Abstract
MXenes undergo water-mediated surface reconstruction under anodic polarization and spontaneously form single-atom centers, reminiscent of single-atom catalysts (SAC), which exhibit great potential for the oxygen evolution reaction (OER). Yet, the as-formed SAC-type MXene motifs still suffer from the conventional activity-stability trade-off under OER conditions. Here, we propose a nitrogen embedding strategy to modulate the local coordination environment of the single-atom centers, thereby enhancing both their stability under anodic polarization and OER activity. Based on density functional theory calculations, we systematically evaluate 53 N-doped SAC-type MXene motifs, covering M2X1O, M3X2O, and M4X3O MXenes, using a three-step screening approach, which identifies promising candidates that (i) allow nitrogen doping under anodic polarization, (ii) are thermodynamically stable under OER conditions, and (iii) exhibit high OER activity. We demonstrate that N-doped Ta2C1O and Ta2N1O SAC-like MXenes break the traditional activity-stability trade-off in OER and provide theoretical guidance for designing efficient and durable SAC-like MXene electrocatalysts.
Introduction
Single-atom catalysts (SAC) have attracted much attention and made significant advances in the field of catalysis due to their atomically dispersed active sites, almost 100% utilization of metal atoms, and tunable electronic structures1. Notably, SAC exhibit excellent activity and selectivity in various electrocatalytic reactions, including the oxygen reduction reaction (ORR)2, oxygen evolution reaction (OER)3, and hydrogen evolution reaction (HER)4, to name a few. To further improve their performance, various support two-dimensional (2D) materials—such as graphene5, carbon nanotubes6, and graphitic carbon nitride (g-C3N4)7—have been developed to stabilize and tailor the single-atom sites.
In the recent years, the rapid development of two-dimensional (2D) materials8 has substantially broadened the design space for SACs. Among these materials, MXenes—a family of 2D transition metal carbides, nitrides, and carbonitrides, with the general formula of Mn+1XnTx, where M is a transition metal, X denotes carbon and/or nitrogen, n typically ranges from 1 to 4, and Tx represents surface adsorbates such as *O, *H, *OH, or *F, which can be tuned by synthesis conditions—have been extensively studied since the first synthesis of Ti3C2Tx by Gogotsi and coworkers9. MXenes have attracted significant attention due to their unique physicochemical properties, including high electrical conductivity, hydrophilicity, large specific surface area, and excellent electrochemical performance10,11,12. Therefore, MXene-based SACs have also recently attracted much attention, exhibiting promising performances in various electrocatalytic reactions such as for nitrogen reduction (NRR) or hydrogen evolution (HER)13,14. While conventional SACs rely on anchoring a foreign metal atom on the MXene basal plane, by means of ab initio molecular dynamics (AIMD) simulations, Wan et al.15 found that Ti2CO2 MXene undergoes spontaneous water-mediated surface reconstruction under anodic polarization, leading to the formation of single-atom motifs, which are reminiscent of an archetypal SAC active site16. This theoretically predicted phenomenon corresponds well with Gogotsi’s report on the influence of applied potential on interfacial behavior of MXene in water environment17.
The main difference between archetypal single-atom catalysts (SACs) and the MXene-SAC motif lies in their formation mechanisms. Conventional SACs are typically synthesized under non-electrochemical conditions, where foreign metal atoms are introduced and anchored ex situ onto a substrate, thereby forming well-defined active sites. In contrast, the MXene-SAC motif does not require any foreign metal atom during its synthesis, since the single-atom center on the MXene basal plane is formed in situ during anodic polarization, with their formation governed by the electrochemical environment. Since these SAC motifs are formed directly under reaction conditions, they may exhibit adaptive behavior and enable in situ regeneration when pulsing the applied electrode potential back and forth18. This appears to be an advantage over conventional SACs, where in situ regeneration is difficult to achieve. Despite this, the electrochemically formed SAC motifs also exhibit limitations. Their formation is sensitive to the applied electrode potential and their applicability is hitherto restricted to specific two-dimensional materials including MXenes. This differs from conventional SACs, which can be synthesized for a plethora of different support materials.
In a recent communication, Razzaq et al.16 have demonstrated using density functional theory (DFT) calculations that the SAC-like motif of MXenes exhibits promising activity for OER, while Faridi et al. showed that these SAC-like sites are selective toward the chlorine evolution reaction in brine19. Notably, the SAC-like motif of on Mo2C1O and Mo2N1O was predicted to selectively catalyze nitrogen reduction upon pulsing of the electrochemical potential18. These previous studies provide detailed mechanistic insights into various energy conversion processes catalyzed by a novel class of electrochemically formed SAC-like MXenes.
Although the MXene-SAC motif is capable of catalyzing OER with moderate overpotentials, as estimated from DFT calculations16, the stability of single atoms anchored to a support under anodic polarization remains a challenge to date. As pointed out by Di Liberto and coworkers20, the single atom site structure can be influenced by the solvent, pH, and applied electrode potential. This can cause deactivation of active sites and structural instability, such as metal diffusion and aggregation, or the formation of new chemical species through reactions with solvent components induced by demetallization. Meng et al.21 screened the stability of the MXene-SAC motif in the homologous series of M2C1O and M2N1O by the construction of Pourbaix diagrams and demonstrated that particularly Mo2C1O-, Mo2N1O-, Ta2C1O-, and Ta2N1O-based systems are stable in the potential regime of about 1.50 V vs. RHE (reversible hydrogen electrode) relevant for OER. However, based on thermodynamic considerations, it was predicted that the other SAC-like MXene of the M2X1O2 series would dissolve under OER conditions, suggesting that stabilization strategies are needed to enhance the stability of the MXene-SAC motif under anodic polarization to achieve efficient and durable OER catalytic performance.
Various studies have shown that the SAC performance is governed by the local coordination environment22,23, particularly the chemical identity of adjacent dopants and their coordination number24. A modulation of charge distribution and electronic structure can ultimately enhance stability and catalytic activity of SACs. Notably, nitrogen doping has emerged as a predominant strategy for archetypal SACs25. For example, Ha et al.26 demonstrated that N-doped graphene-based SACs exhibit superior catalytic performance in both the HER and OER, compared to their undoped C4-coordinated counterparts. In addition, Liu et al.27 found that introducing nitrogen dopants into graphene-based Ir SACs can significantly enhance its stability, and the bond strength between Ir single atom and graphene can also be improved by increasing the nitrogen content.
Given that a nitrogen coordination environment is also commonly employed in archetypal SACs, we present in this work a strategy to embed nitrogen into the MXene-SAC motifs to modulate the electronic structure and local chemical environment of the single-atom site. It is hypothesized that the nitrogen-doped MXene-SAC motif (cf. Fig. 1b) can suppress unfavorable structure dissolution under high anodic potentials and thus further improve the stability of the electrochemically formed single-atom centers. Nitrogen doping of the MXene-SAC motif also influences OER activity, thus potentially offering a way to solve the activity-stability conundrum in oxygen evolution electrocatalysis28,29.
Side (left images) and top (right images) views of a reconstructed p(3 × 3) unit cell of the MXene (0001) surface with M18X9O9-nNn composition under anodic polarization (n = 3 for N-doped situations), in which a metal atom is displaced from the basal plane to form a single-atom motif: a Undoped 3O*-MSAC-*O, with three neighboring *O groups; b nitrogen-doped 3N*-MSAC-*O, obtained by substituting the three surrounding *O groups with three nitrogen atoms. Surface N and O atoms are represented by light blue and red spheres, respectively, while transition metal M and C atoms are shown as dark blue and brown spheres, respectively.
