Copyright [2021]. (B) Linear relation of adsorption free energy ΔG(OCH2O*) and ΔG(HOCH2O*) vs. the d-band center Ed. (C) Projected density of state plots for the d orbital of surface Mo atoms in MXenes, in which the Fermi level was set to 0, and the calculated d-band centers are denoted by dashed vertical red lines. (D) Schematic of the bond formation between the adsorbate (Ads.) and reaction surface of MXenes, represented by Mo2WC2, Mo3C2, and Mo2TiC2. (B)–(D) Reproduced from ref. 77, with permission from Elsevier, Copyright [2021]. (E) Schematic of the L-SERS system for in situ local pH measurement and illustration of pH-sensitive molecules (4-MBA) in the SERS. (F)–(H) Local pH images around the catalyst surface in CO2-saturated electrolytes of (F) KHCO3 (pH 6.8), (G) K2SO4 (pH 5.6), (H) K2SO4 (pH 2.5) with the applied current density of 10![]() ![]() |
The EDL properties of MXenes modified by TMs as single-atom catalysts (SACs) have been extensively explored.79–82 Li et al.77 investigated the Mo3C2 MXene modified with a series of Group IVB, VB, and VIB TMs and explored the mechanism behind the breaking of linear scaling relationships between TMs and the adsorption energies of critical intermediates OCH2O* and HOCH2O* (Fig. 4B). Their findings revealed that substituting TMs on the MXene surface induces an upshift in the d-band center of the molybdenum (Mo) layer surface, selectively tuning the adsorption strength of OCH2O* and HOCH2O*, further lowering UL from −0.651 V for Mo3C2 to −0.350 V for Mo2TiC2 (Fig. 4C). The electron localization function analysis indicated the strong localization of lone electrons on the surface Mo layer upon TM substitution, enhancing its chemical activity due to electronic coupling between the valence states of adsorbates and the TM states, resulting in splitting the bonding and antibonding states (Fig. 4D). In addition, Mo2TiC2 demonstrated the highest conversion performance of CO2 to CH4 among the studied MXenes.
Although electrode properties play a crucial role, the E/E interface is significantly influenced by the electrolyte medium. Aqueous solutions, organic solvents, and ionic liquids (ILs) can stabilize the solid–electrolyte interphase, enhance CO2 solubility, and facilitate efficient ionic transport, all essential for optimizing electrochemical reactions. Among these, aqueous KHCO3 solutions (0.1 to 0.5 M) are widely employed to evaluate the performance of MXene-based catalysts due to their cost-effectiveness, nontoxicity, and buffering capacity, making them the preferred medium for CO2RR investigations. A recent experimental study83 demonstrated that the performance of Ti3C2Tx on a glassy carbon electrode improved for mono- and multicarbon products, achieving selectivity toward CO (42.2%), CH3OH (23.6%), C2H6O (20.1%), and acetone (10.1%) as the KHCO3 concentration in the electrolyte increased from 0.1 to 0.5 M. The authors attributed this performance to the enhanced CO2 adsorption capacity of 0.16 mmol g−1, facilitating a higher current density.
Otgonbayar et al.84 demonstrated enhanced selectivity toward alcohols, such as CH3OH and C2H6O, using a 2D MXene coupled with cuprous oxide (Cu2O)/magnetite (Fe3O4) nanocomposites using various electrolytes, including sodium carbonate (Na2CO3), potassium carbonate (K2CO3), potassium chloride (KCl), and sodium chloride (NaCl). A strong interaction between the catalyst surface and the electrolyte, particularly with alkali metals and halide ions (chlorine [Cl−]), enabled faster adsorption/desorption dynamics, influencing their ability to donate or retain a negative charge. This interaction generated a dipole moment, modified the local surface environment, and enhanced the catalytic activity.
Qu et al.85 demonstrated CO2 electrolysis using seawater as the cost-effective electrolyte. This strategy achieved 92% faradaic efficiency (FE) for CO production using synthesized nitrogen (N)-doped Ti3C2 MXene nanosheets with abundant VTi defects. Mechanistic studies have revealed that N dopants and VTi synergistically modulate the electronic structure of the active titanium (Ti) site, significantly lowering the free energy barriers for *COOH formation and *CO desorption.
Notably, pH is pivotal in influencing aspects of PCET processes in the CO2RR. The local pH at the E/E interface can change during the CO2RR, significantly affecting the selectivity and activity of the catalyst.86,87 For example, Varela et al.88 investigated the influence of the electrolyte concentration and the importance of local pH in controlling the selectivity of the CO2RR on Cu. The study demonstrated that the buffer capacity of the electrolyte is critical in influencing catalytic activity and product selectivity. In dilute KHCO3, the lower proton concentration near the electrode surface suppresses the formation of H2 and CH4, whereas the production rates of CO and C2H4 remain largely unaffected.
Xie et al.78 employed lateral-type in situ surface-enhanced Raman spectroscopy combined with computational simulations using COMSOL to investigate the effect of local pH variations at the E/E interface on CO2RR selectivity and activity (Fig. 4E). The study revealed that, in the diffusion layer, the local pH gradually decreases as the distance from the catalyst surface increases, with a sharp pH drop indicating limited proton mass transport (Fig. 4F–H). As reported, these local pH variations are influenced by the buffering capacity of the electrolyte and can have a considerable influence on the selectivity of C2+ products during the CO2RR (Fig. 4I).89
The pH dependence of CO2 activation differs from that of the competing HER. For example, previous work has demonstrated that the overall product selectivity of graphite-immobilized coprotoporphyrin is highly pH sensitive. At a pH of 1, H2 is the dominant product, whereas at a pH of 3, CO becomes the primary product, illustrating the strong influence of pH on reaction pathways.90 The DFT calculations indicate that the critical intermediate is anionic hydride, which undergoes a nucleophilic attack on CO2 to yield HCOO−. The reaction is facilitated by potential-induced changes in the oxidation state of the catalyst, with the hydride typically residing on the ligand for In and Sn porphyrins. The stability of the resulting species is vital for reactions leading to CO or HCOOH/HCOO− formation.
Pourbaix diagrams, analogous to standard pressure–volume phase diagrams, offer insight into the thermodynamic equilibrium surface structure under varying pH and applied potential (U) conditions. According to the Pourbaix diagram of MXenes with various surface termination groups (Fig. 5),91 MXenes do not exhibit a bare surface at any pH-U conditions in an aqueous environment. At a U value below −0.6 V, the MXene surface is fully hydrogenated, with all reaction sites occupied by H*. As U increases, the surface undergoes progressive oxidation, with OH* gradually replacing H*, highlighting the strong U sensitivity of the process. The intricate relationship between U and MXene functionalization, as demonstrated for other electrochemical reactions, such as the HER92 and N reduction reaction,93 is significant because it directly influences catalytic activity. This understanding can guide the design of MXene-based catalysts for the CO2RR, facilitating the selection of optimal surface terminations for enhanced performance. Future studies should develop Pourbaix diagrams for MXene compositions under CO2RR-relevant pH and U conditions, offering deeper insight into their electrochemical stability and reactivity.
![]() | ||
Fig. 5 Pourbaix diagrams of Mo2C, V2C, Ti2C, and Ti3C2. Reproduced from ref. 91 with permission from Wiley and Sons, Copyright [2022]. |
The electrolyte's metal cation/anion composition is another critical factor influencing the product distribution of the CO2RR. Alkaline–metal cations facilitate CO2 adsorption in aqueous solutions and stabilize critical intermediates, such as COOH*, via noncovalent interactions with adsorbed species or a field effect.94 For instance, Hori et al.95 observed that C2H4 formation increased relative to CH4 as the cation size increased (lithium [Li+] < sodium [Na+] < potassium [K+] < cesium [Cs+]).96 This variation in hydrocarbon selectivity was attributed to differences in specific adsorption or the preferential hydrolysis of various cations. Among anions, halide ions (F−, Cl−, bromine [Br−], and iodine [I−]) could modulate the geometry and electronic structure of metal-based electrocatalysts, stabilize active species, regulate the adsorption and desorption of reaction intermediates, and reduce the overpotential to enhance the selectivity and activity of the CO2RR for multicarbon products.97–99 This finding is attributed to halide ions donating electrons to the empty orbitals of CO2.100
Ni et al.101 synthesized fluorine (F)-doped caged porous C, achieving an FE of 88.3% for CO production at −1.0 V vs. RHE, with a corresponding current density of −37.5 mA cm−2. The microporous structure of the F-doped C shells, particularly at edge positions, induces localized high electric fields, lowering the thermodynamic energy barrier for CO2 reduction. Similarly, Gao et al.102 demonstrated that O2-plasma-activated Cu catalysts, combined with electrolyte design, exhibited enhanced CO2RR activity and selectivity toward multicarbon hydrocarbons and alcohols, achieving a FE of ∼69% and a partial current density of −45.5 mA cm−2 for C2+ products at −1.0 V vs. RHE. Their findings indicated that larger alkali-metal cations and subsurface O species promoted cation adsorption, facilitating C–C coupling on CuOx electrodes. Furthermore, using an electrolyte containing Cs+ and I− induced a significant reorganization of the CuOx surface, forming well-defined Cu species, and enhanced the intermediate stabilization and C2+ product selectivity.
