Next Article in Journal
The Property–Efficiency Relationship over Rh/GaxNby Catalysts in Photothermal Dry Reforming of CH4
Previous Article in Journal
Agricultural Plastic Mulch: A Brief Review of Development, Composition and Catalytic Upcycling Strategies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

In-Depth First-Principles Study of High-Performance M2XO2 MXene Cathode Catalysts for Sodium-Oxygen Batteries

Shandong Key Laboratory of Intelligent Energy Materials, School of Materials Science and Engineering, China University of Petroleum (East China), Qingdao 266580, China
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(4), 311; https://doi.org/10.3390/catal15040311
Submission received: 4 March 2025 / Revised: 21 March 2025 / Accepted: 23 March 2025 / Published: 25 March 2025
(This article belongs to the Section Computational Catalysis)

Abstract

:
Na−O2 batteries are plagued by high cathodic oxygen reduction (ORR)/oxygen evolution (OER) overpotentials during discharging/charging. Herein, we constructed six carbide/nitride MXenes (M2XO2, M = Ti, Zr, and Hf, X = C, and N) and investigated their performance as cathodes for Na−O2 batteries by first-principles calculations. M2CO2 MXenes have a pseudogap, showing semiconducting properties, while M2NO2 MXenes are conductive. The nucleophilic O on the M2XO2 surfaces prefers to bind with the Na atoms of NaxO2 intermediates to activate the Na−O bonds, improving the sodium deintercalation. For all M2XO2 MXenes, the OER overpotential is higher than the ORR overpotential, forming a performance bottleneck of Na−O2 batteries. The overpotentials originate from the too-strong adsorption of NaxO2 on M2XO2 MXenes. Lowering the O p-band center of the M2XO2 MXenes can weaken the NaxO2 adsorption, thereby reducing the overpotential. Consequently, the overpotentials of the M2CO2 carbides are lower than those of the M2NO2 nitrides and further decrease with a decreasing M atomic number. The Ti2CO2 MXene shows extremely low ORR, OER, and total overpotentials (0.23, 0.32, and 0.55 V), suggesting a huge potential as cathodes in Na−O2 batteries.

Graphical Abstract

1. Introduction

The depletion of fossil fuel resources is exacerbated by increasing energy consumption due to population growth and economic expansion. In the quest for more efficient and eco-friendly energy storage solutions, extensive research has been conducted on rechargeable batteries [1,2,3]. Different from conventional battery systems that are self-containing, in the aprotic alkali metal−O2 batteries (such as Li−O2 batteries and Na−O2 batteries), oxygen from the atmosphere participates in the cathode reaction through a gas electrode, which produces a high energy density, resulting in the high energy density battery systems. The electrochemical power devices consist of a Li/Na anode, where Li or Na−metal is oxidized upon discharge to Li+/Na+ ions, and an oxygen cathode, where Li+/Na+ ions recombine with electrons to reduce O2 gas from the ambient environment. The theoretical energy density of alkali metal−O2 batteries (3600 Wh kg−1 for Li−O2 batteries and 1100 Wh kg−1 for Na−O2 batteries) significantly exceeds that of Li−ion batteries (372 Wh kg−1) [4,5]. Compared to Li−O2 batteries, Na−O2 batteries exhibit the following advantages [4,5]: (1) Resource abundance: Sodium exhibits significantly higher crustal abundance (~2.3%) compared to lithium (0.0017%), with correspondingly lower production costs. (2) Efficient discharge product: NaO2 (overpotential ~75 mV) decomposes more readily than Li2O2 (~1.5 V overpotential), enabling >97% energy efficiency. (3) Suppressed dendrites: The lower Young’s modulus of sodium (~1 GPa, one-fifth of lithium’s 4.9 GPa) effectively suppresses metallic dendrite growth, extending the cycle life. (4) Electrolyte stability: Operating at a higher redox potential (−2.71 V vs. −3.04 V for Li), sodium substantially reduces electrolyte decomposition. (5) Temperature adaptability: Na−O2 batteries maintain stable electrochemical performance across a broad temperature range (−20 °C to 80 °C), whereas Li−O2 systems suffer from severe capacity decay below −10 °C. Therefore, the Na−O2 battery is a promising alternative to the Li−O2 battery. During discharge and charge, the oxygen reduction reaction (ORR) and the oxygen evolution reaction (OER) occur at the cathode of Na−O2 batteries, respectively. In the ORR process, atmospheric O2 on the cathode combines with the Na+ + e pairs to form NaxO2 products (x = 1, 2, or 4), which then break down into the original Na+ and oxygen molecules in the OER process [6,7]. However, the sluggish ORR/OER dynamics of Na−O2 batteries strongly hinder their practical applications [6,7]. Therefore, it is urgent to find suitable cathode electrocatalysts for efficient Na−O2 batteries.
Currently, noble metals such as Au, Pt, Ru, and Ir are the most effective catalysts for the cathode of Na−O2 batteries [8,9]. However, these metals are expensive and can only perform one of the two required activities: ORR or OER [10,11,12]. Carbon materials are abundant, inexpensive, and easily shaped, but their low catalytic activity limits their use in the cathode matrix [13,14]. Although transition metal oxides have been reported to have good electrocatalytic activity, their poor conductivity forces them to be organically mixed with conductive substances, making them difficult to produce [15,16]. In recent years, MXenes, a new large family of two-dimensional (2D) materials, have attracted considerable research interest due to their unique electronic and structural properties. MXenes have been applied in capacitors [17,18], metal–ion batteries [19,20], metal−N2 batteries [21], and metal−O2 batteries [22,23]. Especially, MXenes are considered promising candidates for addressing critical challenges in Na–O2 batteries, with key advantages including the following [23,24,25,26,27,28]: (1) MXenes (e.g., Ti2CO2) combine metallic conductivity (an efficient charge transfer) and tunable surface chemistry (e.g., –O terminals) to optimize intermediate adsorption/desorption (e.g., NaO2), lowering the ORR/OER activation barriers. (2) Surface oxygen groups stabilize Na⁺ ions and nucleophilic oxygen species, promoting NaO2 formation/decomposition. (3) Polar surfaces can anchor NaO2 nanoparticles, suppressing their dissolution/aggregation. (4) They are corrosion-resistant under high oxidative potentials (e.g., OER), ensuring cycling stability. (5) Hydrophilic surfaces enable rapid Na⁺ diffusion, reducing localized current density and dendrite nucleation. (6) Enhanced electrolyte wettability ensures a uniform Na⁺ distribution and lowers interfacial resistance. (7) Structural stability from −50 °C to 300 °C supports reliable operation across wide thermal ranges. Min et al. reported that MXene/Cu-BHT heterostructures have ultralow OER (0.24 V) and ORR (0.32 V) overpotentials for Na−O2 batteries [24]. Tang et al. constructed the TMD/MXene heterostructures with 0.55 V/0.20 V for the ORR/OER overpotentials, which show high catalytic activity in Na–O2 batteries [25]. He et al. reported that the h-Ti3C2/CNT composite exhibits a low potential gap of 0.11 V after 70 cycles for Na–O2 batteries [26]. Huang et al. designed a Co–S–N–C@MXene catalyst, which shows excellent ORR activity with a half-wave potential of 0.881 V [27]. The half-wave potential exhibited a minimal decline of only 14 mV after 10,000 cycles of cyclic voltammetry testing. Tan et al. found that the NiFe-P/Ti3C2Tx composite exhibits exceptional activity and stability for OER, achieving a low overpotential of 290 mV and a Tafel slope of 72.3 mV/dec at a current density of 10 mA cm−2 [28].
The M2XO2 monolayers (M = Ti, Zr, and Hf; X = C and N), belonging to the Group 4 Ti-type MXenes, exhibit excellent stability, good machinability, high conductivity, and metallic/semimetallic properties, making them highly suitable as electrocatalysts [29]. Ti2CO2 MXenes were found to have superior storage capacity and a low diffusion barrier for Na ions [30]. Furthermore, recent studies have identified Ti2CO2 MXenes as efficient ORR catalysts [31], while Ti2NO2 MXenes show OER performance [32]. These findings suggest a significant potential for applying Group 4 Ti-type MXenes in Na–O2 batteries. However, the performance of catalysts is closely related to the types of constituent elements. It is interesting to explore the influence of Ti, Zr, and Hf on the catalytic performance of Group 4 Ti-type MXenes towards ORR and OER. In this work, we constructed six M2XO2 MXenes (M = Ti, Zr, and Hf; X = C and N) and explored their ORR/OER performance during discharge/charging for Na−O2 batteries by first-principles density functional theory (DFT) calculations. The purpose of this work is to give our theoretical suggestion about MXenes with different components as cathodes for Na−O2 batteries to guide the experiments. This work merely considers the intrinsic properties of the electrocatalysts, and thus, some external factors are not considered, including the morphology of discharge products, the stability of electrolytes, the corrosion of sodium metal, and the polarization of oxygen electrodes. First, the geometrical and electronic structure of M2XO2 MXenes was carefully studied. All of the M2CO2 MXenes exhibit semiconducting properties, while the M2NO2 MXenes show conductivity. Then, we studied the adsorption properties of the NaxO2 intermediates on M2XO2 MXenes. Finally, the formation/dissociation of NaxO2 on six MXenes in the ORR/OER processes was simulated, and the overpotentials were calculated to assess their catalytic activity quantitatively. The Ti2CO2 MXene was found to exhibit excellent catalytic activity with ultralow ORR, OER, and total overpotentials (0.23, 0.32, and 0.55 V). This work is of great significance in providing available theoretical references of MXenes as cathode catalysts for Na−O2 batteries.