To date, early transition metal MXenes from groups 4 to 6—including Sc, Ti, V, Y, Zr, Nb, Hf, Ta, and W—have been successfully synthesized30. Moreover, Fe species can also be highly dispersed and stably incorporated into MXenes by tuning the synthesis conditions31. In the present work, we systematically investigate the stability of the MXene-SAC motif, and we differentiate between undoped and nitrogen doped cases. Considering realistic electrochemical environments, MXenes in contact with aqueous electrolytes typically form a mixed *OH/*O-terminated surface32. The corresponding Pourbaix diagram analysis further indicates that the *O groups on the SAC site surface are thermodynamically more favorable than *OH groups under OER conditions32,33. Thus, this study focuses only on the O-terminated SAC-like MXene configuration, in which the undoped MXene-SAC motif is referred to as 3O*-MSAC-*O (cf. Fig. 1a), where a single metal atom is coordinated to the basal plane by three surface *O adsorbates, while the latter *O corresponds to the intermediate bound to the SAC site above. Similarly, the N-doped MXene-SAC motif is referred to as 3N*-MSAC-*O (cf. Fig. 1b), where the three surface *O groups are replaced by *N. By building on the existing theoretical support for the formation of the 3*O–MSAC–*O motif of the M2X1O systems, we extend here the investigation to M3X2O and M4X3O MXenes34, where M = Cr, Fe, Hf, Mo, Nb, Ta, Ti, V, W, or Zr, and X = C or N, as usual. Note that during the structural optimization of the 3*O–MSAC–*O motif, some MXenes (such as Cr3C2O, Cr4C3O, Fe2N1O, Fe3X2O, and Fe4X3O) exhibited instabilities; thus, these systems were not further considered. As a result, a total of 53 N-doped MXene-SAC systems were selected as initial candidates for further investigation.
Results
To systematically screen nitrogen-doped SAC-like MXenes as catalysts for the OER, we employ a stepwise funneling strategy based on thermodynamic and activity-related criteria, as shown in Fig. 2. This procedure allows us to progressively narrow down a large number of MXene-SAC motifs systems to a few highly promising candidates. The screening process consists of three fundamental steps: (i) evaluating the thermodynamic feasibility of the nitrogen-doped SAC-like sites in MXenes under anodic polarization (Stage 1 in Fig. 2); (ii) assessing the thermodynamic stability of the nitrogen-doped MXene-SAC motif under OER conditions and comparison with undoped counterparts (Stage 2 in Fig. 2); (iii) analyzing OER activity under anodic polarization and selecting the most promising candidate materials (Stage 3 in Fig. 2). In each stage, increasingly stringent screening criteria are applied to ensure that the final catalysts selected have structural and chemical feasibility, thermodynamic stability, and high electrocatalytic activity under OER conditions.
Thermodynamic feasibility of nitrogen-doped MXene-SAC motifs
The 3O*-MSAC-*O structure (cf. Fig. 1a), as observed in ab initio molecular dynamics (AIMD) simulations35,36, serves as the starting point for the nitrogen substitution process. The three oxygen atoms attached to the SAC metal atom are replaced by three nitrogen atoms in an ammonia-saturated aqueous environment, with in situ formed NH3 serving as the nitrogen source37. To assess the feasibility of forming 3N*-MSAC-*O (cf. Fig. 1b), we analyze the thermochemistry of Eq. (1):
It is worth noting that there is a competition between NH3 and H2O adsorption on the SAC site, although we consider that there is no ammonia-saturated environment when OER takes place. Therefore, even if NH3 has adsorbed on the SAC site, it will be replaced by H2O under OER conditions. In addition, at pH values below 9.25, where NH3 and NH4+ are in equilibrium, NH3 mainly exists as NH4+, whose positive charge limits its propensity to adsorb on the SAC site, thus largely ruling out competitive adsorption.
Structurally, the geometry of 3N*-MSAC-*O preserves the initial planar coordination environment, and the three coordinated oxygen atoms are directly substituted by nitrogen atoms. Given that Eq. (1) is an oxidation process accompanied by proton-coupled electron transfer (PCET), we apply the computational hydrogen electrode (CHE) model38 to evaluate the applied electrode potential, at which the nitrogen doping process is in electrochemical equilibrium (Ueq). Considering that this study focuses on the OER over the MXene-SAC motif, nitrogen doping via ammonia treatment is thermodynamically feasible only when the equilibrium potential of the doping process is lower than the standard equilibrium potential of OER, which amounts to 1.23 V vs. RHE. Therefore, our first-stage screening criterion is referred to as Ueq < 1.23 V vs. RHE, which was applied to a total of 53 N-doped MXene–SAC motifs (cf. Fig. 3).
The Ueq values for the SAC-like motifs of M2X1O, M3X2O, and M4X3O (M = Cr, Fe, Hf, Mo, Nb, Ta, Ti, V, W, or Zr, X = C or N) are summarized in Tables S1–S3 of section S1 in the supporting information (SI). For M2C1O-SAC MXene (cf. Table S1), the results indicate that the feasibility of nitrogen doping is sensitive to the chemical nature of transition metal. Specifically, the equilibrium potentials of W-, Mo-, Ta-, Cr-, and Fe-based M2C1O-SAC are below the threshold of 1.23 V vs. RHE, thus indicating that the nitrogen-doped structure is available below the OER equilibrium potential. In contrast, the equilibrium potentials of Hf-, Zr-, Nb-, Ti-, and V-based M2C1O-SAC exceed 1.23 V, indicating that the doping process is thermodynamically unfavorable below the OER equilibrium potential.
The thermodynamic feasibility of nitrogen substitution in the SAC-like sites of M2N1O MXene exhibits a trend comparable to that observed for the carbides (cf. Table S1). In addition to W-, Mo-, Ta-, and Cr-based M2N1O-SAC, also V2N1O-SAC enables nitrogen substitution at applied electrode potentials below 1.23 V vs. RHE.
We extend the analysis to the SAC-like sites to M3X2O MXenes, as summarized in Table S2 of the SI. For the carbide systems, nitrogen doping is thermodynamically favorable for W-, Nb-, V-, Mo-, and Ta-based SAC sites, with Ueq below 1.23 V vs. RHE. Conversely, Hf-, Zr-, and Ti-based SAC-like motifs of M3C2O exceed the potential threshold, thus ruling out the possibility of nitrogen doping. We note that these results are consistent with the carbides of the SAC-like motifs of the M2C1O MXene. However, Hf- and Zr-based SAC-like sites in M3N2O MXene enable nitrogen substitution according to the screening criterion, in contrast to their carbide counterparts M3C2O-SAC and the general trend observed above.
In the following, the exploration is further extended to SAC-like sites in M4X3O MXenes, as shown in Table S3. Similar to the above carbides systems, W-, Nb-, Mo- and Ta-based SAC-like of M4C3O remain thermodynamically feasible at N doping below 1.23 V vs. RHE. In contrast, Hf-, Zr-, V-, and Ti-based systems only become thermodynamically favorable for N substitution at electrode potentials above 1.23 V. For the nitride MXene-SAC motifs, the trend is largely consistent with that of M4C3O, except for V-based M4N3O, which shows thermodynamic feasibility upon N doping.
Finally, approximately 60% of the initial 53 N-doped MXene–SAC motifs—a total of 32 candidates—meet the thermodynamic feasibility criterion of Ueq < 1.23 V vs. RHE and were selected for further thermodynamic stability screening (Fig. 3). Specially: (i) 10 candidates for M2X1O-SAC—W, Mo, Ta, Cr and Fe for M2C1O-SAC; W, Mo, Ta, Cr and V for SAC-like M2N1O MXenes; (ii) 12 for M3X2O-SAC—W, Mo, Ta, Nb and V for M3C2O-SAC; W, Mo, Cr, Nb, V, Hf and Zr for M3N2O-SAC; (iii) 10 for M4X3O-SAC—W, Mo, Ta and Nb for M4C3O-SAC; W, Mo, Ta, Cr, Nb and V for M4N3O-SAC. These 32 N-doped candidates were selected for further stability analysis and compared with the undoped 3O*-MSAC-*O motifs of SAC-like Mn+1XnO to assess their potential as robust single-atom centers for OER.
Thermodynamic stability of nitrogen-doped MXene-SAC motifs
The thermodynamic stability analysis in this section focuses on single-atom metal deactivation and vacancy (*vac) formation in undoped 3O*-MSAC-*O for the sake of comparison and N-doped 3N*-MSAC-*O of SAC-like MXenes under applied anodic potentials, as illustrated in Fig. 4. In total, 64 structures—including the 32 N-doped MXene–SAC motifs (3N*-MSAC-*O) identified in the first stage of screening and their 32 corresponding undoped 3O*-MSAC-*O counterparts—were investigated. The *vac structures formed after demetallization were fully optimized using DFT calculations.