Considering the well-documented positive influence of halide ions on CO2RR performance, MXenes, with abundant halide surface terminations, present a promising platform for enhancing the selectivity of multicarbon products. Halogen-based etching using molecular Cl, Br, or I can introduce halide terminations onto MXene surfaces, potentially modulating active species and regulating the adsorption–desorption dynamics of reaction intermediates during the CO2RR.103–105 For instance, a Cu-immobilized Ti3C2Clx MXene demonstrated over 58% selectivity for CH3OH, with dominant Cl functional groups residing on the outermost Ti layers.106 Synchrotron-based X-ray absorption spectroscopy and DFT calculations attributed the exceptional performance of the catalyst to single-atom Cu species with an unsaturated electronic structure (Cuδ+, 0 < δ < 2), facilitating a low-energy-barrier conversion from HCOOH* to the CHO* intermediate.
However, despite confirming the abundant Cl terminations, the study did not explicitly address their role in the activity and selectivity of the catalyst. Given the widely accepted understanding that halide ions influence reaction mechanisms by modulating intermediate adsorption and desorption, further investigation of the Cu-immobilized Ti3C2Clx catalyst is necessary to clarify the role of Cl termination groups. Future research should systematically explore the effects of –Cl, –Br, –F, and –I terminations on MXene-based catalysts, offering deeper insight into their influence on CO2RR performance and product distribution.
The electrolyte employed during the CO2RR can also influence product distribution and overall performance. Despite being the most applied electrolyte systems in the CO2RR, aqueous solutions have limited CO2 solubility (about 0.034 M), constraining their application in the H-cell. Research efforts have been directed toward using dipolar aprotic solvents, such as acetonitrile (about 0.27 M)107 and ILs, which can serve as mediators that prevent the HER and as cocatalysts that form complexes with CO2 molecules to overcome this problem and improve CO2 solubility. In one of the earliest studies on MXenes using ILs, Handoko et al.108 combined experimental and theoretical approaches to investigate Ti- and Mo-based MXenes for HCOOH production, achieving a 56% FE in an electrolyte system containing IL 3-butyl-1-methyl-1H-imidazol-3-ium tetrafluoroborate dissolved in an acetonitrile–water solution. This work highlighted the crucial role of –O surface termination groups in suppressing the HER while revealing the limitations of acetonitrile as an electrolyte due to its volatility, causing fluctuations in HCOOH selectivity.
Attanayake et al.109 suppressed the competing HER using the IL 1-ethyl-2-methylimidazolium tetrafluoroborate as an electrolyte in acetonitrile. Under these conditions, Ti3C2 and Mo2C demonstrated a remarkable FE of 90% for CO production. Despite the favorable solubility of CO2 in ILs, their relatively high viscosity compared to conventional solvents poses challenges, including a lower CO2 adsorption rate and increased pumping costs in industrial applications. Cosolvents, such as water or CH3OH, can be introduced to reduce viscosity while maintaining CO2 solubility to mitigate viscosity-related limitations.110
Previous studies on energy storage devices have demonstrated strong synergy between surface terminations of MXenes and IL functional groups, significantly influencing supercapacitor performance.111–114 Applying insight from MXene-IL interactions in energy storage devices to the CO2RR could enhance interfacial interactions, improving product selectivity and addressing stability challenges. This approach presents a promising strategy for applying the complementary properties of MXenes and ILs to enable efficient and scalable CO2 conversion into a diverse range of valuable products.
The catalyst structure can also affect the activity and selectivity of the CO2RR. A Cu-based catalyst is highly capable of producing multicarbon products. However, selectivity toward a specific product is challenging. Nanostructured Cu, including nanoparticles, nanowires, and hierarchical structures with varying compositions, sizes, morphologies, and crystal orientations, has gained considerable research attention due to its ability to enhance the selectivity of the CO2RR.115,116 The exposed crystal facets of the nanostructured Cu play a crucial role in determining the reaction pathway and controlling product distribution. Adjusting the exposed facet on the Cu crystals can change the specific atom arrangements, reaction intermediate affinity, and surface energy, influencing the CO2RR pathway and product distribution.117
Luo et al.118 investigated the facet-dependent selectivity of Cu2O nanocrystals, demonstrating that tailoring the exposed facets enhanced selectivity toward C2H4. Star-shaped Cu2O nanocrystals with (332) facets achieved over 74% selectivity for C2H4. This high selectivity was attributed to a reduction in Gibbs free energy, with the (332) facets exhibiting the lowest energy barrier (0.13 eV) in the initial step of gaseous CO2 hydrogenation, compared to (111) at 0.20 eV and (100) at 0.22 eV. Furthermore, the in situ Raman spectroscopy of star-shaped Cu2O (332) nanocrystals revealed the presence of *COOH and *CHO intermediates, indicative of C–C coupling, facilitating C2H4 formation.
The nanocrystal structure influences the facet dependence of the selectivity of Cu nanocrystals in the CO2RR. An investigation of Cu single crystals in an H-cell has revealed that (100) is suitable for C2H4 formation, whereas (111) favors CH4 formation. For example, Gregorio et al.119 developed a colloidal method to synthesize Cu cubes, spheres, and octahedral nanocrystals and tested them for the CO2RR in flow cells. The study revealed that the Cu octahedral nanocrystals dominantly produced CH4, in line with the presence of the exposed (111) facets. In contrast, the Cu cubes with dominant (100) facets exhibited much higher selectivity (55%) toward C2H4. These studies demonstrated the facet-structure dependence of the selectivity of Cu-based nanocrystals. The influence of external factors, such as temperature and applied voltage (constant and pulse/dynamic), remains underexplored in optimizing MXene performance for the CO2RR. Most studies have focused on ambient temperature conditions,120 either via DFT predictions or experimental investigations, aiming to maintain temperature as a constant parameter to ensure catalyst stability and durability. Conversely, the applied potential is a dynamic factor essential for activating the catalyst surface and forming an E/E interface. Changes in applied voltage alter the charge density on the MXene surface, affecting the organization of ions and solvent molecules in the EDL and affecting CO2 adsorption and intermediate stabilization. Moreover, potential variations can modify the oxidation state of the catalyst, changing its electronic structure and overall surface properties and influencing reaction pathways and product selectivity.
For example, Han et al.121 used in situ X-ray analysis to monitor changes in Sn species' oxidation states and the local chemical environment in the SnO2/MXene catalyst under various applied potentials ranging from −0.6 to −1.2 V vs. RHE. The authors observed a gradual shift toward lower energy on Sn K-edge X-ray absorption near-edge structure (XANES) profiles with a more negative applied potential. They reported a progressive reduction in the oxidation valence state in the CO2RR. The excellent performance and selectivity of the CO2RR to HCOOH in a 0.1 M KHCO3 aqueous solution reached a maximum of 94% at −0.8 V, attributed to the intermediate/mixed Sn oxidation state between metallic Sn0 and Sn4+.
Furthermore, Govindan et al.122 highlighted the influence of the applied potential on MXene-based catalysts, where tuning the cell potential enabled a palladium (Pd)–MXene nanocomposite to achieve a FE of 67.8% for CH3OH at −0.5 V vs. RHE. Moreover, CO and H2 became the dominant products at higher potentials, attributed to CO serving as a crucial intermediate in CH3OH formation. However, as the potential increased, CO desorption from the catalyst surface intensified, restricting its further conversion to CH3OH and shifting the selectivity toward CO and H2 evolution. This observation highlights the vital role of the applied potential in determining product selectivity and reaction pathways. However, despite its importance, no studies have investigated using a pulsed potential for MXenes in the CO2RR, highlighting a notable research gap for future investigation.
The system configuration is a critical factor in determining the overall catalytic performance of MXene-based electrocatalysts in the CO2RR. This configuration encompasses electrode type (e.g., carbon paper, glassy carbon, and gas diffusion electrodes) and reactor design (e.g., H-cells, microfluidic cells, and membrane electrode assemblies, MEAs). Each component plays a distinct role in controlling mass transport behavior, product selectivity, energy efficiency, and scalability (Fig. 6).
While H-cells are the most widely used reactors for fundamental CO2RR studies, they frequently underrepresent real-world performance due to low CO2 solubility (∼34mM), mass transport limitations, and high ohmic resistance from large inter-electrode spacing (>2 cm). As a result, H-cells typically achieve modest current densities (<50 mA cm−2) and produce faradaic efficiencies that are difficult to reproduce under industrial conditions.123,124 In contrast, flow-cell reactors, particularly those employing gas diffusion electrodes (GDEs), directly expose the catalyst to a continuous CO2 gas stream, dramatically improving CO2 mass transport and achieving reliable partial current densities between 50 and 300 mA cm−2. For instance, a SnO2 quantum dot/MXene composite attained 57.8 mA cm−2 with 94% faradaic efficiency for formate in a flow-cell GDE/MEA configuration, significantly outperforming the H-cell benchmark. Similarly, an FePc/MXene composite tested in a flow cell demonstrated ∼98% FE for CO with stability over 24 h, highlighting improved mass transport and HER suppression compared to the H-cell configuration.125
MEA-based reactors integrate solid-state ion-exchange membranes, enabling compact reactor designs, ion-specific transport control, and minimal reactant/product crossover. A Bi2O3/MXene composite demonstrated continuous operation at 300 mA cm−2 with >90% formate selectivity over 60 h in an MEA.126
Abdinejad et al.127 evaluated a Cu–Pd/MXene composite catalyst using both configurations. In an H-cell, the catalyst achieved a CO2-to-formate conversion with a FE of 79% at −0.5 V vs. RHE. In contrast, when tested in a zero-gap MEA reactor, the same catalyst achieved an enhanced FE of 93% at −2.8 V and a full-cell energy efficiency of 47%, showcasing the MEA's ability to sustain higher current densities and reduce resistance. These comparisons underscore that while H-cells are valuable for mechanistic insights, they are limited in reflecting the behavior of MXene catalysts under realistic, high-throughput conditions. MEAs offer enhanced gas transport, stable interfaces, and minimized side reactions, allowing MXenes to achieve higher selectivity, greater stability, and industrially relevant current densities. This reinforces the need to assess MXene catalysts under MEA configurations to realize and report their catalytic potential. Table 2 shows comparison of MXene catalyst performance across different reactor configurations.