2. Results and Discussion

2.1. Structure Properties

The schematic models of M2X and M2XO2 MXenes (M = Ti, Zr, and Hf, X = C, and N) are shown in Figure 1, and the corresponding fully optimized configurations are shown in Figures S1 and S2. The lattice constants of Ti2C, Ti2N, Zr2C, Zr2N, Hf2C, and Hf2N are calculated as 3.042, 2.994, 3.274, 3.239, 3.115, and 3.198 Å, respectively, which are consistent with other published findings (3.039, 2.985, 3.280, 3.246, 3.211, and 3.171 Å) [33,34]. As shown in Figure 1, three possible O adsorption sites (i.e., the top, hcp, and fcc sites) were considered. The formation energy (see Table S1) indicates that all adsorption sites can form stable O-terminated MXenes, where the most stable adsorption site is the fcc site. Therefore, in the following part, M2X-MXenes with full O-coverage of the fcc sites (M2XO2: Ti2CO2, Zr2CO2, Hf2CO2, Ti2NO2, Zr2NO2, and Hf2NO2) were chosen to study further.
To explore the electronic properties of M2XO2 MXenes, the projected density of the states (PDOS) was plotted (see Figure 2). For all M2XO2 MXenes, the valence band is mainly composed of the p-states of the O and C/N atoms, while the conduction band is dominated by the d-states of the transition metal atoms. In addition, the M2CO2 carbides (Ti2CO2, Zr2CO2, and Hf2CO2) exhibit semiconducting properties, as reflected by the Fermi level passing through the pseudogap (Figure 2a–c). The M2NO2 nitrides (Ti2NO2, Zr2NO2, and Hf2NO2) are conductive due to the Fermi level crossing the conduction band (Figure 2d–f). With an increasing metal atomic number, the pseudogap of M2XO2 MXenes gradually widens, indicating that the system stability is gradually improved by increasing covalent interaction. The p-band center (ɛp(O)) of the surface O atoms is calculated to be −2.57, −2.52, −2.50, −2.31, −2.22, and −2.16 eV with respect to the Fermi level for Ti2CO2, Zr2CO2, Hf2CO2, Ti2NO2, Zr2NO2, and Hf2NO2, respectively, suggesting that the ɛp(O) of the M2CO2 carbides is apparently lower than that of the M2NO2 nitrides. This indicates that the M2CO2 carbides exhibit weaker adsorption of the NaxO2 intermediates than the M2NO2 nitrides. Moreover, with the decreasing metal atomic number, the ɛp(O) gradually shifts downwards, suggesting that the adsorption ability of the M2XO2 surfaces towards NaxO2 intermediates becomes weaker progressively.
The Hirshfeld charge population analyses of M2XO2 MXenes (see Figure 3) show that surface O groups are negatively charged by −0.23 to −0.28 e, while the metal atoms are positively charged by 0.38 to 0.47 e, suggesting strong electron transfer from the metal centers to the surface O groups. As the atomic number of the metal increases, the positive charge localized on the metal center gradually increases. This trend arises from the lower electronegativity of metals with higher atomic numbers, resulting in the reduced binding strength of valence electrons to the atomic nucleus. Consequently, this enhances the charge transfer from the metal center to oxygen atoms. Furthermore, the surface oxygen groups of M2CO2 carbides carry more negative charges than those of M2NO2 nitrides, indicating stronger nucleophilicity of the M2CO2 surfaces. This situation can be further confirmed by the sliced deformation electron density (SDED) maps (see Figure 4). As shown in Figure 4, the surface O groups exhibit apparent electron aggregation (the red color region) for all M2XO2 MXenes. Especially, the red region of the O atoms on the M2CO2 carbides is larger than that on the M2NO2 nitrides, indicating its stronger electron accumulation.
As illustrated in Figure 2, M2XO2 MXenes exhibit enhanced electrical conductivity due to the hybridization of oxygen p-orbitals with transition metal d-orbitals. This interaction increases the density of electronic states near the Fermi level. Additionally, the strong electronegativity of surface oxygen leads to partial oxidation of the Ti atoms, as shown in Figure 3. This partial oxidation creates conductive pathways enriched with delocalized electrons, significantly reducing resistivity by improving charge carrier mobility. Moreover, the –O terminations can electrostatically attract Na⁺ ions and electrons, facilitating interfacial charge transfer and thus minimizing polarization losses at the electrode–electrolyte interface. Overall, these mechanisms demonstrate that surface oxygen terminations play a crucial role in enhancing the electronic conductivity of M2XO2 MXenes.