Top views of the p(3 × 3) SAC-like MXene (0001) surface under anodic polarization: a undoped 3O*-MSAC-*O and b doped 3N*-MSAC-*O. In both cases, the single metal atom is displaced from the basal plane, forming a metal–oxygen (M–O) bond. Upon the dissolution process, the M–O bond is removed through demetallation, resulting in the formation of a surface vacancy (*vac).
To quantitatively assess the dissolution behavior of metal active sites in SAC-like MXenes, a thermodynamic Born–Haber cycle approach was adopted, as originally proposed by Di Liberto et al.20 to evaluate the stability of SAC structures under OER conditions using first-principles calculations. We further go beyond this approach by extending the analysis to both pre-OER and OER scenarios, as discussed by Meng et al.21 recently (Figs. S1 and S2). In both cases, the single metal atom at the SAC sites is dissolved into an ionic species (Mn+) with various oxidation states (see Table S4). Note that the pre-OER and OER scenarios exhibit a Gibbs free-energy difference of 4.92 eV ), corresponding to the overall OER Gibbs free-energy change at U = 0 V vs. RHE. Therefore, here, for simplicity, we assume that the dissolution potential Udiss (cf. Eq. 10 in “Methods” section) refers only to the pre-OER scenario. Accordingly, 3N*-MSAC-*O and 3O*-MSAC-*O systems with Udiss exceeding 1.60 V vs. RHE are considered thermodynamically stable under OER conditions (Fig. 5). The threshold criterion for catalyst stability was chosen to be 1.60 V vs. RHE as this value is slightly above typical OER conditions of 1.53 V vs. RHE, corresponding to an applied overpotential of 0.30 V. An overpotential on this order of magnitude is commonly required for active OER catalysts to achieve a current density of about 10 mA/cm2, which is relevant for solar cell applications39,40,41.
We start with the stability analysis of the 3N*-MSAC-*O systems together with their 3O*-MSAC-*O precursors of M2C1O-SAC MXenes (M = Cr, Fe, Mo, Ta, W) as a reference in the pre-OER scenario. Note that both Udiss values of the pre-OER and OER scenarios for 3N*-MSAC-*O and 3O*-MSAC-*O are summarized in Table S5, S6 and Tables S7, S8 of the SI, respectively. Compared to the 3O*-MSAC-*O systems (Table S6), nitrogen doping enhances the stability of the SAC sites of W-, Mo-, and Ta-based M2C1O-SAC motifs (Table S5). In particular, Mo- and Ta-based 3N*-MSAC-*O remains stable up to an electrode potential exceeding 4.0 V, while for 3N*-WSAC-*O a significant stabilization is witnessed compared to its highly unstable 3O*-WSAC-*O counterpart. However, nitrogen doping leads to decreased stability for Cr- and Fe-based systems, especially for Fe, which tends to dissolve and forms Fe3+ at electrode potential below 0 V vs. RHE. Despite this, nitrogen doping contributes positively to the stability of most metal sites in M2C1O-SAC MXenes.
Similar to the above analysis, we also constructed Pourbaix diagrams to evaluate the stability behavior of 3N*-MSAC-*O in the pre-OER scenario, using 3O*-MSAC-*O as a reference, as shown in Figs. 6, 7, respectively. The analysis is based on the M2C1O-SAC motifs under various pH conditions, with a focus on the potential range of 0.8–2.0 V vs. SHE to represent the oxidative environment relevant to the OER. In the pre-OER scenario, the dissolution of the metal atom involves proton consumption, and thus the boundary lines between M2C1O-SAC and Mn+ exhibits a positive slope on the SHE scale, indicating that the SAC sites are becoming thermodynamically more stable at higher pH values, as shown in Fig. S1 of the SI and Eq. 11 in the “Methods” section.
Pourbaix diagrams assessing the thermodynamic stability of SAC-like sites on the MXene basal plane, 3N*-MSAC-*O, of a W2C1O, b Mo2C1O, c Ta2C1O, d Cr2C1O, and e Fe2C1O MXenes in the context of the pre-OER scenario (cf. Fig. S1). The stability of MSAC is referenced to different oxidative ion states resulting from the dissolution of the corresponding metals (cf. Table S4 in the SI). Regions where 3N*-MSAC-*O is thermodynamically stable are colored blue, while a yellow background indicates that metal cations are energetically favored. The black, dashed line indicates the OER equilibrium potential on the SHE scale.
Pourbaix diagrams assessing the thermodynamic stability of SAC-like sites on the MXene basal plane, 3O*-MSAC-*O, of a W2C1O, b Mo2C1O, c Ta2C1O, d Cr2C1O, and e Fe2C1O MXenes in the context of the pre-OER scenario (cf. Fig. S1). The stability of MSAC is referenced to different oxidative ion states resulting from the dissolution of the corresponding metals (cf. Table S4 in the SI). Regions where 3O*-MSAC-*O is thermodynamically stable are colored blue, while a yellow background indicates that metal cations are energetically favored. The black, dashed line indicates the OER equilibrium potential on the SHE scale. Panels (a–e) are reproduced with permission from ref. 21.
Taking 3N*-CrSAC-*O in Fig. 6 as an example, its stability critically depends on the applied electrode potential and the pH value. Under acidic conditions of pH < 5, 3N*-CrSAC-*O exhibits poor thermodynamic stability within the anodic potential range of 0.8 to 2 V vs. SHE, and tends to decompose into Cr3+ species. As the pH increases, its stability improves significantly, and at pH > 11, 3N*-CrSAC-*O becomes thermodynamically stable at an electrode potential of 1.60 V vs. SHE. In contrast, 3N*-FeSAC-*O is unstable under any U and pH values, dissolving to form Fe3+ species, which is consistent with its low Udiss of −1.44 V listed in Table S5. The Pourbaix diagrams under OER conditions are shown in Figs. S3–S4, following the same analysis method as used for the pre-OER scenario in Figs. 6 and 7.
A detailed comparison of the stability of the M2N1O-SAC motifs in the pre-OER and OER scenarios between the N-doped 3N*-MSAC-*O and undoped 3O*-MSAC-*O configurations is presented in Tables S9, S10 and Tables S11, S12 of the SI, respectively. The corresponding Pourbaix diagrams are shown in Figs. S5–S8 of the SI. Similar with M2C1O-SAC, the results of M2N1O-SAC also indicate that nitrogen doping (cf. Table S9 of the SI) significantly reduces the dissolution tendency of metal sites, thereby enhancing their structural stability under pre-OER scenario—especially for Mo- and Ta-based systems with the highest dissolution potential around 4.0 V vs. RHE—thereby broadening the electrochemical window for structural stability during OER. However, Cr-based systems show a deterioration in stability upon nitrogen doping, with an even higher dissolution tendency than that in the undoped state.
We next examine the stability evolution of M3C2O-SAC MXenes, and the relevant results for the pre-OER and OER scenarios of 3N*-MSAC-*O and 3O*-MSAC-*O are summarized in Tables S13, S14 and Tables S15, S16 of the SI, respectively. The corresponding Pourbaix diagrams are shown in Figs. S9–S12 of the SI. Under pre-OER scenario, nitrogen doping significantly enhances the stability for all metal centers, i.e., W-, Mo-, Ta-, Nb-, and V-based systems (see Table S13 of the SI). Among all systems, 3N*-MoSAC-*O and 3N*-TaSAC-*O still indicate the highest resistance to dissolution, with 3N*-MoSAC-*O in particular showing a dissolution potential of approximately 4.55 V vs. RHE.
Furthermore, we explored the stability of M3N2O-SAC MXenes for 3N*-MSAC-*O and 3O*-MSAC-*O under pre-OER and OER conditions, and the results are summarized in Tables S17, S18 and Tables S19, S20 of the SI. The corresponding Pourbaix diagrams are shown in Figs. S13–S16. In the pre-OER scenario, it can be observed that nitrogen doping significantly improves the stability of metal sites in W-, Nb-, and V-based systems, which were highly unstable in the undoped state (cf. Fig. S13 of the SI). However, nitrogen doping still fails to improve the inherent instability of Hf- and Zr-based systems, which readily dissolve as HfO2+(aq) and Zr4+(aq), respectively, even under different pH conditions. Notably, 3N*-MoSAC-*O still exhibits the highest stability under the pre-OER scenario, with a dissolution potential of 4.78 V vs. RHE (see Table S17).