MXene catalyst | Cell configuration | Current density (mA cm−2) | Product & FE (%) | Ref. |
---|---|---|---|---|
Au/Ti3C2Tx MXene | H-cell | 17.3 | CO (48.3%), H2 (25.6%) | 128 |
Cu–Pd/MXene aerogel | H-cell | ∼20 | Formate (∼79%) | 127 |
Cu–Pd/MXene aerogel | MEA | 150 | Formate (>90%) | 127 |
FePc/MXene composite | Flow-cell | >100–200+ | CO (∼98%) stable | 125 |
FePc/MXene composite | MEA | 200+ | CO (∼98%), 96% stability | 125 |
Pd–Ti3C2Tx MXene | H-cell | ∼10–20 | CO (∼48%), H2 (∼26%) | 122 |
SnO2 QD/MXene composite | H-cell | <30 | Formate (<70%) | 121 |
SnO2 QD/MXene composite | Flow-cell | 57.8 | Formate (∼94%) | 121 |
![]() | ||
Fig. 7 (A) Periodic tables presenting MAX phase and MXene compositions. (B) Top-down etching methods for MXene synthesis. (C) MXene forms synthesized using top-down etching methods. (B) and (C) Reproduced from ref. 132 with permission from Springer Nature, Copyright [2025]. (D) Schematic demonstrating the large-scale synthesis of MXenes using the HCl:HF:H2O etching method. (E) X-ray diffraction pattern of MXenes synthesized via a large-scale process and a small batch. No notable change in structure is observed. (D) and (E) Reproduced from ref. 133 with permission from Wiley and Sons, Copyright [2020]. |
The MAX phases are synthesized at elevated temperatures (800 °C to 1800 °C) in an inert atmosphere. The process involves mixing elemental powders (M and A) or carbide/nitride mixtures, followed by sintering. The precursor selection, stoichiometry, and heating rate strongly influence the properties and purity of the MAX phase. Since the discovery of MXenes, extensive research has focused on optimizing MAX phase synthesis, refining etching protocols and improving delamination methods. The addition of excess Ti and aluminum (Al) beyond the stoichiometric ratio during the reactive sintering of Ti3AlC2 enhances the stability and conductivity of the resulting MXene.134,135 This improvement is attributed to the excess Al, which reacts with O in the system, promoting the formation of more stoichiometric Ti3AlC2 with fewer defects, including O substitution in the C sublattice.
Recently, Michałowski et al.136 employed atomic-resolution ultralow-energy secondary-ion mass spectrometry to detect O incorporation in the C sublattice. The study revealed that when a stoichiometric Ti:
Al
:
C ratio is used, Ti3AlC2 can contain up to 30% O in the C sublattice, forming an oxycarbide MXene upon etching. In contrast, using excess metals in synthesis led to Ti3AlC2 with no detectable O. Excessive A-layer content during the MAX phase synthesis reduces O substitution and C vacancies, improving MXene stability and electronic properties.134 Similarly, the composition of X in the MAX phase can influence MXene properties.
For example, Shuck et al.137 used three C sources, graphite, TiC, and C lampblack, to synthesize the MAX phase. The produced MXene exhibited distinct stability and conductivity depending on the composition and morphology. Therefore, choosing elemental precursors, mixing ratios, and sintering conditions is crucial to achieving desirable MAX phase characteristics for CO2RR applications. Future research should explore the influence of these factors, particularly the M, A, and X compositions, on the electrocatalytic performance of MXenes.
MXene synthesis involves the removal of the A layer from the MAX phase to form multilayered MXenes, with the reaction's Gibbs free energy determining the etching effectiveness. This approach is called the top-down approach for MXene synthesis (Fig. 7B). MXenes produced using this approach come in M2XTx, M3X2Tx, M4X3Tx, and M5X4Tx forms (Fig. 7C). Naguib et al.138 synthesized the first MXene by etching Ti3AlC2 with concentrated hydrofluoric acid (HF), selectively weakening Ti–Al bonds while preserving Ti–C bonds. During etching, the initial adsorption of H and F atoms onto Ti atoms selectively weakens the less stable Ti–Al bonds, leaving the stronger Ti–C bonds intact. This process creates interlayer spacing, facilitating the intercalation of HF and H2O for sequential layer-by-layer etching.139 This process exposed undercoordinated Ti metallic surfaces, which were saturated with termination groups, such as –O, –OH, and –F, denoted as Tx.140,141 Since the first reported synthesis of MXenes, significant efforts have been directed toward optimizing synthesis protocols and minimizing the use of hazardous HF. Therefore, numerous alternative methods have been developed, including low HF etching, electrochemical, alkaline, molten salt, and halogen etching approaches.
Low concentrations or reduced quantities of HF yield high-quality MXenes with fewer defects. Acid mixtures have been explored as alternative etching solutions to minimize HF usage while maintaining efficient MAX-phase etching. For example, a comparative study evaluated HF/HCl and HF/H2SO4 etching systems to remove the Al layer from Ti3AlC2 and reported that MXenes synthesized via HF/HCl displayed larger interlayer spacing and higher structural water compared to HF and HF/H2SO4, which might be due to the –Cl termination.142 Recently, an optimized HCl:
HF
:
H2O etching solution has gained widespread adoption due to its reduced HF content and ability to produce high-quality MXene sheets. This approach uses a 6
:
1
:
3 ratio of HCl, HF, and H2O per gram of MAX powder.143
Shuck et al.133 applied the HCl:HF:H2O etching method for the scalable synthesis of up to 50 g of Ti3C2Tx MXene (Fig. 7D). Notably, the large-batch synthesized MXene exhibited identical structural and chemical characteristics to small-batch samples, demonstrating that this approach enables scaling up without compromising material quality (Fig. 7E), making it a promising route for commercialization.
In a three-electrode configuration, the electrochemical etching method selectively removes the A atomic layer by applying a potential while using the MAX phase as an electrode. The applied potential disrupts the M–A bond using electrolyte solutions, such as NaCl, HCl, or HF.144 Effective control over the etching potential and time ensures selective A atom removal, enabling precise control over MXene synthesis. An uncontrollable increase in the applied potential can eliminate the M-layer, yielding amorphous C materials.145,146
Chen et al.147 synthesized Ti3C2Tx via electrochemical etching in a mixed lithium hydroxide (LiOH) and lithium chloride (LiCl) aqueous solution, achieving over 90% etching efficiency. The synthesis was performed at 5.5 V for 5 h, using two identical Ti3AlC2 blocks as symmetric electrodes. As the etching progressed, the Ti3AlC2 cathode remained intact, whereas the Ti3AlC2 anode was partially consumed. The presence of Li+ ions in the etching solution facilitated the etching process by intercalating into the layers and promoting the delamination of the MXene with –Cl surface termination.
Similarly, Shen et al.148 prepared an F-free Ti3C2Cl2 MXene using a molten-salt-assisted electrochemical etching technique. During the electrochemical etching process, the surface termination was modified from –Cl to –O and sulfur (–S), considerably shortening the modification steps and enriching the variety of surface terminations.
Electrochemical etching is a green and safe synthesis method with low energy consumption. However, challenges remain, including forming an amorphous C layer under uncontrolled etching conditions and the relatively low yield of the MXene. Although the MAX phase electrode can be reused multiple times, the typical etching process results in limited MXene production, making it unsuitable for large-scale synthesis. Despite these limitations, this approach can potentially prepare MXenes for the CO2RR.
Alkaline etching is a nonacid etching method for synthesizing MXenes with functional groups, such as –OH and –O, making it hydrophilic and suitable for fabricating electrodes for aqueous applications, such as the CO2RR. Despite the limited toxicity of this method, MXene synthesis using this approach is challenging because the reaction is spontaneous at elevated temperatures, making it difficult to control oxidation at elevated temperatures and low concentrations. The fast oxidation is attributed to –OH termination groups that oxidize MXenes quickly.
For example, Li et al.149 successfully prepared Ti3C2(OH)2 MXenes using KOH in a hydrothermal reactor. Replacing the Al atoms with –OH groups allows the formation of 2D Ti3C2(OH)2. Similarly, NaOH-assisted hydrothermal alkali etching at 270 °C yielded Ti3C2Tx with a yield of 92% and improved interlayer spacing.150 The primary reaction pathway involves converting Al into Al(oxide) hydroxides, followed by their dissolution in an alkaline medium. Elevated reaction temperatures and concentrated NaOH facilitate the rapid dissolution of Al(oxide) hydroxides, forming F-free MXenes with abundant –OH and –O surface terminations. The abundant –OH terminations on MXenes synthesized using this method may enhance C2 product formation.
Literature reports indicate that the presence of –OH terminations on Cu catalysts can lower the binding energy of CO and improve the charge equilibrium between C atoms in the adsorbed OCCO intermediate. This interaction reduces the energy barrier for C2H4 formation by facilitating CO dimerization.151–153
Molten salt etching can be divided into fluoride-containing and fluoride-free molten salt etching.154 The fluoride-containing approach allows the in situ formation of HF during synthesis. By combining a strong acid (e.g., HCl or H2SO4) and fluoride salts (e.g., LiF, NaF, KF, NH4F, and FeF3) or using bifluoride salts (e.g., NH4HF2, NaHF2, and KHF2), HF can be formed in situ during the etching process.154,155 During synthesis, metal cations (Li+, Na+, and K+) enter negatively charged MXene layers and increase interlayer spacing, eliminating the need for an extra intercalation step. The etching temperature and concentration of the acid and fluoride salt can significantly affect the quality of MXene sheets.