2.2. NaxO2 Adsorption

During the charging and discharging processes, the intermediate species involved on the cathode surface of Na−O2 batteries mainly include NaO2, Na2O2, and Na4O2 [2,35,36]. Herein, the adsorption of these species (NaxO2, x = 1, 2, and 4) was characterized. The calculated adsorption configurations are shown in Figure 5 and Figures S3–S5, and the corresponding structural parameters are tabulated in Tables S2–S4.
As shown in Figure 5 and Figures S3–S5, the surface nucleophilic O groups of M2XO2 MXenes prefer to combine with the electropositive Na atoms of NaxO2 (x = 1, 2, and 4) to activate the Na–O bonds. Therefore, compared to free NaxO2, the Na–O bonds in NaxO2* are elongated efficiently, while the O–O bonds in NaxO2* are shortened (see Tables S2–S4), which is conducive to the deintercalation of sodium in the charging process. The charge population analyses in Tables S5–S7 indicate that a strong electron transfer from Na to surface O groups occurs on the NaxO2*-adsorbed M2XO2 MXenes, which weakens the ionic interaction in the Na–O bonds of NaxO2*. Therefore, surface oxygen terminals can effectively enhance the OER process during charging. After the NaxO2* (x = 1, 2, and 4) adsorption on M2XO2 MXenes, the changes in the O–O and Na–O bonds in (Na2O)2* are the smallest (Tables S2–S4), suggesting that (Na2O)2* is the most difficult to activate among the NaxO2* intermediates, which is in good accordance with the experimental reports [25,37]. Furthermore, for all considered M2XO2 MXenes, M2CO2 carbides with a lower metal atomic number have a stronger activation effect on the NaxO2* intermediates, as reflected by the greater change in the Na–O and O–O bond lengths of NaxO2* (see Tables S2–S4).
With the increasing Na content, the adsorption energy of NaxO2* on the M2XO2 surfaces increases gradually (see Figure 6), indicating that the larger the NaxO2 particles, the more difficult they are to remove from the catalyst surfaces. Compared to the M2NO2 nitrides, the M2CO2 carbides show weaker adsorption energy towards NaxO2*. In addition, with the decreasing atomic number of M, the adsorption energy of NaxO2* on M2XO2 MXenes is further weakened, improving the accumulation of NaxO2 particles. Therefore, the Ti2CO2 MXene with the lowest NaxO2* adsorption energy can best promote the electrode reaction, thus avoiding the stacking of insulating and insoluble discharge products. The chemisorption of catalysts is known to be determined by the underlying electronic structure. In this work, the close correlation between the NaxO2* adsorption energy and the p-band center (εp(O)) of the surface O atoms of M2XO2 MXenes is confirmed by a linear relationship of Eads(NaO2) = −2.90εp(O) − 9.27, Eads(Na2O2) = −4.98εp(O) − 16.92, and Eads(Na4O2) = −8.86εp(O) − 30.90 (see Figure 7a–c), where the adsorption energy of NaxO2* shows a weakened trend with the decrease of εp(O).