Finally, the data for SAC-like M4C3O (Tables S21–S24 and Figs. S17–S20) and M4N3O (Tables S25–S28 and Figs. S21–S24) are summarized accordingly in the SI. The overall conclusion aligns with the above findings, i.e., nitrogen doping can significantly enhance the stability of metal sites, except for Cr-based system (Tables S21 and S25). In addition, Mo- and Ta-based systems, which are already relatively stable before nitrogen doping (Tables S22 and S26), exhibit the greatest thermodynamic stability upon nitrogen doping, as evidenced by their significantly extended potential window (Figs. S17 and S21).
During the second stage of the thermodynamic stability screening from the initial set of 32 candidates, 11 representative MXenes-SAC motifs were selected for the subsequent activity evaluation in the OER. Specifically, the selected candidates include: (i) Mo- and Ta-based 3N*-MSAC-*O for M2X1O-SAC MXenes; (ii) Mo- and Ta-based systems for M3C2O-SAC MXenes; 3N*-MoSAC-*O for M3N2O-SAC MXenes; (iii) Mo- and Ta-based systems for M4X3O-SAC MXenes (cf. Fig. 5).
Catalytic activity of nitrogen-doped MXene-SAC motifs for OER
To evaluate the OER catalytic activity of the above candidates, we performed DFT calculations within the CHE framework to describe the proton-electron transfer steps in the mononuclear-Walden mechanism42 for OER and applied the thermodynamic descriptor Gmax(U)43,44 to assess the electrocatalytic activity (cf. “Methods” section). We use an applied electrode potential of U = 1.53 V vs. RHE in the analysis, as an applied overpotential of 300 mV typically corresponds to the experimental condition for achieving a current density of approximately 10 mA/cm39. Thus, a threshold16 of Gmax < 0.60 eV @U = 1.53 V vs. RHE is used to evaluate the 11 N-doped candidates identified above. We note that the threshold of 0.60 eV16,43 to distinguish between active and inactive catalysts is related to previous work16 on the topic.
The corresponding limiting Gibbs free-energy spans for the N-doped SAC-like MXene are shown in Fig. S25, and the values of the reaction activity descriptor Gmax(U) are summarized in Table 1. The related optimized intermediate structures, 3N*-MSAC-*O, 3N*-MSAC-*O OH, 3N*-MSAC-*OO and 3N*-MSAC-*OH, within the mononuclear-Walden framework are presented in Fig. S26. Among all 11 candidate materials, we observed that at the potential of U = 1.53 V vs. RHE, the limiting step in the approximation of the Gmax(U) descriptor consistently corresponds to the formation of the *OOH intermediate (3N*-MSAC-*O + H2O → 3N*-MSAC-*OOH + H+ + e‒), suggesting that this step might be reconciled with the rate-determining step (RDS). However, when the potential is reduced to 1.23 V vs. RHE, the limiting step for N-doped SAC-like Ta2C and Ta2N MXenes shifts to 3N*-MSAC-*OH → 3N*-MSAC-*OOH, whereas no such transition is observed in the other 9 N-doped SAC-like MXenes. This observation aligns with the findings of Razzaq et al.16, who reported that MXene-SAC motifs with high electrocatalytic activity exhibit a shift in the RDS under varying applied potentials during OER, which corresponds to a change in the limiting span of Gmax(U). Such behavior reflects a favorable reaction kinetics due to a change in the Tafel slope with increasing overpotential43,44.
Notably, N-doped Ta2C1O- and Ta2N1O-SAC exhibit higher catalytic activity compared to the other candidates—particularly Ta2N1O-SAC, which has a remarkably low Gmax(U = 1.53 V) value of 0.39 eV. Furthermore, Mo-based MXene-SACs, such as Mo2C1O, Mo3C2O, and Mo4C3O, which combine high nitrogen doping feasibility with robust structural stability, exhibit 0.60 eV < Gmax(U = 1.53 V) < 1.00 eV (Table 1), and thus are not sufficiently active for OER.
While a previous study45 suggested that a Walden-type step (cf. Eq. 15 in “Methods” section) serves as the RDS for the oxygen reduction reaction (ORR) over archetypal SAC catalysts, we conclude based on the evaluation of the Gmax(U) descriptor that the desorption-adsorption Walden step is not related to the RDS. This finding suggests that the contribution of desorption-adsorption Walden steps to the electrocatalytic activity is not universal for different catalysts, although for deeper insight the identification of transition states is necessary, which is beyond the scope of the present manuscript.
In summary, during the third stage of screening 11 nitrogen-doped MXene-SAC motifs related to their OER activity, Ta2C1O- and Ta2N1O-SAC exhibit the most promising performance, with Gmax values of 0.56 eV and 0.39 eV at 1.53 V vs. RHE, respectively. These values fall within the proposed activity threshold of Gmax (U = 1.53 V) < 0.60 eV. Combined with their thermodynamic feasibility for nitrogen doping, structural stability under both (i) pre-OER and (ii) OER scenarios, and their remarkable catalytic activity, Ta2C1O- and Ta2N1O-SAC emerge as the most promising candidates for OER (Fig. 8).
Finally, to gain further insight into how nitrogen doping affects the structural stability and activity of MXene-SAC motifs, a Bader charge analysis was performed, as illustrated in Figs. S27 to S32. Here, we focus on the average charge transfer of the nitrogen or oxygen atoms coordinated to the metal center in 3N*-MSAC-*O and 3O*-MSAC-*O, respectively. The result indicates that, in most systems, the 3N* environment exhibits significantly stronger electron accumulation compared to their O*-coordinated counterparts. This can be seen as the primary reason for the enhanced structural stability and catalytic performance of the nitrogen-doped structures. In contrast, Fe- and Cr-based systems deviate from this trend—nitrogen doping leads to reduced electron accumulation (Fig. S27), which in turn decreases the structural stability. We highlight that particularly Ta-based systems show the most prominent effect related to electron accumulation (Figs. S27–S29, S31–S32), indicating their particular ability in stabilizing MXene-SAC configurations.
Discussion
This study systematically investigated the feasibility of nitrogen doping for fully O-terminated MXene-SAC motifs and provides an in-depth analysis of its role in enhancing the stability and activity of a single metal site under anodic polarization conditions, as well its potential application in OER. Thermodynamic analysis based on DFT calculations combined with the CHE model reveals that nitrogen doping can significantly regulate the coordination environment of metal sites, effectively improving their resistance to dissolution under harsh electrochemical conditions while enhancing catalytic activity, thereby breaking the conventional activity-stability trade-off46.
Our results indicate that the thermodynamic feasibility of nitrogen substitution in MXene-SAC motifs strongly depends on the chemical nature of the transition metal. More precisely, W-, Mo-, Ta-, Cr-, Nb- and V-based systems generally exhibit a favorable nitrogen doping process across different MXene types, with equilibrium potentials (Ueq) lower than 1.23 V vs. RHE. In contrast, Ti-, Hf- and Zr-based MXene-SAC motifs are unfavorable for nitrogen incorporation regardless of their structural composites.
Nitrogen doping in the MXene-SAC motif also depends on structure and composition, but this effect is limited. Nb favors thicker MXenes (M3X2O, M4X3O), while V prefers nitrides composition over carbides. For Hf and Zr, nitrogen doping can only occur in M3N2O-SAC. In addition, nitrides MXene-SAC motifs are generally more favorable for N doping than carbides. Among all MXene structures considered in this work, M3N2O shows the widest adaptability and allows nitrogen doping even in challenging systems.
The second stage of our stability analysis reveals that nitrogen doping generally enhances the stability of metal sites by increasing their dissolution potential under anodic conditions, thereby expanding the electrochemical operating window. In particular, Mo- and Ta-based systems exhibit excellent intrinsic stability due to nitrogen incorporation. Notably, Mo-based system can achieve a dissolution potential of about 4.50 V vs. RHE, significantly extending their stability range. These observations are also consistent with previous studies47,48 highlighting the stabilizing role of nitrogen doping in catalysts. However, the stabilizing effect of nitrogen incorporation is not universal across all metal species, as Cr- and Fe-based systems exhibit decreased stability upon nitrogen doping. Similarly, the effect of nitrogen doping on the thermodynamic stability of metal sites mainly depends on the intrinsic properties of the transition metal rather than the specific configuration of MXene-SAC motifs. The impact of MXene structure and composition on metal site dissolution is minimal.