Wang et al.156 reported the synthesis of an MXene with an accordion-like structure by introducing LiF into a NaCl–KCl molten salt etchant with CuCl2. The reaction was kept for 5 h at 750 °C, and monolayer flakes of MXene nanosheets were synthesized by incorporating the prepared MXene into a tetrabutylammonium hydroxide (TBAOH) solution during agitation, achieving an MXene yield of ≈15% to 20%.
The Lewis acidic molten salt etching approach is proposed to prepare fluoride-free MXenes by adjusting MAX precursors with various Lewis acid salts at elevated temperatures. For example, Li et al.157 used an element replacement approach that replaces the A-layer atom in the MAX phase with Zn atoms in molten ZnCl2. This approach synthesizes and etches several MAX phases, including Ti3ZnC2, Ti2ZnC, Ti2ZnN, and V2ZnC, to achieve their respective MXene derivatives. The MXene synthesis was achieved by the MXene Al-MAX with ZnCl2 in a 1:
6 molar ratio and heating at 550 °C for 5 h to achieve pure Cl-terminated MXene sheets.
Li et al.105 expanded this approach using Lewis acidic etching to synthesize MXenes from the MAX phase precursors containing A elements, such as Si, Zn, and gallium (Ga; Fig. 8A). As illustrated in Fig. 8B, this approach can prepare MXenes with other types of A-layer atoms. By tuning the chemistry of the MAX precursor and the composition of the Lewis acid melt, a direct redox interaction between the A element and the cation of the Lewis acid molten salt enables the prediction of MAX phase reactivity in the molten salt, facilitating MXene synthesis. For instance, Ti3SiC2 was immersed in molten CuCl2 at 750 °C. During the reaction, the exposed Si atoms, which are weakly bonded to Ti in the Ti3C2 sublayers, were oxidized to silicon (Si4+) cations using Lewis acid Cu2+, forming volatile SiCl4 and the concomitant reduction of Cu2+ to metallic Cu. The metallic Cu was removed by immersing the Ti3C2Cl2 product in an ammonium persulfate solution. High-resolution scanning transmission electron microscopy (STEM) revealed that the resulting MXene exhibited a lamellar microstructure similar to that of the MXene synthesized using the HF etching approach (Fig. 8C). Considering the high activity of Cu for the CO2RR, this research should optimize this approach to preserve Cu metallic particles in the resulting MXene structure and test it for the CO2RR.
![]() | ||
Fig. 8 (A) Schematic of Ti3C2Cl2 synthesis via the Lewis acid etching route. (B) Gibbs free energy mapping (700 °C) guiding the selection of Lewis acid Cl salts based on the electrochemical redox potential. (C) High-resolution transmission electron microscopy imaging of an MXene. (A)–(C) Reproduced from ref. 158 with permission from Springer Nature, Copyright [2020]. (D) Etching MAX phases in Lewis acidic molten salts and atomic-resolution high-angle annular dark-field (HAADF) image of Ti3C2Br2. (E) HAADF images of Ti3C2Te, Ti3C2S, and Ti3C2NH2 MXenes. (B) and (E) Reproduced from ref. 159 with permission from American Association for the Advancement of Science, Copyright [2020]. (F) Schematic of the reaction zone and proposed mechanisms of CVD-Ti2CCl2. (G) Scanning transmission electron microscopy (STEM) images of spherical MXenes. (H) High-resolution image of spherical MXenes. (F)–(H) Reproduced from ref. 160 with permission from American Association for the Advancement of Science, copyright [2023]. (I) Schematic of a fluidized reactor bed for the CVD process. (J) X-ray diffraction pattern of Ti2CCl2 obtained using FBR-CVD before delamination. (K) Cross-sectional HAADF-STEM image of the (0001) plane of Ti2CCl2 flakes. (I)–(K) Reproduced from ref. 161 with permission from Elsevier, Copyright [2024]. |
Kamysbayev et al.159 employed substitution and elimination reactions in molten organic salts to synthesize MXenes with varied surface termination groups using CdCl2 or CdBr2 molten salts (Fig. 8D). The study demonstrated that Cl− or Br− terminated MXenes can actively participate in surface reactions, where halide ion exchange enables precise control over the surface chemistry and properties of MXene sheets. The surface terminations of the synthesized MXenes were further modified by dispersing them in molten alkali-metal halides, such as Li2Te, Li2S, Li2Se, Li2O, and NaNH2 (Fig. 8E), allowing for tailored functionalization and enhanced material properties.
Recently, Liu et al.162 prepared Ti3C2Tx (Tx: Cl− and O−) via a molten-salt-etching route in acetonitrile-based electrolyte. Moreover, CuCl2 was applied as the main molten salt etching, and NaCl/KCl was employed as a supporting electrolyte. The synthesis was performed at 680 °C for 24 h in an argon (Ar)-filled furnace. The obtained MXene displayed enhanced electrochemical stability.
The literature has reported several new top-down MXene synthesis approaches, including hydrothermal-assisted HCl etching163 and microwave-assisted molten salt etching.164–166 These methods are still in their early stages and require more investigation to optimize them to produce high-quality MXenes.
Most studies on MXene synthesis are based on a top-down approach. Three bottom-up synthesis approaches, chemical vapor deposition (CVD), the template method, and plasma-enhanced pulsed laser deposition, have been reported for synthesizing MXenes. The CVD method allows the growth of ultrathin MXene sheets at elevated temperatures and a nonterminated surface.
Xu et al.167 produced defect-free molybdenum carbide (Mo2C), tungsten carbide (WC), and tantalum carbide (TaC) thin films using CVD at elevated temperatures (1085 °C). Recently, Wang et al.160 synthesized Ti2CCl2 using CVD. The reaction of CH4 and TiCl4 on a Ti surface enables direct CVD growth at 950 °C of Ti2CCl2 carpets and complex spherulite-like morphologies that form via buckling and the release of the MXene carpet to expose a fresh surface for further reactions (Fig. 8F). Scanning electron microscopy imaging revealed that the synthesized MXene evolved from bulges into spherical MXene vesicles (Fig. 8G) with sheets radiating from the center and oriented normally to the surface (Fig. 8H). The template method uses 2D TM oxide (TMO) nanosheets as templates. During the synthesis, the TMO nanosheets are carbonized or nitrided to form a carbide or nitride MXene, respectively.168
For example, Xiao et al.169 produced molybdenum nitride (MoN) using 2D molybdenum trioxide (MoO3) nanosheets as templates. MoN was synthesized by annealing the MoO3 cover with NaCl at 280 °C for 2 h. MoN exhibited very uniform nanosheets with a thickness of about 0.71 nm. The 2D tungsten nitride and vanadium nitride nanosheets were also synthesized using this method. Plasma-enhanced CVD and pulse laser deposition can be combined to prepare 2D MXenes.
Zhang et al.170 prepared an ultrathin large-area Mo2C film on sapphire by combining plasma-enhanced CVD and pulse laser deposition. The sapphire substrate was heated to 700 °C for depositing the high-quality Mo2C film using CH4 plasma as the C source.
Xiang et al.161 reported a scalable gas-phase technology for synthesizing Cl-terminated Ti2CCl2. The synthesis was conducted in a fluidized bed CVD reactor, where TiCl3 was introduced into the reactor at 770 °C and was rapidly sublimated to form a gaseous precursor for nucleation (Fig. 8I). The gaseous precursors were transported by Ar gas to react with CH4 in the upper region of the fluidized reactor bed, forming TiCCl2 powders. The process yielded about 0.1 kg per batch, underscoring the high efficiency of the synthesis method. X-ray diffraction (Fig. 8J) and atomically resolved high-angle annular dark-field (HAADF)-STEM images (Fig. 8K) revealed the typical characteristics of Ti2CCl2.
Following the initial discovery of MXenes in 2011, Mashtalir et al.173 demonstrated in 2013 that multilayer MXenes could be delaminated into monolayer MXene nanosheets via DMSO intercalation. After delamination, the X-ray diffraction analysis revealed that the interlayer spacing of the Ti3C2Tx MXene increased from 1.95 to 3.50 nm, reducing van der Waals interactions and promoting exfoliation via ultrasonication. Delamination into the monolayer MXene increases surface terminations and enhances the hydrophilicity and the negatively charged surface of MXene nanosheets, facilitating their dispersion and the formation of stable colloidal solutions. Although DMSO has promising results as an intercalation molecule for Ti3C2Tx, it is not effective for delaminating other MXene types.
Hydrazine monohydrate, N,N-dimethylformamide, and urea have been explored as intercalation agents to exfoliate multilayer Ti3C2Tx into a monolayer Ti3C2Tx MXene.173 However, these methods have demonstrated limited efficiency due to the aggregation of monolayers, resulting in thicker flakes (20 to 50 nm). Naguib et al.174 demonstrated that TBAOH, hydroxyl choline, and n-butylamine could facilitate the delamination of V2Cx and Ti3CNTx from their multilayered structures into single layers via simple handshaking in water. Following intercalation and delamination, the interlayer spacing increased from 2.14 to 3.86 nm for Ti3CNTx and from 1.99 to 3.86 nm for V2CTx. The similarity in interlayer spacing for both MXenes highlights the crucial role of intercalating molecules in determining the final interlayer spacing of delaminated MXenes.
Han et al.175 demonstrated that hydrothermal-assisted intercalation of TMAOH can efficiently intercalate multilayer Ti3C2Tx, increasing the monolayer Ti3C2Tx MXene yield to over 73% while achieving a thickness of 1.7 nm. The hydrothermal-assisted intercalation process facilitates diffusion and incorporation of TMAOH between layers. Ascorbic acid was introduced as a mild reductant to prevent the oxidation of the MXene at elevated temperatures.