2.3. Evaluation of Catalytic Activity

According to the previous studies [2,35,36,38], the ORR and OER occur at the cathodes of Na−O2 batteries during the discharging and charging processes, respectively, as in the Li−O2 electrochemical system. The 4e oxygen transfer pathways are involved in the cycling process: (1) O2 + Na+ + e  NaO2*; (2) NaO2* + Na+ + e  Na2O2*; (3) Na2O2* + 2Na+ + 2e  (Na2O)2*, where * denotes the active sites of the reaction intermediates [2,35,36]. During the discharge process, the ORR on the cathodic M2XO2 MXenes proceeds via an initial combination of O2 with (Na+ + e) to form an adsorbed NaO2*. Secondly, the adsorbed NaO2* reacts with (Na+ + e) to generate Na2O2*. After that, Na2O2* further reacts with (Na+ + e) to form the discharge product (Na2O)2*. During charging, the OER occurs by gradually removing sodium from (Na2O)2* to produce oxygen.
Figure 8 shows the ORR/OER free energy profiles for the M2CO2 carbides (Ti2CO2, Zr2CO2, and Hf2CO2) and M2NO2 nitrides (Ti2NO2, Zr2NO2, and Hf2NO2). In Figure 8a–f, the green arrowheads represent the ORR process from left to right when discharging, while the purple arrowheads indicate the OER process from right to left when recharging. The ORR is the process of (Na2O)2 nucleation on M2XO2 MXenes. At the open-circuit potential U = 0 V, all three steps have a downward free energy trajectory on both the M2CO2 and M2NO2 surfaces (see the black paths in Figure 8a–f), indicating the spontaneous nucleation (i.e., ORR) and endothermic decomposition (i.e., OER) of (Na2O)2. As shown in Figure 8a–f, due to the weakened adsorption of the NaxO2 (x = 1, 2, and 4) intermediates, the total free energy changes for O2 → (Na2O)2* (ΔG(O2 → (Na2O)2*)) on M2XO2 MXenes decrease with the decreasing metal atomic number, leading to the easier decomposition of (Na2O)2* in the OER process. Furthermore, the ΔG(O2 → (Na2O)2*) for the M2CO2 carbides is less than that for the M2NO2 nitrides, suggesting that the carbides can enhance the (Na2O)2 decomposition in the OER process.
At an equilibrium potential (Ue), the ORR and OER processes reach an electrochemical equilibrium state (see the orange paths in Figure 8a–f). In this work, Ue is calculated to be 1.50, 2.53, 2.81, 2.89, 2.76, and 3.46 V for Ti2CO2, Zr2CO2, Hf2CO2, Ti2NO2, Zr2NO2, and Hf2NO2, respectively. When Ue is applied, the formation of NaO2* and Na2O2* along the ORR pathway is still downhill for Ti2XO2 and Zr2XO2, but the step of (Na2O)2* formation is changed to be uphill and endothermic (see Figure 8a,b,d,e). For Hf2XO2, both Na2O2* and (Na2O)2* formation steps are uphill on the ORR free energy profiles, but the last (Na2O)2* formation step must overcome a larger energy platform (see Figure 8c,f). Thus, the rate-determining step (RDS) of the ORR is the last formation of (Na2O)2* (i.e., Na2O2* → (Na2O)2*) on all M2XO2 MXenes. For the OER pathway, after the downhill (Na2O)2* decomposition on Ti2XO2 and Zr2XO2, both the Na2O2* and NaO2* decomposition steps are uphill, but the NaO2* decomposition requires more energy than the Na2O2* decomposition (see Figure 8a,b,d,e). Moreover, only the last NaO2* decomposition step in the OER process is uphill on Hf2XO2 (see Figure 8c,f). Therefore, the NaO2* decomposition forms the RDS of the OER on all considered M2XO2 MXenes.
The discharging potential (UDs) is the highest voltage that drives all ORR steps energetically downhill on free energy profiles (the blue paths from left to right in Figure 8a–f). Meanwhile, the charging potential (UC) is the lowest potential that drives all OER steps energetically downhill on the free energy profiles (the pink paths from right to left in Figure 8a–f). For various M2XO2 MXenes, their UDc/UC values are different, which is caused by the different free energy changes of RDS during the ORR/OER process. To further evaluate the discharge–charge efficiency of Na−O2 batteries, the ORR ( η O R R ) and OER ( η O E R ) overpotentials were calculated by η O R R = U e U D c and η O E R = U C U e , respectively, while the total overpotential ( η T O T ) was calculated by η T O T = η O R R + η O E R , which are tabulated in Table 1. The η O R R / η O E R overpotentials are calculated to be 0.23/0.32, 0.47/0.88, 0.76/1.05, 1.03/2.68, 0.87/0.92, 0.76/1.05, and 1.25/2.85 V for Ti2CO2, Zr2CO2, Hf2CO2, Ti2NO2, Zr2NO2, and Hf2NO2, respectively. For all considered M2XO2 MXenes, the OER overpotentials are higher than the ORR overpotentials, which suggests that the OER kinetics are slower than the ORR kinetics, forming a bottleneck in the performance of Na−O2 batteries. This situation is in good agreement with the experimental results, where the hard decomposition of the discharge products results in high charging overpotentials, low round trip efficiency, and poor rechargeability [2]. As shown in Table 1, the total overpotential ( η T O T ) is calculated to be 0.55, 1.35, 3.71, 1.79, 1.81, and 4.10 V for Ti2CO2, Zr2CO2, Hf2CO2, Ti2NO2, Zr2NO2, and Hf2NO2, respectively, suggesting the M2CO2 carbides exhibit lower overpotential than the M2NO2 nitrides. Furthermore, the overpotentials of M2XO2 MXenes decrease with the decreasing atomic number of M. Among all considered M2XO2 MXenes, Ti2CO2 has the lowest overpotentials of η O R R (0.23 V), η O E R (0.32 V), and η T O T (0.55 V), indicating the highest catalytic activity. As shown in Figure 9, the η O R R , η O E R , and η T O T overpotentials of Ti2CO2 MXenes are much lower than those of Si2Se2 (0.62, 0.38, and 1.00 V) [39], MoSSe (0.49, 0.59, and 1.08 V) [3], SiSe2 (0.46, 0.73, and 1.19 V) [39], and NCF (0.56, 0.83, and 1.49 V) [40], indicating Ti2CO2 possesses an outstanding catalytic performance as the cathode for Na−O2 batteries.
For all catalysts, the RDS of the ORR process is the reduction of Na2O2* to (Na2O)2*. Therefore, the η O R R overpotential is determined by the adsorption energy of Na2O2* on M2XO2 MXenes. Meanwhile, the RDS of the OER process is the decomposition of NaO2*, and thus the NaO2* adsorption energy plays a dominant role in the η O E R overpotential. For Hf2NO2 MXene, the ultra-strong adsorption (see Figure 6b) makes the NaO2* and Na2O2* intermediates firmly adhere to the MXene surfaces, which is difficult to react further, thus forming the highest overpotentials (see Table 1). When Hf2CO2 carbides are used as catalysts, the adsorption energy of Na2O2* and NaO2* is reduced effectively, leading to improved overpotentials. Similarly, after reducing the metal atomic number, Ti2NO2 exhibits weaker adsorption energy towards Na2O2* and NaO2* and thus has lower overpotentials than Hf2NO2. For Ti2CO2, it has the weakest Na2O2* and NaO2* adsorption energy, resulting in the lowest η O R R and η O E R potentials.
As shown in Figure 5 and Figures S3–S5, the surface nucleophilic O groups of M2XO2 MXenes prefer to combine with the electropositive Na atoms of NaxO2 (x = 1, 2, and 4) to activate the Na–O bonds. Therefore, compared to free NaxO2, the Na–O bonds in NaxO2* are elongated efficiently, while the O–O bonds in NaxO2* are shortened (see Tables S2–S4), which is conducive to the deintercalation of sodium in the charging process. The charge population analyses in Tables S5–S7 indicate that a strong electron transfer from Na to surface O groups occurs on the NaxO2*-adsorbed M2XO2 MXenes, which weakens the ionic interaction in the Na–O bonds of NaxO2*. After the NaxO2* (x = 1, 2, and 4) adsorption on M2XO2 MXenes, the change in the O–O and Na–O bonds in (Na2O)2* is the smallest (Tables S2–S4), suggesting that (Na2O)2* is the most difficult to activate among NaxO2* intermediates, which is in good accordance with the experimental reports [25,37]. Furthermore, for all considered M2XO2 MXenes, M2CO2 carbides with a lower metal atomic number have a stronger activation effect on NaxO2* intermediates, as reflected by the greater change in the Na–O and O–O bond lengths of NaxO2* (see Tables S2–S4).
To further explore the effect of metal (M) and nonmetal (X) atoms, the PDOS analyses were carried out for the NaxO2* (x = 1 and 2)-adsorbed Ti2CO2, Ti2NO2, Ti2CO2, and Hf2NO2 MXenes (see Figure 10). As shown in Figure 10, after adsorption, NaxO2* forms strong ionic bonding and anti-bonding states with the MXene substrates. The bonding states located below the Fermi energy level are mainly composed of surface O p-orbitals (the orange color in Figure 10a–h), while the anti-bonding states located above the Fermi energy level are primarily composed of Na-s orbitals (the purple color in Figure 10a–h), indicating the electrons transfer from Na of NaxO2* to the surface O groups of M2XO2 substrates. The main peaks of the bonding and anti-bonding states are shown as the orange and purple dotted lines in Figure 10, respectively. The difference between the two main peaks represents the energy gap between the bonding and anti-bonding states. The energy gap is calculated to be 7.01, 7.53, 7.82, and 8.22 eV for NaO2*-adsorbed Ti2CO2, Ti2NO2, Hf2CO2, and Hf2NO2, respectively (see Figure 10a–d), while it is 7.28, 7.65, 7.82, and 8.56 eV for Na2O2*-adsorbed Ti2CO2, Ti2NO2, Hf2CO2, and Hf2NO2, respectively (see Figure 10e–h). The larger the energy gap, the stronger the interaction between the Na-s orbitals and surface O p-orbitals; that is, the stronger the adsorption of NaxO2 on the substrates. On the same substrate, the energy gap of Na2O2* is larger than that of NaO2*, suggesting stronger adsorption of Na2O2* than NaO2*. For the same NaxO2 adsorbent, the energy gap on the M2NO2 nitrides is greater than that on the M2CO2 carbides, indicating that the M2NO2 nitrides exhibit stronger adsorption towards NaxO2 than the M2CO2 carbides. Ti2CO2 has the smallest energy gap, so the interaction with the NaxO2 intermediate is the weakest, resulting in the lowest overpotential in the ORR/OER processes. Furthermore, the weakest interaction between Ti2CO2 and NaxO2 originates from the lowest p-band center of the surface O for Ti2CO2 among all M2XO2 MXenes (see Figure 7).