In the third stage of our screening protocol, Ta2N1O-SAC is identified as the most active catalyst among all thermodynamically stable candidates, exhibiting the lowest Gmax(U) value of 0.39 eV at U = 1.53 V vs. RHE, followed by Ta2C1O-SAC, which also shows promising OER activity with a Gmax(U) value of 0.56 eV at the same potential. The limiting span of the Gmax(U) descriptor is related to 3N*-MSAC-*OH → 3N*-MSAC-*OOH, which changes to 3N*-MSAC-*O → 3N*-MSAC-*OOH at higher overpotentials. This indicates that both Ta2C1O- and Ta2N1O-SAC show a change in the rate-determining step upon increasing overpotential, which is accompanied with a favorable reaction kinetics.
Overall, this study reveals that nitrogen doping is an effective strategy to improve both the stability and activity of MXene-SAC motifs in OER. N-doped Ta2C1O-SAC and N-doped Ta2N1O-SAC, as representative systems, break the conventional activity-stability trade-off in the OER46, and this finding is consistent with the work by Shim et al.49, who revealed that single-atom Ta doping enhances both the activity and stability of RuO2 under acidic OER conditions. It is worth noting that N-doped Ta2C1O-SAC and Ta2N1O-SAC remain unaffected as promising candidates when the OER scenario is adopted for the stability analysis (cf. section “Thermodynamic stability of nitrogen-doped MXene-SAC motifs”). This finding could motivate future experimental studies to validate the stability and catalytic performance of N-doped SAC-like MXenes. In particular, the synthesis process—especially the NH3 treatment—must be optimized to achieve precise nitrogen incorporation while avoiding over- or incomplete doping. The present work may stimulate experimental investigations in this direction by using the reported design principles to develop stable and durable OER electrocatalysts based on N-doped MXene-SAC motifs.
Methods
All calculations were performed using electronic structure theory in the density functional theory (DFT) framework within the Vienna ab initio simulation package (VASP)50,51. To accurately describe the electronic structure properties of the studied MXene systems, the generalized gradient approximation (GGA) using the Perdew-Burke-Ernzerhof (PBE) exchange-correlation functional52 was employed, which has also been widely used for 2D materials53,54. In addition, Grimme’s D3 dispersion approach55 was applied to consider intermolecular interactions, as the PBE functional is underestimating the adsorption energies unless dispersive interactions are duly accounted56. To accurately describe the influence of core electrons on the valence electron density, the projector-augmented wave (PAW) method was employed57,58. Based on this approach, the valence electron density was expanded using a plane-wave basis set with a kinetic energy cutoff of 440 eV. During optimization, the electronic self-consistent field convergence criterion was set to 10‒6 eV, and the degrees of freedom of atoms were relaxed until the force on each atom was below 0.01 eV·Å⁻1. Although hybrid functionals (e.g., HSE06) may provide more accurate descriptions of electronic structures59—particularly band gaps and charge transfer—they were not considered in this work, as they offer no significant improvement over PBE in predicting equilibrium potentials or evaluating energetics60, and their high computational cost makes them unsuitable for large-scale screening61. Moreover, a recent study has noted that gas-phase errors in hybrid functionals may propagate through free energy profiles and affect activity predictions62. Note that solvation effects were not considered in this study, as our previous work21 demonstrated that the inclusion of solvation does not significantly alter the overall trends of the investigated MXene-SAC motifs. All calculations were performed using periodic models, effectively infinite in the basal plane directions. To eliminate interactions between periodic MXene layers, a vacuum spacing of at least 19 Å was introduced along the out-of-plane direction. Integration in the Brillouin zone was carried out using a Γ-centered Monkhorst-Pack k-point mesh of 5 × 5 × 1. Furthermore, all studied SAC-like MXenes—including 3N*-MSAC-*O and 3O*-MSAC-*O types in M2X1O, M3X2O, and M4X3O MXenes (where M = Cr, Fe, Hf, Mo, Nb, Ta, Ti, V, W, or Zr, and X = C or N)—were treated with spin polarization. Gas-phase calculations of isolated atoms and molecules were performed in a symmetry-broken cell of 9 × 10 × 11 Å to ensure accurate orbital occupation and energetics.
Besides the above calculation of electronic energy (EDFT), harmonic vibrational frequencies of the studied systems were obtained by constructing and diagonalizing the relevant block of the Hessian matrix via finite difference of the analytical gradients with a step size of 0.015 Å. Based on the resulting vibrational frequencies, the zero-point energy (\({E}_{{ZPE}}\)) and entropy (TS) were further estimated (vide infra). In the frequency calculations, only the displacements of the termination adsorbates—9*O in 3O*-MSAC-*O and 6*O + 3*N in 3N*-MSAC-*O, together with the MSAC and *O atoms of MSAC-*O center, as well the first M layer of the Mn+1XnO were considered, while the remaining M and X atoms of Mn+1XnO were fixed at their optimized positions.
Furthermore, the CHE model38 proposed by Nørskov et al. was employed to simulate the effect of the applied electrode potential on the reaction free energy. Under standard equilibrium conditions—\({p}_{{\text{H}}_{2}}\,\)= 1 bar, T = 298.15 K, U = 0 V vs. SHE, and pH = 0—the electrochemical equilibrium between a proton-electron pair and gaseous hydrogen reads:
Note that the present study primarily focuses on a thermodynamic screening protocol based on the CHE model without explicitly including kinetic barriers or solvent dynamics during the reaction. Future work based on AIMD simulations could provide deeper insights into the stability of the MXene-SAC site under constant potential conditions by extending previous work on the topic15.
Accordingly, the Gibbs free energies (G) of the systems were determined as follows:
Regardless of the underlying reaction equation, the change in Gibbs free energy (ΔG) was calculated using the following Eq. (4), which is derived from the individual Gibbs free energies defined in Eq. 3, where ΔG(0, 0) represents the free-energy change at an electrode potential of 0 V vs. SHE and pH = 0.
where, ΔEDFT refers to the change in electronic energy (EDFT), ΔEZPE represents the change in zero-point energy (\({E}_{{ZPE}}\)), and ΔS is the entropy change.
To evaluate the thermodynamics of reaction processes in the framework of Eq. 4, we set up explicit reaction equations involving relevant intermediates, with Eq. 1 presented above serving as an example that illustrates the thermodynamic feasibility of the N-doping process,
Here, the EDFT obtained from the structure optimization process can be directly used to calculate the ΔEDFT term,
where, \({E}_{3{\rm{N}}* -{{\rm{M}}}_{{\rm{SAC}}}-* {\rm{O}}}\) and \({E}_{3{\rm{O}}* -{{\rm{M}}}_{{\rm{SAC}}}-* {\rm{O}}}\) represent the electronic energies of the 3N*-MSAC-*O and 3O*-MSAC-*O systems, respectively. \({E}_{{{\rm{H}}}_{2}{\rm{O}}}\), \({E}_{{{\rm{H}}}_{2}}\), and \({E}_{{{\rm{NH}}}_{3}}\) are the energies of isolated H2O, H2, and NH3 molecules in vacuum. Similarly, the change in the zero-point energy contribution (∆EZPE) is obtained as:
where, \({E}_{3{\rm{N}}* -{{\rm{M}}}_{{\rm{SAC}}}-* {\rm{O}}}^{{ZPE}}\) and \({E}_{3{\rm{O}}* -{{\rm{M}}}_{{\rm{SAC}}}-* {\rm{O}}}^{{ZPE}}\) represent the zero-point energies (ZPE) of the 3N*-MSAC-*O and 3O*-MSAC-*O systems, respectively. \({E}_{{{\rm{H}}}_{2}{\rm{O}}}^{{ZPE}}\), \({E}_{{{\rm{H}}}_{2}}^{{ZPE}}\), and \({E}_{{{\rm{NH}}}_{3}}^{{ZPE}}\) are the ZPE of H2O, H2, and NH3 molecules in vacuum. The zero-point energy term, EZPE, is calculated from the vibrational frequencies obtained through DFT calculations as follows:
where h is the Planck constants, vi is the vibrational frequency of the ith mode. For a system with N atoms, the normal vibrational modes (NVMs) are determined from its molecular geometry. For gas-phase linear molecules such as H2, the NVM are 3N‒5; for nonlinear molecules like H2O and NH3, it is 3N−6. For the studied extended systems, such as 3N*-MSAC-*O and 3O*-MSAC-*O, all 3N degrees of freedom are treated as vibrational modes, since translational and rotational modes become frustrated, turning into vibrational modes. Similar to the calculation of ∆EZPE, the entropy contribution ∆S is evaluated as follows:
where, \({S}_{3{\rm{N}}* -{{\rm{M}}}_{{\rm{SAC}}}-* {\rm{O}}}\) and \({S}_{3{\rm{O}}* -{{\rm{M}}}_{{\rm{SAC}}}-* {\rm{O}}}\) represent the entropies of the 3N*-MSAC-*O and 3O*-MSAC-*O systems, respectively, while \({S}_{{{\rm{H}}}_{2}{\rm{O}}}\), \({S}_{{{\rm{H}}}_{2}}\), and \({S}_{{{NH}}_{3}}\) denote the entropies of gas-phase H2O, H2, and NH3 molecules in vacuum. The entropy contributions of these gas-phase molecules were taken from standard thermodynamic data sources63. For the studied systems such as 3N*-MSAC-*O and 3O*-MSAC-*O, due to the frustrated translational and rotational modes, only the vibrational entropy (Svib) was considered,
where vi is the vibrational frequency of the ith mode, n denotes the total number of vibrational modes, T represent the temperature in Kelvin, and kB, h are the Boltzmann and Planck constants, respectively.