With the continuous expansion of the 2D MXene family, alternative delamination solvents have been explored for various MXene compositions. Montazeri et al.176 applied NaOH to intercalate Na+ ions into Nb2CTx and Mo2Ti2C3Tx multilayers following a washing step with TBAOH. The resulting delaminated MXenes exhibited increased d-spacing values of 1.6 and 1.5 nm for Nb2CTx and Mo2Ti2C3Tx, respectively. In addition to assisting with delamination, NaOH also reduced the surface oxidation of the flakes. Similarly, Mashtalir et al.177 reported amine-assisted delamination of the Nb2C MXene, where the intercalation of isopropylamine between the Nb2CTx layers followed by mild sonication in water for 18 h at room temperature led to successful exfoliation. X-ray diffraction analysis of the delaminated Nb2CTx revealed an increased interlayer spacing of 1.23 nm, which is sufficiently large to accommodate more than one isopropylamine molecule and water between layers.
In addition, LiCl can be employed as an intercalator for multilayer MXenes to enlarge their interlayer spacing by inserting Li+.143 Zhang et al.178 applied LiCl as an etchant to delaminate Ti3C2Cl2, synthesized via Lewis acid molten salt etching. A primary challenge associated with Lewis acid molten salt etching is the difficulty of achieving monolayer nanosheets due to the hydrophobic nature and strong interlayer interactions of halogen-terminated MXenes. Delamination was accomplished using a LiCl-assisted DMSO intercalation approach, where the sample was treated for 24 h followed by centrifugation. As Ti3C2Cl2 is hydrophobic, hydrated cations struggle to intercalate between layers. The experiment was conducted in a moisture-free environment to avoid forming a hydration shell around Li+, which could hinder intercalation.179
Notably, the HCl/LiF molten salt etching approach eliminates the need for additional intercalation because MXenes synthesized via this method can be directly delaminated. This result is attributed to the spontaneous insertion of Li+ into interlayers, expanding and weakening interlayer interactions, facilitating separation into monolayers via ultrasonication or simple shaking by hand.180
Song et al.181 proposed a freeze-sonication delamination strategy for exfoliating a multilayer MXene into a monolayer MXene with the yield exceeding 74%. This approach applies the synergistic effect of ultrasonic treatment and ionic intercalation, facilitating the penetration of numerous water molecules into the interlayer space. The volume expansion, followed by ultrasonic treatment in a frozen state, forms the monolayer MXene. The delaminated MXene demonstrated an excellent gravimetric capacitance of 261.1 F g−1 and satisfactory cycling stability. Many of the reported delamination approaches produced MXenes with smaller flakes, limiting their application for large MXene flakes.
Method | Scalability and throughput | Termination control | Advantages | Industrial suitability | Key limitations |
---|---|---|---|---|---|
HF etching141 | High (tens of grams per batch) | Poor (–F dominated) | Simple, fast, widely adopted | Mature, used in academic and lab-scale work | Highly toxic; uncontrolled terminations; defect formation |
HCl–HF–H2O/LiF–HCl105,133 | Moderate (∼5–15 g per batch) | Moderate (mixed –F/–OH) | Safer than pure HF; better delamination via Li+ | Feasible at lab and pilot scales (∼50 g) | Involves fluoride; sensitive to processing parameters |
Electrochemical etching147 | Low (batch); moderate (flow reactors: ∼10 g h−1) | High (tunable –O, –Cl) | Fluoride-free; selective and tunable; potential-controlled | Promising for industrial translation | Lower yield; possible defect formation; scaling challenges |
Alkaline etching149,150 | Moderate (gram-scale hydrothermal setups) | Good (–O, –OH rich) | Fluoride-free; environmentally benign; improved hydrophilicity | Feasible for specific MAX phases | Limited to few MAX phases; low efficiency |
Molten salt etching162,184 | Moderate to high (up to tens of grams per run) | Moderate (–Cl, –O, –S possible) | Fluoride-free; high crystallinity; controlled terminations | Pilot studies reported | Requires >600 °C; substrate dependent |
Halogen etching159,161 | Moderate–high (∼0.1 kg per batch via a CVD or halide vapor route) | Tunable (–Cl, –Br, –I) | Unique terminations; dry etching; scalable in CVD setups | Emerging technique | Toxic gases; process maturity is low |
As MXene synthesis scales toward industrial production, reducing structural defects and chemical impurities becomes vital for preserving material performance and enabling application-specific functionality.133 Several complementary strategies have been proposed to mitigate these challenges throughout the synthesis workflow. Continuous-flow etching systems offer more homogeneous reaction environments than static batch processing, ensuring consistent exposure of MAX phases to etchants and minimizing local concentration gradients that can induce uneven etching.185 This leads to improved uniformity in layer thickness and flake size.133,185 Additionally, precursor engineering, such as stoichiometric optimization of MAX phases with slight excess of Ti or Al, can suppress intrinsic carbon vacancies and oxygen substitutions that degrade the final MXene quality.134,137 The choice of etching media significantly affects structural integrity and termination control. Mixed-acid systems, such as the commonly used 6:
1
:
3 HCl
:
HF
:
H2O ratio, offer safer and more reproducible etching than pure HF, reducing structural collapse and enhancing monolayer yield.186 Shuck et al.187 demonstrated that this protocol could be scaled to synthesize up to 50 g of Ti3C2Tx MXene per batch, while maintaining flake morphology and surface chemistry similar to those obtained in small-scale syntheses. Among the emerging approaches, soft delamination has shown significant promise.188,189 This method eliminates ultrasonication, shaking, or centrifugation while separating MXene sheets. Instead, the intercalated MXene mixture is left undisturbed for ∼30 minutes, allowing gravity-assisted flake separation and the formation of a dark colloidal suspension. Although this method tends to yield more bi- and tri-layered flakes and is relatively slow, it produces large, low-defect flakes ideal for catalytic applications where surface integrity is critical.188,189
Finally, storage and environmental stability must be considered, as freshly synthesized MXenes are highly susceptible to oxidation and hydrolysis when exposed to air or moisture. Best practices include storing MXenes under inert atmospheres (e.g., Ar or N2), applying freeze-drying techniques to avoid hydrolysis, or using protective encapsulation (e.g., polymer or carbon coatings) to retain surface activity before electrochemical deployment.190,191 Collectively, these defect-mitigation strategies, spanning reactor design, etchant optimization, purification, delamination, and post-synthesis stabilization, are critical for producing high-quality, scalable MXenes suitable for the industrial CO2RR and other energy applications.
DFT calculations further indicate that CO2 interacts with MXene surfaces via physisorption, driven by noncovalent interactions, or chemisorption, involving covalent bonds with surface metal sites. The computed Gibbs free energy (ΔG) for chemisorption varies substantially depending on the MXene composition, ranging from −3.19 to −1.29 eV, as summarized in Table 4. These energetic variations, combined with the electronic effects introduced by different surface terminations, underscore the importance of functional group engineering in tuning MXene-based catalyst activity and selectivity toward the targeted CO2RR products.120,201 These energetic variations and the electronic effects introduced by different surface terminations underscore the critical importance of tailoring MXene functional groups to direct specific CO2RR pathways. Table 5 compares the effects of MXene surface terminations on CO2RR intermediates and product selectivity.
Species/M3C2 | Group IV | Group V | Group VI | |||||
---|---|---|---|---|---|---|---|---|
Ti3C2 | Zr3C2 | Hf3C2 | V3C2 | Nb3C2 | Ta3C2 | Cr3C2 | Mo3C2 | |
*CO2 | −0.59 | 0.17 | 0.18 | 0.29 | 0.35 | 0.25 | 0.15 | |
**CO2 | −3.01 | −3.19 | −3.05 | −1.47 | −1.60 | −2.30 | −1.29 | −2.11 |
**OCHO | −2.04 | −2.25 | −2.89 | −1.40 | −1.71 | −1.58 | −1.61 | −1.74 |
**HOCO | −2.06 | −2.49 | −2.79 | −1.41 | −1.54 | −1.92 | −1.74 | −1.91 |
**OCH2O | −3.51 | −4.08 | −4.31 | −1.93 | −2.22 | −2.86 | −1.60 | −1.64 |
**HCOOH | −1.01 | −2.19 | −2.47 | −0.15 | −0.12 | −0.32 | 0.01 | −0.78 |
**CO | −1.18 | −1.11 | −1.54 | −1.45 | −1.39 | −1.80 | −2.00 | −2.27 |
**HOCH2O | −2.47 | −2.82 | −3.11 | −1.59 | −1.88 | −2.56 | −1.85 | −2.15 |
**HOCH2OH | −1.09 | −1.07 | −3.56 | −0.61 | −0.68 | −1.10 | −0.69 | −0.90 |
**H2CO | −2.43 | −3.21 | −3.31 | −1.81 | −2.16 | −2.39 | −1.78 | −1.86 |
**C2OH | −1.58 | −2.05 | −1.89 | −1.26 | −1.36 | −1.94 | −1.51 | −1.64 |
**CH3O | −2.93 | −3.12 | −3.33 | −2.20 | −2.36 | −2.98 | −2.12 | −2.53 |
**CH2 | −1.81 | −1.88 | −1.21 | −0.97 | −1.47 | −2.33 | −1.65 | −2.11 |
**CH3OH | 0.06 | 0.07 | 0.01 | 0.33 | 0.08 | 0.22 | 0.33 | 0.17 |
**O | −4.80 | −5.24 | −5.23 | −3.45 | −3.95 | −4.27 | −3.53 | −3.57 |
**CH3 | −2.04 | −2.37 | −2.86 | −2.26 | −2.47 | −3.27 | −2.52 | −2.98 |
**OH | −4.59 | −4.65 | −4.80 | −3.70 | −3.90 | −4.35 | −3.73 | −3.91 |
**CH4 | −1.18 | −0.70 | −0.94 | −2.15 | −0.77 | −0.82 | −0.55 | −0.70 |
**H2O | −3.03 | −2.96 | −3.04 | −2.46 | −2.37 | −2.64 | −2.55 | −2.88 |
Termination | Representative MXenes | Key intermediaries | Favored product(s) | Mechanistic insight |
---|---|---|---|---|
–O203,204 | Ti3C2O2, Ti2CO2 | *COOH, *HCOO | CO, HCOOH | Strong binding of O-bound species; promotes proton-coupled electron transfer |
–OH120,205,206 | Ti3C2(OH)2 | *COOH | CO | Increases local proton availability; improves surface hydrophilicity |
–F207–210 | Ti3C2Fx | *CHO, weak *CO2 | CH3OH, CH4 (low rate) | Weakens CO2 adsorption; shifts path to *CHO; often lowers activity |
–Cl209 | Ti3C2Cl2 | Not fully studied; charge effects | (under study) | Alters electronic distribution; stabilizes some adsorbates; promising but immature |
Among TM-based MXenes, Group IV (e.g., Ti3C2, Zr3C2, and Hf3C2) exhibits a stronger binding affinity for CO2 compared to Group V (e.g., V2C and Nb2C) or VI (e.g., Mo3C2 and Cr3C2).120 This trend can be attributed to the TM atoms' electronic configuration and d-band center, influencing the overlap between metal orbitals and CO2 antibonding orbitals.202,211,212 For example, Mo3C2 and Cr3C2 MXenes preferentially interact with CO2 over H2O, making them highly promising for the CO2RR in aqueous environments. The mechanistic pathway for CH4 formation on Mo3C2 involves successive hydrogenation steps of intermediates, such as OCHO*, ˙OCH2O*, and HOCH2O*, yielding CH3O* and CH4 as products (Fig. 9A). The DFT calculations suggest that CH3O* is thermodynamically favored over H2COH* during the fifth H+/e− addition, directing the reaction away from CH3OH and toward CH4 formation.