3. Materials and Methods

In this work, the Cambridge Serial Total Energy Package (CASTEP) was adopted for the DFT calculations [41,42]. The spin-polarized method was used to account for different wavefunctions for different spins. The Perdew–Burke–Ernzerhof (PBE) functional was chosen as the DFT exchange-correlation potential [43,44]. PBE may underestimate the band gaps and overestimate adsorption energies due to their self-interaction error and insufficient description of long-range electronic correlations. However, the metallic/semimetallic properties of MXenes make the underestimation of the bandgap have little impact on the catalytic activity (such as ORR/OER). The study of MXene systems demonstrates that computational approaches, including PBE, PBE+U, the meta-GGA SCAN functional, and HSE06, yield highly comparable results in structural, electronic, and magnetic properties [45,46]. Additionally, the error in adsorption energy can be partially corrected using DFT-D3. The PBE functional combined with DFT-D3 dispersion correction is the most widely used and validated approach for MXenes systems, particularly in catalytic studies involving surface adsorption [47,48,49]. The DFT-D3 method has proven effective for both molecules and solids, achieving a CCSD(T) accuracy typically within 10% [50]. Therefore, PBE-D3 was used in this work. The valence electrons were described using a plane wave basis, and the norm-conserving pseudopotential represented the ionic cores [51,52]. After convergence testing (see Figures S6 and S7), the 6 × 6 × 1 k-point Monkhorst–Pack mesh and the 500 eV cutoff energy of the plane wave were chosen for all calculations. The convergence criteria for geometric optimization were as follows: an energy less than 10−5 eV atom−1, a force less than 0.03 eV Å−1, a stress less than 0.05 Gpa, and a displacement less than 0.001 Å. A 20 Å vacuum layer was added to the surface of M2XO2 MXenes to eliminate the influence of mirror structures.
The formation energy (Ef) of M2XO2 MXenes was determined by
E f = E t o t a l n M E M n X E X n O E O / n t o t a l
where Etotal is the energy of the M2XO2 system, EM is the energy of a single M (Ti, Zr, or Hf) atom in bulk metal, EX is the energy of a single X (C or N) atom in graphene or N2, ntotal is the total atom number in the M2XO2 system, and nM, nX, and nO are the numbers of the M, X, and O atoms in the M2XO2 system, respectively.
The adsorption energy (Eads) was determined by
E a d s = E t o t E s u b s t r a t e E a d s o r b a t e
where Etot, Esubstrate, and Eadsorbate denote the energy of the adsorption system, the substrate, and the adsorbate, respectively.
During the discharging/charging process, the free energy change (ΔG) of the intermediates in each step was described as:
G = E + E Z P E T S n e U
where ΔE, ΔEZPE, and ΔS denote the calculated changes in the energy, zero point energy, and entropy at the temperature of T = 298 K, respectively. n is the number of transferred electrons at the potential of U. The theoretical overpotentials ƞ for the ORR and OER were calculated by ƞORR = UeUDc and ƞOER = UCUe, respectively. In this definition, UDc and UC are the discharge and charge potentials calculated by UDc = min (ΔG/ne) and UC = max (ΔG/ne), respectively. Ue is the equilibrium potential calculated by Ue = −ΔG/ne.
The p-band center (ɛp) was defined as
[ ε p = E ρ p ( E ) d E ρ p ( E ) d E ]
where ρ p ( E ) is the density of the p-state at the level of E.