Furthermore, as mentioned in the section “Thermodynamic stability of nitrogen-doped MXene-SAC motifs”, the Born–Haber thermodynamic cycle was originally proposed by Di Liberto et al.20 to estimate the stability of SAC structures under OER electrochemical conditions; however, the simultaneous O2 evolution was not considered in the original model. Thus, to provide a more comprehensive analysis, we considered two distinct dissolution pathways in this study following our previous work21: (a) pre-OER scenario (U < 1.23 V vs. RHE), with no O2 generation during demetallization of the single atom center of MXene (cf. Fig. S1); (b) OER scenario (U > 1.23 V vs. RHE), with O2 evolution during demetallization of the single atom center of MXene (cf. Fig. S2).
In both scenarios, the single metal atom may dissolve into ionic species (Mn+) with various oxidation states (see Table S4 of the SI). The overall dissolution process, illustrated in Figs. S5 and S6 of the SI includes three steps: (i) removing metal atom from the MSAC site to the gas phase (ΔG1); (ii) cohesive energy to convert the gas-phase atom into the solid phase (ΔG2); (iii) oxidizing the solid metal atom into an ionic species (ΔG3). Among them, only ΔG1 of (i) step is calculated via DFT using the CHE model, while ΔG2 and ΔG3 are derived from data in the thermodynamic tables64,65 and standard electrode potential (ΔE°)66,67, respectively. The total Gibbs free energy change, referred to as ΔG1-3, is used to estimate the dissolution potential, Udiss:
where n is the number of electrons transferred in the overall reaction.
The free energy change ΔG(pH, U) as a function of pH and U can be derived as follows:
where ν(H⁺) and ν(e⁻) denote the stoichiometric coefficients of protons and electrons involved in the reaction equation for the elementary step. Based on this approach, pH- and U-dependent Pourbaix diagrams can be constructed by determining the equilibrium lines between various dissolved species, which are derived from the above-calculated ΔG(pH, U) to illustrate the thermodynamically stable regions of the system32.
Similarly, as described in the section “Thermodynamic feasibility of nitrogen-doped MXene-SAC motifs”, the thermodynamic feasibility of the N-doping process was evaluated based on Eq. 1:
The corresponding Gibbs free energy change (ΔG) can be obtained based on the above reaction equation and invoking the CHE model38. The ΔG value is used to estimate the equilibrium potential (Ueq) according to the following equation:
where 3 corresponds to the number of electrons transferred in Eq. 1. Values of Ueq below 1.23 V vs. RHE indicate that the N-doping process is thermodynamically favorable before OER can proceed. Conversely, when Ueq exceeds 1.23 V vs. RHE, the incorporation of nitrogen atoms is thermodynamically unfavorable below the OER equilibrium potential.
As mentioned in the section “Catalytic activity of nitrogen-doped MXene-SAC motifs for OER”, the OER reaction mechanism employed here deviates from the conventional mononuclear pathway68,69, as reported for MXenes in our previous work16. Specifically, the adsorption of the reactant H2O and the release of the product O2 proceeds in a concerted manner, while the formation of reaction intermediates does not follow the conventional sequence of a mononuclear mechanism68,69. Notably, previous studies45,70 indicate that highly active OER catalysts tend to follow such a pathway, deviating from the traditional mononuclear route. The elementary steps of the mononuclear-Walden OER mechanism are shown in Eqs. 13–16,
The values of ∆G1, ∆G2, ∆G3, and ∆G4 were determined using density functional calculations an invoking the CHE model. These corresponding ∆G values and the related intermediate structures within the mononuclear-Walden framework are presented in Figs. S25 and S26. In addition, we included gas-phase error corrections for H2O(l) and O2(g) involved in Eq. 15. Specifically, the chemical potential of gaseous H2O was corrected under the conditions of T = 298.15 K and p = 0.035 bar71, at which liquid and gaseous H2O are in thermodynamic equilibrium. To address the well-known deviation associated with the GGA functional evaluating the free energy of O2 molecules72, a correction62,73 was also applied based on the experimental enthalpy (\({\Delta}_{f}{H}^{\circ }\)) of –2.51 eV74 for the water formation reaction, as shown in Eqs. 17 and 18:
where \({\Delta}_{f}{H}^{\circ }\) is expressed as \({\Delta}_{f}{E}_{D{FT}}\) + \({\Delta}_{f}{E}_{Z{PE}}\), thus, the gaseous O2 can be corrected as Eq. 19;
After the correction of the gaseous O2 Gibbs free energy, the Gibbs free energy of the H2O(aq) → O2(g) + H2(g) reaction agrees with the experimental value when referring to the standard equilibrium potential of OER”.
In addition, to quantify the electrocatalytic activity of the N-doped MXene-SAC motifs for OER, we employed the activity descriptor Gmax(U)43,44, which refers to a free-energy span model that identifies the largest energy difference between the OER intermediate states at a given electrode potential. This energy span exhibits a Brønsted–Evans–Polanyi (BEP) relationship with the highest transition state in the catalytic cycle and can therefore be associated with the rate-determining step (RDS). Ten different free-energy spans between the relevant intermediate states of Eqs. 20–29 are possible:
Based on the free-energy spans of Eqs. 20–29, the descriptor Gmax(U) is defined as follows:
where n denotes the above ten free-energy spans, and Gmax(U) representing the largest span, that is, the span between the reaction intermediate with the lowest and highest free energy in the free energy landscape, in dependence of the applied electrode potential. Albeit the definition of the Gmax(U) descriptor is reminiscent of the free-energy span model of Kozuch and Shaik75 related to homogeneous catalysis, the notion of Gmax(U) relies on the thermodynamic information and does not capture any kinetic parameters, such as the transition state with the highest free energy, in its framework. The reason for this choice is that the calculation of transition states in an electrochemical environment is still in its infancy53,76,77, considering that most theoretical works only determine selected transition states of chemical steps, whereas the calculation of barriers for charge-transfer steps at constant electrode potential is accompanied with additional challenges78. In the framework of the original free-energy span model of Kozuch and Shaik75, the Gmax(U) descriptor estimates electrocatalytic activity based on the turnover-determining intermediate (TDI), which appears to be a valid approximation for electrocatalytic processes on surfaces, as discussed elsewhere43. Compared with the conventional approach that relies solely on the thermodynamic overpotential ηTD, Gmax(U) offers a more comprehensive assessment of the electrocatalytic activity by enabling potential-dependent activity analysis. Further details on the Gmax(U) descriptor can be found in previous publications79,80.
Data availability
The DFT data generated in this study have been deposited in the Zenodo repository database without accession code. Link: https://doi.org/10.5281/zenodo.17076457.
References
Li, S. et al. Recent advances in the development of single atom catalysts for oxygen evolution reaction. Int. J. Hydrogen Energy. 82, 1081–1100 (2024).
Zhao, C. X., Li, B. Q., Liu, J. N. & Zhang, Q. Intrinsic electrocatalytic activity regulation of M–N–C single-atom catalysts for the oxygen reduction reaction. Angew. Chem., Int. Ed. Engl. 60, 4448–4463 (2021).