![]() | ||
Fig. 9 (A) CO2 conversion mechanism into *CH4 and **H2O catalyzed by Mo3C2 (Reprinted with permission from ACS Nano 2017, 11, 11, 10825–10833. Copyright [2017] American Chemical Society). (B) Volcano-type relationship between UL and the Bader charge of *HCOOH on M2C (Reprinted with permission from Ind. Eng. Chem. Res. 2023, 62, 48, 20716–20726. Copyright [2023] American Chemical Society). (C) Reaction mechanism of electrochemical reduction of CO2 on O-terminated MXenes (Reproduced from ref. 211 with permission from Royal Society of Chemistry, Copyright [2022]). (D) and (E) Density of states (DOS) corresponding to d-orbitals of the adsorbed dual-atom, where the d-band center is denoted by dashed lines. (F) Adsorption energy of *CH3CO and *CH4 intermediates on dual-atom catalysts/Ti2CO2. (D)–(F) Reproduced from ref. 213 with permission from Royal Society of Chemistry, Copyright [2025]). |
Moreover, M2C MXenes, such as V2C and Cr2C, have demonstrated potential for HCOOH production, as evidenced by a volcano-type relationship between the adsorption strength and catalytic activity (Fig. 9B). This relationship suggests that intermediate adsorption strengths facilitate efficient CO2-to-HCOOH conversion. Excessively strong adsorption impedes intermediate desorption, whereas weak adsorption hinders activation, demonstrating the critical role of adsorption energy in catalytic performance.121,214–216
The surface functionalization of MXenes significantly enhances their catalytic performance by modifying electronic properties, adsorption energies, and reaction pathways.217,218 Oxygen-terminated MXenes (e.g., Ti2CO2 and V2CO2) are effective for the CO2RR because the pathway to *HCOOH is preferred over the *CO pathway due to the stabilizing effect of –O groups on reaction intermediates (Fig. 9C).211 Oxygen vacancies, forming during the reaction, further improve selectivity by stabilizing transition states and creating active sites for intermediate binding. For instance, O-terminated MXenes have been reported to preferentially catalyze CO2 to HCOOH with reduced overpotentials, applying the accessibility of H-coordinated mechanisms over C-coordinated pathways.219 In addition, Sc2C(OH)2 and Y2C(OH)2 are promising candidates for CH4 production due to the reactive H atom in the –OH group. This reactive H atom facilitates stable intermediate formation, lowering overpotentials and enhancing selectivity for CH4 formation.205 Fluoride-containing terminations influence product pathways differently. Furthermore, Ti3C2 MXenes with –F terminations tend to favor a path involving formaldehyde intermediates, forming CH3OH. In contrast, –F-free MXenes follow a distinct mechanism, producing HCOOH and methylene glycol, decomposing into CH3OH and water.220
Sun et al.213 used DFT to evaluate the activity and selectivity of a dual-atom-modified MXene catalyst for the CO2RR to C2H6O. They demonstrated that the Co–Co dual-atom catalyst, with its asymmetric C–C coupling mechanism, achieves high catalytic activity due to its moderate d-band center, optimally balancing electron occupancy in antibonding orbitals, ensuring efficient adsorption of reaction intermediates. The d-band analysis revealed that vanadium (V)–V and chromium (Cr)–Cr dual-atom catalysts, characterized by higher d-band centers, exhibit stronger adsorption of intermediates (*CH3CH2O and *CH2OHCH2O), making hydrogenation steps energetically demanding, with energy barriers of 0.81 and 0.93 eV, respectively (Fig. 9D). In contrast, manganese (Mn)–Mn, iron (Fe)–Fe, and cobalt (Co)–Co dual-atom catalysts have d-band centers positioned farther from the Fermi level (Fig. 9E), allowing antibonding orbitals to be more readily occupied by electrons. This positioning weakens the adsorption of intermediates (Fig. 9F), facilitating smoother reaction pathways without excessive energy barriers and increasing chemical reactivity.221
Table 6 presents innovative MXene-based electrocatalysts, their electrocatalytic activities, working electrolytes, performance, and CO2RR products, further driving innovation in MXene-based electrocatalysts toward the CO2RR.
Electrocatalyst | Potential | Current density (mA/cm2) | Electrolyte | Faradaic efficiency (%) | Prod. | Ref. |
---|---|---|---|---|---|---|
Pd-MXene | 0.5 V | 17 | CO2-saturated 1.0 M KHCO3 | 67.8 | CH3OH | 122 |
SA-Cu-MXene | −1.4 V vs. RHE | −21.3 | 0.1 M KHCO3 | 59.1 | CH3OH | 106 |
CdS/Ti3C2 | −1.0 VRHE | ∼6.4 | 0.1 M KHCO3 | 94 | CO | 222 |
VS-CdS/Ti3C2 | −1.0 VRHE | ∼−6 | 0.1 M KHCO3 | 96 | CO | |
MxOy/MAX hybrid | −0.4 to −0.6 V | 2.4 | 0.5 M NaHCO3 | 67 | CO | 223 |
Ag–ZnO/Ti3C2Tx | −0.87 VRHE | 22.5 | 0.5 M KHCO3 | 98 | CO | 224 |
d-Ti3C2Tx | −2.2 V vs. SCE | −1.5 | Acetonitrile, 1 ethyl-3 methylimidazolium tetrafluoroborate EMIMBF4 | 65 | CO | 108 |
d-Mo2CTx | −2.2 V vs. SCE | −2.5 | Acetonitrile, 1 ethyl-3 methylimidazolium tetrafluoroborate EMIMBF4 | 90 | CO | 108 |
Cu–Pd/MXene | −2.8 V | 150 | 0.1 m KHCO3 | 93 | Formate | 127 |
SnO2/MXene | 1.1 V | −57.8 | CO2-saturated 0.1 M KHCO3 | 94 | Formate | 121 |
ZnO–Fe-MXene | 1.0 V | 18.745 | 0.5 M NaOH | — | Formate | 225 |
Cu-/Ti3C2Tx | −1.5 V vs. Ag/AgCl | −1.08 | 0.1 M NaHCO3 | 58.1 | HCOOH | 76 |
MXene (Ti3C2Tx) modified with boron-doped diamond | −1.3 V vs. Ag/AgCl | 0.5 M KOH | 97 | HCOOH | 226 |
Liu et al.247 demonstrated the synthesis of MXene/metal–organic framework (MOF) heterostructures through electrostatic attraction, where the surface terminations of the Ti3C2Tx MXene strongly interacted with MOF precursors, forming well-integrated composites that effectively enhanced CO2 adsorption and catalytic selectivity. Similarly, a covalent organic framework (COF)–Ti3C2 heterostructure achieved >90% CO selectivity at −0.6 V vs. RHE due to the synergistic effect of the MXene's large exposed surface area and the COF's catalytic functionalities, substantially suppressing the HER.248 This significant performance was primarily attributed to the large, exposed surface of the MXene, which effectively disperses the COF, endows the heterojunction with more active sites, and facilitates efficient transport channels. The 3D MXene/graphene oxide/perylene diimide aerogel heterostructure was synthesized via impregnation and freeze-drying for the photocatalytic CO2RR.233 The resulting heterostructure featured a large surface area, an enhanced photogenerated carrier, and an electron transfer network facilitated by π–π stacking via electrostatic attraction. This structural configuration, interconnecting the faces of the heterostructure, promotes the efficient transfer of photogenerated electron–hole pairs, enabling rapid carrier movement and separation. Additionally, perylene diimide functions as an electron donor, activating catalytic sites for enhanced photocatalytic CO2RR.