4. Conclusions

In this work, six carbide/nitride MXene models (M2XO2, M = Ti, Zr, and Hf, X = C and N) were constructed to evaluate their catalytic activity as cathodes for Na−O2 batteries. The first-principles calculations suggest that the M2CO2 MXenes show semiconductor-like properties, mirrored by the pseudogaps around the Fermi level, while the M2NO2 MXenes are conductors, as reflected by the high intensity of DOSs at the Fermi level. For all M2XO2 MXenes, the OER overpotential during charging is greater than the ORR overpotential during discharging, forming a performance bottleneck of Na−O2 batteries. The surface nucleophilic O groups of M2XO2 MXenes tend to bind with the Na atoms of NaxO2 intermediates to activate the Na−O bonds, which is conducive to the deintercalation of sodium in the charging process. However, the too-strong adsorption of NaxO2 intermediates leads to high overpotentials on M2XO2 MXenes. Due to the downward shift of the p-band centers of the surface O groups, the NaxO2 adsorption on the M2CO2 carbides is weaker than that on the M2NO2 nitrides and is further weakened with a decreasing M atomic number. Therefore, the overpotentials decrease in the order of Hf2NO2 > Hf2CO2 > Zr2NO2 > Ti2NO2 > Zr2CO2 > Ti2CO2. Notably, the Ti2CO2 MXene shows ultralow ORR, OER, and total overpotentials (0.23, 0.32, and 0.55 V), making it an excellent candidate for cathodes in Na−O2 batteries.
Building on these findings, future studies could focus on (1) exploring diverse MXene compositions (e.g., M₃X2, M₄X₃) or transition metals (V, Mo) to tailor electronic/adsorption properties, (2) surface modifications (e.g., heteroatom doping) to fine-tune intermediate binding and reduce overpotentials, and (3) experimental validation of the stability and catalytic performance of Ti2CO2 MXenes for practical application. Integrating machine learning with computational screening may expedite MXene discovery, while in situ techniques could unravel dynamic reaction mechanisms. These efforts would bridge theoretical insights with real-world battery performance, fostering the development of efficient and durable Na–O2 systems.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal15040311/s1, Figure S1: Top and side views of M2X Mxenes; Figure S2: Top and side views of M2XO2 Mxenes; Figure S3: Top and side views of NaO2-adsorbed M2CO2 and M2NO2 Mxenes; Figure S4: Top and side views of Na2O2-adsorbed M2CO2 and M2NO2 Mxenes; Figure S5: Top and side views of (Na2O)2-adsorbed M2CO2 and M2NO2 Mxenes; Figure S6: The total energy of Ti2CO2 as a function of the k-point; Figure S7: The total energy of Ti2CO2 as a function of the cutoff energy; Table S1: Formation energy of M2XO2 MXenes; Table S2: The Na−O and O−O bond lengths in NaO2 on M2XO2 Mxenes; Table S3: The Na−O and O−O bond lengths in Na2O2 on M2XO2 Mxenes; Table S4: The Na−O and O−O bond lengths in Na4O2 on M2XO2 Mxenes; Table S5: Hirshfeld charge of the NaO2-adsorbed M2XO2 Mxenes; Table S6: Hirshfeld charge of the Na2O2-adsorbed M2XO2 Mxenes; Table S7: Hirshfeld charge of the Na4O2-adsorbed M2XO2 Mxenes.

Author Contributions

L.Z.: investigation, conceptualization, methodology, validation, and writing—original draft; Z.J.: investigation; T.D.: data curation; Z.P.: visualization; M.L.: methodology; H.R.: writing—review and editing; J.X.: writing—review and editing, methodology, supervision, and funding acquisition; W.X.: resources and project administration. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (U23B2087) and Shandong Provincial Natural Science Foundation (ZR2022MB094).