Yang, M. Q. et al. Iridium single-atom catalyst coupled with lattice oxygen activated CoNiO2 for accelerating the oxygen evolution reaction. J. Mater. Chem. A 10, 25692–25700 (2022).
Zhou, K. L. et al. Platinum single-atom catalyst coupled with transition metal/metal oxide heterostructure for accelerating alkaline hydrogen evolution reaction. Nat. Commun. 12, 3783 (2021).
Qi, S., Li, C., Chen, G. & Zhao, M. Single-atom catalysts supported on graphene/electride heterostructures for the enhanced sulfur reduction reaction in lithium-sulfur batteries. J. Energy Chem. 97, 738–746 (2024).
Georgakilas, V. et al. Decorating carbon nanotubes with metal or semiconductor nanoparticles. J. Mater. Chem. 17, 2679–2694 (2007).
Chai, H. et al. Single-atom anchored g-C3N4 monolayer as efficient catalysts for nitrogen reduction reaction. Nanomater 13, 1433 (2023).
Khan, K. et al. Recent developments in emerging two-dimensional materials and their applications. J. Mater. Chem. C 8, 387–440 (2020).
Naguib, M. et al. Two-dimensional nanocrystals produced by exfoliation of Ti3AlC2. Adv. Mater. 23, 4248–4253 (2011).
Jussambayev, M. et al. MXenes for sustainable energy: a comprehensive review on conservation and storage applications. Carbon Trends 19, 100471 (2025).
Singh, D., Razzaq, S., Tayyebi, E. & Exner, K. S. Selectivity control in the nitrogen reduction reaction over Mo2C MXene by a nitrogen-rich environment. ACS Catal 15, 5589–5600 (2025).
Singh, D., Bulatova, K. & Exner, K. S. Strain engineering of Mo2C MXene to steer the selectivity in the nitrogen reduction reaction toward ammonia. Chem. Eng. J. 518, 164451 (2025).
Cao, S. et al. MXene-based single atom catalysts for efficient CO2RR towards CO: a novel strategy for high-throughput catalyst design and screening. Chem. Eng. J. 461, 141936 (2023).
Yan, Y. et al. MXene-supported single-atom and nano catalysts for effective gas-phase hydrogenation reactions. Mater. Today Chem. 2, 100010 (2023).
Wan, P., Chen, Y. & Tang, Q. Electrochemical stability of MXenes in water based on constant potential AIMD simulations. ChemPhysChem 25, e202400325 (2024).
Razzaq, S. et al. MXenes spontaneously form active and selective single-atom centers under anodic polarization conditions. J. Am. Chem. Soc. 147, 161–168 (2025).
Parker, T. et al. In situ raman and fourier transform infrared spectroscopy studies of MXene−electrolyte interfaces. ACS Nano https://doi.org/10.1021/acsnano.5c03810 (2025).
Singh, D., Razzaq, S., Faridi, S. & Exner, K. S. Selective nitrogen reduction reaction on single-atom centers of molybdenum-based MXenes by pulsing the electrochemical potential. Mater. Today 86, 575–579 (2025).
Faridi, S. et al. Trends in competing oxygen and chlorine evolution reactions over electrochemically formed single-atom centers of MXenes. J. Mater. Chem. A 13, 16481–16490 (2025).
Di Liberto, G., Giordano, L. & Pacchioni, G. Predicting the stability of single-atom catalysts in electrochemical reactions. ACS Catal 14, 45–55 (2024).
Meng, L., Viñes, F., Illas, F. & Exner, K. S. Stability of single-atom centers of MXenes under anodic polarization conditions. J. Phys. Chem. C 129, 9589–9601 (2025).
Zhang, Y. et al. The effect of coordination environment on the activity and selectivity of single-atom catalysts. Coord. Chem. Rev. 461, 214493 (2022).
Wang, L., Wang, H. & Lu, J. Local chemical environment effect in single-atom catalysis. Chem. Catal. 3, 100492 (2023).
Meng, L., Viñes, F. & Illas, F. Unveiling the synergy between surface terminations and boron configuration in boron-based Ti3C2 MXenes electrocatalysts for nitrogen reduction reaction. ACS Catal 14, 15429–15443 (2024).
Shi, Z., Yang, W., Gu, Y., Liao, T. & Sun, Z. Metal-nitrogen-doped carbon materials as highly efficient catalysts: progress and rational design. Adv. Sci. 7, 2001069 (2020).
Ha, M. et al. Tuning metal single atoms embedded in NxCy moieties toward high-performance electrocatalysis. Energy Environ. Sci. 14, 3455–3468 (2021).
Liu, X., Jiang, Z. & Guo, Y. Mechanistic understanding on effect of doping nitrogen with graphene supported single-atom Ir toward HER and OER: A computational consideration. Chem. Phys. Lett. 834, 140971 (2024).
Binninger, T., Moss, G. C., Rajan, Z. S. H. S., Mohamed, R. & Eikerling, M. H. How to break the activity-stability conundrum in oxygen evolution electrocatalysis: mechanistic insights. ChemCatChem 16, e202400567 (2024).
Zheng, X., Jiang, L., Pan, H. & Sun, W. Breaking the activity-durability trade-off of anode catalysts for proton exchange membrane water electrolyzers. ACS Energy Lett. 10, 3471–3484 (2025).
Gogotsi, Y. & Anasori, B. The rise of MXenes. ACS Nano 13, 8491–8494 (2019).
Wang, Y. et al. Controlled etching to immobilize highly dispersed Fe in MXene for electrochemical ammonia production. Carbon Neutraliz 1, 117–125 (2022).
López, M., Exner, K. S., Viñes, F. & Illas, F. Computational pourbaix diagrams for MXenes: a key ingredient toward proper theoretical electrocatalytic studies. Adv. Theory Simul. 6, 2200217 (2023).
Meng, L., Yan, L. K., Viñes, F. & Illas, F. Effect of terminations on the hydrogen evolution reaction mechanism on Ti3C2 MXene. J. Mater. Chem. A 11, 6886–6900 (2023).
Peng, J., Chen, X., Ong, W. J., Zhao, X. & Li, N. Surface and heterointerface engineering of 2D MXenes and their nanocomposites: insights into electro- and photocatalysis. Chem 5, 18–50 (2019).
Tian, Y. et al. Theoretical insights on potential-dependent oxidation behaviors and antioxidant strategies of MXenes. Nat. Commun. 15, 10099 (2024).
Hou, P. et al. Unraveling the oxidation behaviors of MXenes in aqueous systems by active-learning-potential molecular-dynamics simulation. Angew. Chem. Int. Ed. 62, e202304205 (2023).
Ustavytska, O., Kurys, Y., Koshechko, V. & Pokhodenko, V. One-step electrochemical preparation of multilayer graphene functionalized with nitrogen. Nanoscale Res. Lett. 12, 175 (2017).
Nørskov, J. K. et al. Origin of the overpotential for oxygen reduction at a fuel-cell cathode. Phys. Chem. B. 108, 17886–17892 (2004).
Yang, T. et al. High-throughput identification of exfoliable two-dimensional materials with active basal planes for hydrogen evolution. ACS Energy Lett. 5, 2313–2321 (2020).
McCrory, C. C. L., Jung, S., Peters, J. C. & Jaramillo, T. F. Benchmarking heterogeneous electrocatalysts for the oxygen evolution reaction. J. Am. Chem. Soc. 135, 16977–16987 (2013).
McCrory, C. C. L. et al. Benchmarking hydrogen evolving reaction and oxygen evolving reaction electrocatalysts for solar water splitting devices. J. Am. Chem. Soc. 137, 4347–4357 (2015).
Exner, K. S. Importance of the walden inversion for the activity volcano plot of oxygen evolution. Adv. Sci 10, 2305505 (2023).
Razzaq, S. & Exner, K. S. Materials screening by the descriptor Gmax(η): the free-energy span model in electrocatalysis. ACS Catal 13, 1740–1758 (2023).
Exner, K. S. A universal descriptor for the screening of electrode materials for multiple-electron processes: beyond the thermodynamic overpotential. ACS Catal 10, 12607–12617 (2020).