Computational and experimental studies have highlighted the potential of MXenes as promising support materials for facilitating strong metal–support interactions in CO2RR catalysts. The abundant surface-functional groups and metal vacancy defects in MXenes serve as ideal anchoring sites for single metal atoms, primarily due to the high surface energy, adjustable electronic structure, and uniform atomic arrangement of MXenes.214,255–257 The synthesis of MXene-based SACs can be achieved using three strategies: surface adsorption, metal vacancy anchoring, and anchoring at surface-functional group vacancies.258 According to a computational study, the adsorption of single metal atoms is possible on the top, hollow (hcp and fcc), and bridge sites.
Chen et al.259 employed electrostatic adsorption and in situ reduction to synthesize a Co-Ti3C2Tx SAC for photocatalytic CO2 reduction. Co-cations are initially adsorbed onto the negatively charged functional groups on the MXene surface during synthesis, forming ionic bonds facilitated by electrostatic attraction. The ionic bond formation is followed by reducing Co cations on the MXene, leading to covalent bonds between the metal atoms and surface-functional groups (–F and –O). Sodium borohydride was employed as a reducing agent for Co2+ ions and as an oxidation inhibitor of MXenes to overcome its fast oxidation.
Zhou et al.248 designed and synthesized the MXene@Por-COF-Co heterostructure. The dispersed COF structures and exposed MXene nanosheets offer more accessible reactive sites and quicker ion transfer channels to the heterostructure because the covalent interactions between the aldehyde groups in the COF structures and the amino groups of MXene can facilitate the in situ formation of COFs on the surface of amino-functionalized MXene nanosheets. TEM images of MXene@Por-COF-Co-7 reveal a homogeneous distribution of COF nanosheets across the surface of the MXene nanosheets. Furthermore, MXene@Por-COF-Co-7 exhibits a remarkable CO FE of 97.28% at −0.6 V, significantly higher than that of Por-COF-Co (0%) at the same potential. In addition, MXene@Por-COF-Co-7 also maintains a high FE in the potential range of −0.5 to −1 V vs. RHE, suggesting its good selectivity for CO formation. The bias current density of CO increased with a rise in voltage.
Zhao et al.260 employed self-reduction stabilization to anchor platinum (Pt) single atoms onto Ti vacancies of Ti3C2 for CO2 activation with amines and silane, producing formamides. The single Pt atoms on the Ti3−xC2Ty support exhibited partial positive charges and atomic dispersion. Adsorbing and reducing Pt4+ simultaneously were successful without adding a reductant. Moreover, HAADF imaging revealed that the Pt single atoms were anchored at the Ti site in Ti3C2 rather than at the lattice gap (Fig. 10A). These Ti vacancies strongly correlate with the etching conditions during MXene synthesis, especially the etchant. Fig. 10B reveals that Pt1/Ti3−xC2Ty SAC displayed superior catalytic performance for converting CO2 compared to that of Pt nanoparticles (NPs).
![]() | ||
Fig. 10 (A) High-angle annular dark-field (HAADF) scanning transmission electron microscopy (STEM) image of Pt/Ti3−xC2Tx. (B) Catalytic performance of catalyst systems. (A) and (B) Reprinted with permission from J. Am. Chem. Soc. 2019, 141, 9, 4086–4093. Copyright [2019] American Chemical Society). (C) Aberration-corrected (AC) HAADF-STEM image of the SA-Cu-MXene. (D) Faradaic efficiency (FE) of product formation on the SA-Cu-MXene. (E) Current stability and corresponding FE for CH3OH formation on the SA-Cu-MXene. (C)–(E) Reprinted with permission from ACS Nano 2021, 15, 3, 4927–4936. Copyright [2021] American Chemical Society). (F) FE of Cu-SA–Ti3C2Tx. (G) X-ray absorption near-edge structure spectra at the Cu K-edge with CuO, Cu2O, and Cu foil as a reference. (H) Extended X-ray absorption fine structure (EXAFS) fitting curve of Cu-SA/Ti3C2Tx; inset illustrates the Cu-SA–Ti3C2Tx structure. Yellow, blue, dark yellow, red, and white balls represent Cu, Ti, C, O, and H, respectively. (F)–(H) Reproduced from ref. 261 with permission from Springer Nature, Copyright [2021]). (I) AC HAADF-STEM of Ti3C2Tx demonstrating Ti vacancies.262 (J) Potential-dependent FE of H2 and CO on Cu SA@N-Ti3C2Tx at applied potentials. (K) EXAFS fitting of Cu SA@N-Ti3C2Tx; inset presents the atomic interface structure model. (I)–(K) Reproduced from ref. 262 with permission from Springer Nature, Copyright [2024]). |
Zhao et al.106 developed single-atom Cu loaded on MXene layers by selectively etching Al layers from quaternary MAX phases [Ti3(Al1−xCux)C2] for CH3OH synthesis, applying easy sublimation of AlCl3 and leaving unreacted Cu on the MXene. The improved selectivity for CH3OH arises from the capacity of atomically dispersed Cu sites to impede the C–C coupling of *CO, facilitating the formation of CH3OH (Fig. 10C). This coordination lowers the energy barrier for converting HCOOH* into an absorbed CHO* intermediate, enhancing electrocatalytic activity for CO2 conversion. The SA-Cu–MXene catalyst exhibited an increased FE of 59.1% for CH3OH production with high stability and a low energy barrier for the rate-determining step (HCOOH* to CHO*; Fig. 10D and E).
Bao et al.261 synthesized a Cu SAC anchored on Ti3C2Tx nanosheets via chemical reduction, followed by freeze-drying. The resulting Cu-NP/Ti3C2Tx SAC reduced CO well, achieving over 98% selectivity toward C2+ products with a high C2H4 selectivity of 71% (Fig. 10F). The catalyst promotes the formation of the *CO–CHO intermediate, facilitating C–C coupling. The XANES analysis revealed that the Cu valence state in the Cu-NP/Ti3C2Tx SAC lies between that of metallic Cu and Cu+, indicating the presence of O coordination and formation of Cu–O3 species (Fig. 10G and H).
Similarly, Liu et al.262 reported a monoatomic Cu catalyst featuring Cu–N1C1 coordination anchored on the N-doped Ti3C2Tx MXene for the efficient CO2 reduction to CO. This catalyst achieved over 97% selectivity toward CO at an applied potential of −0.58 V vs. RHE (Fig. 10J). The excellent performance was attributed to a potential-dependent valence transition of the Cu species. Aberration-corrected HAADF imaging indicated a high density of Ti vacancies in the MXene lattice, serving as preferential anchoring sites for immobilizing isolated Cu atoms (Fig. 10I). The XANES analysis revealed a negative shift in the Cu absorption edge relative to the pristine Ti3C2Tx, which is indicative of strong electronic interactions between Cu single atoms and the MXene support (Fig. 10K).
In developing SACs, a primary consideration is their stability because several factors, such as decomposition, dissolution, and atom migration, can promote cluster formation and morphological degradation. The high surface energy and mobility of isolated atoms drive these transformations.263 Instability of SACs often results in diminished current density and FE.264,265 Although many studies on MXene-based SACs for the CO2RR have emphasized activity and selectivity, the stability of the single atoms under reaction conditions remains underexplored.
Future investigations should prioritize evaluating catalyst stability by conducting detailed post-reaction analyses. Such efforts could provide crucial insight into the structural and chemical evolution of catalysts, facilitating the rational design of MXene–single-atom interactions for enhanced catalytic activity, durability, and selectivity.
DFT has been pivotal in advancing 2D material-based electrocatalysis by predicting performance under several atomic configurations.82,266,267 This theory has been critical in developing SAC-supported 2D material-based electrocatalysts. Numerous studies have employed the DFT to design SAC-MXene-based catalysts for the CO2RR. For instance, Li et al.258 demonstrated the high catalytic activity of single-atom scandium (Sc), Ti, and V-supported Ti2CN2 to produce CO with a low overpotential of 0.37, 0.27, and 0.23 eV, respectively. In contrast, Mn and Fe supported on Ti2CN2 primarily produce HCOOH with a low overpotential of 0.32 and 0.43 eV. The high catalytic activity of single atoms on Ti2CN2 can be attributed to N-functionalization, stabilizing SACs effectively by anchoring TM atoms. This functionalization also lowers the energy barrier for CO2 reduction and improves catalytic selectivity. These SACs on Ti2CN2 exhibit high catalytic activity with much lower overpotentials.
Similarly, Athawale et al.234 explored the feasibility of MXenes serving as an anchoring site for isolating TM SACs for the CO2RR. Several SAC systems containing 3d (Sc, Ti, V, Cr, and Mn), 4d (yttrium [Y], zirconium [Zr], niobium [Nb], and Mo), and 5d (hafnium [Hf]) transition metals, supported on an O-terminated MXene (TM@Ti2CO2), were designed using DFT calculations. The findings indicate that TMs anchored on top of the C atom of Ti2CO2 (hollow-C site) present the most stable configuration. The hollow-C site, primarily for Nb, Mo, Zr, V, Cr, and Ti atoms, exhibits the most negative Eb values, indicating higher stability and better suitability for the CO2RR.