Data Availability Statement

Data will be made available on request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zhu, Y.; Ge, M.; Ma, F.; Wang, Q.; Huang, P.; Lai, C. Multifunctional electrolyte additives for better metal batteries. Adv. Funct. Mater. 2024, 34, 2301964. [Google Scholar] [CrossRef]
  2. Khan, Z.; Vagin, M.; Crispin, X. Can hybrid Na–Air batteries outperform nonaqueous Na–O2 batteries? Adv. Sci. 2020, 7, 1902866. [Google Scholar] [CrossRef] [PubMed]
  3. Wu, Q.; Ma, Y.; Wang, H.; Zhang, S.; Huang, B.; Dai, Y. Trifunctional electrocatalysts with high efficiency for the oxygen reduction reaction, oxygen evolution reaction, and Na–O2 battery in heteroatom-doped Janus monolayer MoSSe. ACS Appl. Mater. Interfaces 2020, 12, 24066–24073. [Google Scholar] [CrossRef] [PubMed]
  4. Lutz, L.; Dachraoui, W.; Demortière, A.; Johnson, L.R.; Bruce, P.G.; Grimaud, A.; Tarascon, J.-M. Operando monitoring of the solution-mediated discharge and charge processes in a Na–O2 battery using liquid-electrochemical transmission electron microscopy. Nano Lett. 2018, 18, 1280–1289. [Google Scholar] [CrossRef]
  5. Von Gunten, A.; Velinkar, K.; Nikolla, E.; Greeley, J. Elucidation of parasitic reaction mechanisms at interfaces in Na–O2 batteries. Chem. Mater. 2023, 35, 5945–5952. [Google Scholar] [CrossRef]
  6. Qin, B.; Wang, L.; Tsui, C.K.J.; Ho, C.-K.; Lam, W.-Y.A.; Li, F.; Li, C.-Y.V.; Chen, G.; Chan, K.-Y. Impacts of carbon mesopores on superoxide degradation in the cathode of a Na–O2 battery. Energy Fuels 2024, 38, 5522–5533. [Google Scholar] [CrossRef]
  7. Yuan, R.; Tan, C.; Zhang, Z.; Zeng, L.; Kang, W.; Liu, J.; Gao, X.; Tan, P.; Chen, Y.; Zhang, C. Topological engineering electrodes with ultrafast oxygen transport for super-power sodium-oxygen batteries. Adv. Mater. 2024, 36, 2311627. [Google Scholar] [CrossRef]
  8. Beermann, V.; Gocyla, M.; Willinger, E.; Rudi, S.; Heggen, M.; Dunin-Borkowski, R.E.; Willinger, M.-G.; Strasser, P. Rh-doped Pt–Ni octahedral nanoparticles: Understanding the correlation between elemental distribution, oxygen reduction reaction, and shape stability. Nano Lett. 2016, 16, 1719–1725. [Google Scholar] [CrossRef]
  9. Egorova, K.S.; Ananikov, V.P. Which metals are green for catalysis? Comparison of the toxicities of Ni, Cu, Fe, Pd, Pt, Rh, and Au salts. Angew. Chem. Int. Ed. 2016, 55, 12150–12162. [Google Scholar] [CrossRef]
  10. Jeong, Y.S.; Park, J.-B.; Jung, H.-G.; Kim, J.; Luo, X.; Lu, J.; Curtiss, L.; Amine, K.; Sun, Y.-K.; Scrosati, B.; et al. Study on the catalytic activity of noble metal nanoparticles on reduced graphene oxide for oxygen evolution reactions in lithium–air batteries. Nano Lett. 2015, 15, 4261–4268. [Google Scholar] [CrossRef]
  11. Yao, Y.; Wu, F. Turning waste chemicals into wealth—A new approach to synthesize efficient cathode material for an Li–O2 battery. ACS Appl. Mater. Interfaces 2017, 9, 31907–31912. [Google Scholar] [PubMed]
  12. Zhang, S.; Wen, Z.; Rui, K.; Shen, C.; Lu, Y.; Yang, J. Graphene nanosheets loaded with Pt nanoparticles with enhanced electrochemical performance for sodium–oxygen batteries. J. Mater. Chem. A 2015, 3, 2568–2571. [Google Scholar] [CrossRef]
  13. Benti, N.E.; Tiruye, G.A.; Mekonnen, Y.S. Boron and pyridinic nitrogen-doped graphene as potential catalysts for rechargeable non-aqueous sodium–air batteries. RSC Adv. 2020, 10, 21387–21398. [Google Scholar] [PubMed]
  14. Munuera, J.M.; Paredes, J.I.; Enterría, M.; Villar-Rodil, S.; Kelly, A.G.; Nalawade, Y.; Coleman, J.N.; Rojo, T.; Ortiz-Vitoriano, N.; Martínez-Alonso, A.; et al. High performance Na-O2 batteries and printed microsupercapacitors based on water-processable, biomolecule-assisted anodic graphene. ACS Appl. Mater. Interfaces 2020, 12, 494–506. [Google Scholar] [PubMed]
  15. Liu, Y.; Chi, X.; Han, Q.; Du, Y.; Yang, J.; Liu, Y. Vertically self-standing C@NiCo2O4 nanoneedle arrays as effective binder-free cathodes for rechargeable Na−O2 batteries. J. Alloys Compd. 2019, 772, 693–702. [Google Scholar]
  16. Wang, J.; Gao, R.; Zheng, L.; Chen, Z.; Wu, Z.; Sun, L.; Hu, Z.; Liu, X. CoO/CoP heterostructured nanosheets with an O–P interpenetrated interface as a bifunctional electrocatalyst for Na–O2 battery. ACS Catal. 2018, 8, 8953–8960. [Google Scholar] [CrossRef]
  17. Brady, A.; Liang, K.; Vuong, V.Q.; Sacci, R.; Prenger, K.; Thompson, M.; Matsumoto, R.; Cummings, P.; Irle, S.; Wang, H.-W.; et al. Pre-sodiated Ti3C2Tx MXene structure and behavior as electrode for sodium-ion capacitors. ACS Nano 2021, 15, 2994–3003. [Google Scholar]
  18. Verma, S.; Padha, B.; Young, S.-J.; Chu, Y.-L.; Bhardwaj, R.; Mishra, R.K.; Arya, S. 3D MXenes for supercapacitors: Current status, opportunities and challenges. Prog. Solid State Chem. 2023, 72, 100425. [Google Scholar]
  19. Shetti, N.P.; Mishra, A.; Basu, S.; Aminabhavi, T.M.; Alodhayb, A.; Pandiaraj, S. MXenes as Li-ion battery electrodes: Progress and outlook. Energy Fuels 2023, 37, 12541–12557. [Google Scholar]