Yu, S., Levell, Z., Jiang, Z., Zhao, X. & Liu, Y. What is the rate-limiting step of oxygen reduction reaction on Fe–N–C catalysts? J. Am. Chem. Soc. 145, 25352–25356 (2023).
Tang, Y. et al. Breaking the activity-stability trade-off of Au catalysts by depth-controlled TiO2 nanotraps. J. Catal. 423, 145–153 (2023).
Peera, S. G. et al. Effect of N-doped carbon on the morphology and oxygen reduction reaction (ORR) activity of a xerogel-derived Mn(II)O Electrocatalyst. Catalysts 14, 792 (2024).
Li, Y. et al. Carbon corrosion mechanism on nitrogen-doped carbon support—a density functional theory study. Int. J. Hydrogen Energy. 46, 13273–13282 (2021).
Shim, J. et al. Atomically dispersed high-valent d0-metal breaks the activity–stability trade-off in proton exchange membrane water electrolysis. J. Am. Chem. Soc. 147, 16179–16188 (2025).
Kresse, G. & Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 54, 11169–11186 (1996).
Kresse, G. & Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 6, 15–50 (1996).
Perdew, J. P., Burke, K. & Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865–3868 (1996).
López, M., Exner, K. S., Viñes, F. & Illas, F. Theoretical study of the mechanism of the hydrogen evolution reaction on the V2C MXene: Thermodynamic and kinetic aspects. J. Catal. 421, 252–263 (2023).
Huang, K. et al. Mechanistic insights for the electrocatalytic CO2 reduction on M2C-type MXenes. Ind. Eng. Chem. Res. 62, 20716–20726 (2023).
Grimme, S., Antony, J., Ehrlich, S. & Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 132, 154104 (2010).
Ventura-Macias, E., Martinez, P. M., Pérez, R. & Vilhena, J. G. Quantum mechanical derived (VdW-DFT) transferable Lennard–Jones and morse potentials to model cysteine and alkanethiol adsorption on Au(111). Adv. Mater. Interfaces 11, 2400369 (2024).
Blöchl, P. E. Projector augmented-wave method. Phys. Rev. B 50, 17953–17979 (1994).
Kresse, G. & Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 59, 1758–1775 (1999).
Islam, M. A comprehensive investigation on the physical properties of SiC polymorphs for high-temperature applications: a DFT study based on GGA and hybrid HSE06 exchange correlation functionals. Nucl. Mater. Energy. 38, 101631 (2024).
Ontiveros, D., Vela, S., Viñes, F. & Sousa, C. Tuning MXenes towards their use in photocatalytic water splitting. Energy Environ. Mater. 7, e12774 (2024).
Deák, P., Aradi, B., Frauenheim, T., Janzén, E. & Gali, A. Accurate defect levels obtained from the HSE06 range-separated hybrid functional. Phys. Rev. B 81, 153203 (2010).
Urrego-Ortiz, R., Builes, S., Illas, F. & Calle-Vallejo, F. Gas-phase errors in computational electrocatalysis: a review. EES Catal 2, 157–179 (2024).
NIST Standard Reference Database. SRD Number 69, https://doi.org/10.18434/T4D303 (2025).
Kaxiras, E. Atomic and Electronic Structure of Solids (Cambridge University Press, 2003).
Kittel, C. Introduction to Solid State Physics, 8th edn (John Wiley & Sons, Inc, 2005).
IUPAC. Compendium of Chemical Terminology, 2nd edn (the “Gold Book”). Online corrected version, 2006, “Standard electrode potential, E. https://doi.org/10.1351/goldbook.S05912 (1997).
IUPAC. Compendium of Chemical Terminology, 3rd edn. International Union of Pure and Applied Chemistry. Online version 3.0.1, 2019, “electrode potential”. https://doi.org/10.1351/goldbook.E01956 (2006).
Man, I. C. et al. Universality in oxygen evolution electrocatalysis on oxide surfaces. ChemCatChem 3, 1159–1165 (2011).
Rossmeisl, J., Logadottir, A. & Nørskov, J. K. Electrolysis of water on (oxidized) metal surfaces. Chem. Phys. 319, 178–184 (2005).
Usama, M., Razzaq, S., Hättig, C., Steinmann, S. N. & Exner, K. S. Oxygen evolution reaction on IrO2(110) is governed by Walden-type mechanisms. Nat. Commun. 16, 6137 (2025).
Yu, M., Li, A., Kan, E. & Zhan, C. Substantial impact of spin state evolution in OER/ORR catalyzed by Fe–N–C. ACS Catal 14, 6816–6826 (2024).
Grindy, S., Meredig, B., Kirklin, S., Saal, J. E. & Wolverton, C. Approaching chemical accuracy with density functional calculations: diatomic energy corrections. Phys. Rev. B 87, 075150 (2013).
Urrego-Ortiz, R., Builes, S. & Calle-Vallejo, F. Fast correction of errors in the DFT-calculated energies of gaseous nitrogen-containing species. ChemCatChem 13, 2508–2516 (2021).
Linstrom, P. J. & Mallard, W. G. (eds). NIST Chemistry WebBook, NIST Standard Reference Database 69 (National Institute of Standards and Technology, 1997).
Kozuch, S. & Shaik, S. How to conceptualize catalytic cycles? The energetic span model. Acc. Chem. Res. 44, 101–110 (2011).
Li, N. et al. Understanding of electrochemical mechanisms for CO2 capture and conversion into hydrocarbon fuels in transition-metal carbides (MXenes). ACS Nano 11, 10825–10833 (2017).
Guo, Z. et al. M2C-type MXenes: promising catalysts for CO2 capture and reduction. Appl. Surf. Sci. 521, 146436 (2020).
Melander, M. M., Wu, T., Weckman, T. & Honkala, K. Constant inner potential DFT for modelling electrochemical systems under constant potential and bias. NPJ Comput. Mater. 10, 5 (2024).
Razzaq, S. & Exner, K. S. Method to determine the bifunctional index for the oxygen electrocatalysis from theory. ChemElectroChem 9, e202101603 (2022).
Exner, K. S. Beyond the thermodynamic volcano picture in the nitrogen reduction reaction over transition-metal oxides: implications for materials screening. Chin. J. Catal. 43, 2871–2880 (2022).
Acknowledgements
K.S.E. acknowledges funding by the RESOLV Cluster of Excellence, funded by the Deutsche Forschungsgemeinschaft under Germany’s Excellence Strategy, EXC 2033-390677874, RESOLV. S.R., D.S., and K.S.E. acknowledge funding by the Ministry of Culture and Science of the Federal State of North Rhine-Westphalia (NRW Return Grant). F.V. and F.I. acknowledge the Spanish Ministerio de Ciencia e Innovación and Agencia Estatal de Investigación (AEI) MCIN/AEI/10.13039/501100011033 and, as appropriate, by “European Union Next Generation EU/PRTR”, through grants PID2021-126076NB-I00, la Unidad de Excelencia María de Maeztu CEX2021-001202-M and Generalitat de Catalunya for 2021SGR00079 grant. F.V. is thankful for the ICREA Academia Award 2023 with ref. Ac2216561. L.M. thanks the China Scholarship Council (CSC) for financing her PhD (CSC202108390032).
Funding
Open Access funding enabled and organized by Projekt DEAL.
Author information
Authors and Affiliations
Contributions
L.M. performed the theoretical calculations, data analysis, and figure preparation for the section of the feasibility and stability of MXene-SAC motifs. L.M. also wrote the first draft of this manuscript. S.R. carried out the theoretical calculations, data analysis, and figure preparation for the OER mechanism section. D.S. contributed to this study by optimizing the MXene-SAC models. F.V. and F.I. supervised the study, provided computational resources, and contributed to manuscript revision. K.S.E. initiated the project, provided computational resources, supervised the study, co-wrote the paper with L.M., and contributed to manuscript revision. All authors contributed to the analysis and discussion of the results.
Corresponding authors
Ethics declarations
Competing interests
The authors declare no competing interests.
Additional information
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Supplementary information
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.
About this article
Cite this article
Meng, L., Razzaq, S., Singh, D. et al. Nitrogen embedding enhances stability and activity of single-atom motifs of MXenes under anodic polarization. npj 2D Mater Appl 9, 91 (2025). https://doi.org/10.1038/s41699-025-00610-z
Received:
Accepted:
Published:
Version of record:
DOI: https://doi.org/10.1038/s41699-025-00610-z