In the electrocatalytic domain, Liu et al.276 reported a hybrid composed of TiO2 and SnO2 nanowires self-assembled onto the Ti3C2 MXene via van der Waals interactions. The MXene suppressed MMO aggregation provided efficient electron pathways and improved structural integrity. This integration enhanced CO2 adsorption and activation, significantly improving CO2RR activity.277
Hao et al.224 fabricated an Ag–ZnO/Ti3C2Tx hybrid catalyst via a cation exchange and self-assembly method—the MXene-regulated interface featured undercoordinated sites and mesoporous nanostructures. The catalyst achieved nearly 100% CO FE and a partial current density of 22.59 mAcm−2 at −0.87 V vs. RHE (Fig. 11A). DFT calculations confirmed that MXene addition shifted the d-band center, enhanced *COOH binding, and lowered the energy barrier for intermediate formation (Fig. 11B). In situ attenuated total reflectance infrared spectroscopy offered valuable insight into the reaction intermediates, elucidating the origin of CO selectivity (Fig. 11C and D). Within the applied potential window of −0.123 to −1.423 V, the attenuated total reflectance infrared spectra revealed characteristic signals corresponding to CO2, CO32−, and adsorbed *CO species. The *CO bands exhibited a bipolar profile, indicative of Fano line shape modulation. For the Ag–ZnO/Ti3C2Tx catalyst, the inverted band observed at 1920 cm−1 was attributed to linearly bonded CO (COL), whereas the positive band at 1800 cm−1 was assigned to bridge-bonded CO (COB). These observations suggest that forming *CO intermediates from CO2 is more favorable on Ag–ZnO/Ti3C2Tx than on Ag–ZnO. DFT revealed that adding the MXene facilitated stronger binding abilities of *COOH compared to Ag–ZnO. Moreover, the MXene regulated the Ag–ZnO interface by reducing the electron filling of antibinding sites and optimizing the electronic structure by lowering the formation energy barrier of the intermediate.
![]() | ||
Fig. 11 (A) Faradaic efficiency (FE) of CO on Ag–ZnO–Ti3C2Tx. (B) Density of states (DOS) for Zn 3d, Ag 4d, and O 2p orbitals of Ag−ZnO and Ag−ZnO/Ti3C2Tx (left pane) and the corresponding d-band center (right pane). (C) and (D) In situ attenuated total reflectance infrared spectra while stepping the potential from −0.123 to −1.423 V on (C) Ag−ZnO and D) Ag−ZnO/Ti3C2Tx. (A)–(D) Reproduced from ref. 224 with permission from Wiley and Sons, Copyright [2023]. (E) Low-magnification high-angle annular dark-field scanning transmission electron microscopy image and corresponding electron-dispersive spectroscopy element images of SnO2/MXene. (F) High-resolution transmission electron microscopy image and lattice plane. (G) Sn K-edge X-ray absorption near-edge structure spectra of the SnO2/MXene catalyst, Sn foil, and SnO2 references. (H) Potential-dependent FE of formate. (E)–(H) Reproduced from ref. 121 with permission from PNAS, Copyright [2022]. |
Similarly, Cao et al.232 reported a ZnO/N–Ti3C2Tx catalyst that achieved an FE exceeding 96% for CO production. ZnO provided the primary active sites for CO2 conversion, while the N-doped MXene improved textural properties and conductivity, facilitating PCET. Similarly, SnO2 quantum dots grown on ultrathin Ti3C2Tx MXene sheets (via hydrothermal synthesis) delivered a 94% FE and a partial current density of 57.8 mA cm−2 for formate production. In situ XANES measurements revealed that SnO2 was partially reduced to metallic Sn during operation, which acted as the true catalytic site. Coordination environment changes observed via EXAFS confirmed strong Sn–Ti interfacial coupling unique to the MXene-based hybrid (Fig. 11F–H).121 In addition, the absorption edge position of the Sn K-edge XANES spectrum of the SnO2/MXene is situated between that of metallic Sn foil (Sn0) and SnO2 (Sn4+; Fig. 11G). The extended X-ray absorption fine structure spectra of SnO2/MXene in the R-space and K-space differ from those of pure SnO2, suggesting that the local coordination environment of Sn in SnO2/MXene is unlike that in pristine SnO2.
In a similar study, Yu et al.210 synthesized a TiO2/Ti3C2 MXene photocatalyst via thermal calcination. By adjusting the temperature (350–650 °C), TiO2 nanoparticle loading was modulated, influencing CH4 production. Though photocatalysis, this system again underscores the broader relevance of MXene–MMO interactions across catalytic modalities.
Recently, a Cu2O/MXene 0D/2D hybrid catalyst demonstrated excellent selectivity for propane (C3H8) production at −1.3 V vs. RHE.278 The interface between Cu2O and the MXene created cooperative binding sites; the MXene favored *CO adsorption while Cu2O stabilized *C2 intermediates, facilitating C–C–C coupling and efficient hydrogenation to C3 products. This hybrid highlights the unique potential of MXene–MMO heterostructures in enabling multicarbon product formation, which is rarely achieved using either component alone. These examples illustrate how electronic coupling, vacancy engineering, and interfacial coordination at the MXene–MMO boundary promote enhanced CO2 adsorption, intermediate stabilization (e.g., *COOH and *CHO), and improved selectivity. Moreover, MXenes can help suppress competing hydrogen evolution reactions due to their hydrophilicity and binding site modulation, thereby further enhancing CO2RR selectivity. Future work should focus on understanding how the oxidation states, vacancy density, and structural morphology of MXenes and MMOs evolve under electrochemical operating conditions, and how these changes influence catalytic stability and product distribution. In particular, quantifying the role of MXenes in HER suppression and tailoring interfaces for selective multi-electron/multicarbon pathways are promising directions for designing next-generation MXene-based hybrid electrocatalysts.
Beyond intrinsic catalyst design, the practical implementation of MXene-based electrocatalysts is significantly influenced by reactor-level constraints inherent to industrial electrolyzer systems (as discussed in Section 2.1). Industrial electrolyzers require operation under stringent conditions, including continuous high current densities (>200 mA cm−2), efficient heat and water management, effective control of CO2 crossover, and stable long-term performance. Traditional laboratory-scale H-cell reactors, which dominate fundamental studies, fail to replicate these conditions due to inherent limitations such as low CO2 solubility (∼34 mM), diffusion-controlled mass transport, large electrode spacing, and dilute electrolytes, resulting in low achievable current densities (<100 mA cm−2). Consequently, performance metrics derived from H-cells rarely translate effectively to industrial-scale systems. These reactor designs significantly enhance CO2 transport through direct gas-phase delivery, minimize ohmic losses, control reaction interfaces more effectively, and offer improved electrolyte management, addressing many limitations of conventional H-cell setups.
Nevertheless, a lack of standardized evaluation protocols hinders widespread adoption and advancement in the MXene-based CO2RR. Variations in cell designs, electrode materials, electrolyte compositions, gas flow conditions, and performance metrics currently impede meaningful comparison between studies. Establishing community-wide benchmarking guidelines and uniform testing standards will enhance performance evaluation reproducibility, transparency, and reliability.
Moreover, the catalytic performance of MXenes remains closely tied to their synthesis routes, which influence structural characteristics such as flake size, surface termination chemistry (–O, –OH, –F), and defect density. Oxygen-terminated MXenes have demonstrated promising catalytic properties by lowering reaction barriers and stabilizing critical reaction intermediates. However, achieving reproducible and controlled termination profiles is challenging. To address this, scalable, cost-effective, and precisely controllable synthesis methods must be developed. Additionally, defect engineering and doping offer strategic routes to optimize the catalytic properties of MXenes. Introducing oxygen vacancies or doping with heteroatoms (N, P, or transition metals) can modify the electronic structure, enhance active site densities, and tune intermediate binding energies. However, excessive defects may negatively affect structural stability and induce undesired side reactions. Therefore, careful optimization and systematic evaluation of doping strategies under realistic electrochemical conditions are necessary.
Significant gaps remain between computational predictions and experimental outcomes. Many theoretical studies utilize idealized MXene structures without realistic surface heterogeneities and defects. Experimentally synthesized MXenes typically feature mixed terminations, variable flake sizes, and structural imperfections. Future theoretical efforts should incorporate realistic structural models that account for these variations to enhance predictive accuracy, thereby more effectively guiding experimental development.
Currently, most experimental CO2RR studies focus on Ti3C2Tx MXenes. Exploring under-investigated MXene compositions such as Mo2CTx, Nb2CTx, and V2CTx could uncover unique catalytic properties, improved stability, and enhanced activity. Leveraging machine learning and high-throughput computational screening methods may accelerate the discovery of promising MXene candidates and guide targeted experimental validation.
The operational stability of MXenes under realistic electrochemical conditions remains relatively unexplored. Prolonged exposure to negative potentials, variable pH environments, and continuous gas flow can induce oxidation, structural deformation, and changes in surface termination composition. To address this, in situ and operando characterization techniques such as transmission electron microscopy (TEM), X-ray absorption spectroscopy (XAS), and Raman spectroscopy should be employed to monitor catalyst evolution under reaction conditions. Such insights are crucial for designing MXenes with enhanced durability and reliable long-term performance. Finally, emerging fabrication technologies, particularly additive manufacturing and 3D printing, offer novel opportunities for transitioning MXene catalysts into practical, scalable electrode architectures. Printable MXene inks, already successfully demonstrated in energy storage and electronics, could facilitate customized, high-surface-area electrode designs, enhancing mass transport and reaction interface stability. Combining MXenes with complementary materials through advanced printing techniques further opens new possibilities for scalable integration into commercial CO2RR systems. In summary, the successful industrial deployment of MXene-based electrocatalysts for the CO2RR demands a comprehensive approach that integrates advanced material synthesis and surface engineering, optimized reactor design, standardized evaluation methods, and scalable fabrication techniques.
Footnote |
† These authors contributed equally. |
This journal is © The Royal Society of Chemistry 2025 |