  20. Zhang, Y.; Lu, Q.; Zhang, L.; Zhang, L.; Shao, G.; Zhang, P. Adjustable MXene-based materials in metal-ion batteries: Progress, prospects, and challenges. Small Struct. 2024, 5, 2300255. [Google Scholar]
  21. Dai, X.; Zhang, W.; Sun, Y.; Du, Z.; Tao, Z.; Wang, J.; Fang, W.; Xing, X.; Chen, Y.; Li, H.; et al. Niobium oxide/MXene heterostructure for simultaneous production of ammonia and energy via rechargeable Zn-N2 battery system. J. Energy Chem. 2025, 103, 448–457. [Google Scholar]
  22. Zhu, L.; Wang, J.; Liu, J.; Wang, R.; Lin, M.; Wang, T.; Zhen, Y.; Xu, J.; Zhao, L. First principles study of the structure–performance relation of pristine Wn+1Cn and oxygen-functionalized Wn+1CnO2 MXenes as cathode catalysts for Li-O2 batteries. Nanmaterials 2024, 14, 666. [Google Scholar]
  23. Mudassar Aslam, M.; Noor, T.; Iqbal, N. Advances in MXenes synthesis and MXenes derived electrocatalysts for oxygen electrode in metal-air batteries: A review. Mater. Sci. Eng. B 2023, 292, 116400. [Google Scholar]
  24. Min, Y.; Yuan, H.; Wang, W.; Xu, L. Design of heterostructures of MXene/Two-dimensional organic frameworks for Na–O2 batteries with a new mechanism and a new descriptor. J. Phys. Chem. Lett. 2021, 12, 2742–2748. [Google Scholar] [PubMed]
  25. Tang, C.; Min, Y.; Chen, C.; Xu, W.; Xu, L. Potential applications of heterostructures of TMDs with MXenes in sodium-ion and Na–O2 batteries. Nano Lett. 2019, 19, 5577–5586. [Google Scholar]
  26. He, X.; Jin, S.; Miao, L.; Cai, Y.; Hou, Y.; Li, H.; Zhang, K.; Yan, Z.; Chen, J. A 3D hydroxylated MXene/carbon nanotubes composite as a scaffold for dendrite-free sodium-metal electrodes. Angew. Chem. Int. Ed. 2020, 59, 16705–16711. [Google Scholar]
  27. Huang, W.; Zhang, J.; Deng, G.; Zhu, G.; Chen, Y.; Xu, C.; Cheng, J. MXene-supported Co–S–N–C catalysts with enhanced oxygen reduction reaction activity for anion exchange membrane fuel cells. ACS Appl. Energy Mater. 2025, 8, 2612–2619. [Google Scholar]
  28. Tan, L.; Wang, J.; Zhou, S.; Zhu, H.; Guo, J.; Chen, Y.; Li, X.; Dong, Z.; Zhang, Q.; Cong, Y. NiFe phosphides coupled on Ti3C2Tx MXene nanosheets for high-efficiency oxygen evolution reaction in alkaline medium. J. Colloid Interface Sci. 2025, 689, 137263. [Google Scholar]
  29. Shen, Y.; Lv, H.; Chen, L. Recent advances in two-dimensional MXenes for zinc-ion batteries. Mater. Chem. Front. 2023, 7, 2373–2404. [Google Scholar]
  30. Liu, C.; Yang, Y.; Tang, K.; Wu, F.; Liu, Y.; Yang, Z.; Chai, Y.; Sun, J. Properties of Ti2CO2 and Ti2CO2/G heterostructures as anodes of sodium-ion batteries by first-principles study. Theor. Chem. Acc. 2024, 143, 58. [Google Scholar]
  31. Liu, C.-Y.; Li, E.Y. Termination effects of Pt/v-Tin+1CnT2 MXene surfaces for oxygen reduction reaction catalysis. ACS Appl. Mater. Interfaces 2019, 11, 1638–1644. [Google Scholar] [PubMed]
  32. Wang, L.; Dou, Y.; Gan, R.; Zhao, Q.; Ma, Q.; Liao, Y.; Cheng, G.; Zhang, Y.; Wang, D. The single atom anchoring strategy: Rational design of MXene-based single-atom catalysts for electrocatalysis. Small 2025, 21, 2410772. [Google Scholar]
  33. Yorulmaz, U.; Özden, A.; Perkgöz, N.K.; Ay, F.; Sevik, C. Vibrational and mechanical properties of single layer MXene structures: A first-principles investigation. Nanotechnology 2016, 27, 335702. [Google Scholar] [PubMed]
  34. Liu, M.-Z.; Li, X.-H.; Cui, X.-H.; Yan, H.-T.; Zhang, R.-Z.; Cui, H.-L. The influence of different functional groups on quantum capacitance, electronic and optical properties of Hf2C MXene. Appl. Surf. Sci. 2022, 605, 154830. [Google Scholar]
  35. Mekonnen, Y.S.; Christensen, R.; Garcia-Lastra, J.M.; Vegge, T. Thermodynamic and kinetic limitations for peroxide and superoxide formation in Na–O2 batteries. J. Phys. Chem. Lett. 2018, 9, 4413–4419. [Google Scholar]
  36. Krishnamurthy, D.; Hansen, H.A.; Viswanathan, V. Universality in nonaqueous alkali oxygen reduction on metal surfaces: Implications for Li–O2 and Na–O2 batteries. ACS Energy Lett. 2016, 1, 162–168. [Google Scholar]
  37. Kumar, S.; Kishore, B.; Munichandraiah, N. Electrochemical studies of non-aqueous Na–O2 cells employing Ag-RGO as the bifunctional catalyst. RSC Adv. 2016, 6, 63477–63479. [Google Scholar]
  38. Tatara, R.; Leverick, G.M.; Feng, S.; Wan, S.; Terada, S.; Dokko, K.; Watanabe, M.; Shao-Horn, Y. Tuning NaO2 cube sizes by controlling Na+ and solvent activity in Na–O2 batteries. J. Phys. Chem. C 2018, 122, 18316–18328. [Google Scholar]
  39. Wang, J.; Zhang, K.; Pan, M.; Liu, Z.; Deng, H. Theoretically evaluating two-dimensional tetragonal Si2Se2 and SiSe2 nanosheets as cathode catalysts for alkali metal–O2 batteries. J. Phys. Chem. C 2023, 127, 21033–21046. [Google Scholar]
  40. Zheng, Z.; Jiang, J.; Guo, H.; Li, C.; Konstantinov, K.; Gu, Q.; Wang, J. Tuning NaO2 formation and decomposition routes with nitrogen-doped nanofibers for low overpotential Na-O2 batteries. Nano Energy 2021, 81, 105529. [Google Scholar]
  41. Clark, S.J.; Segall, M.D.; Pickard, C.J.; Hasnip, P.J.; Probert, M.I.J.; Refson, K.; Payne, M.C. First principles methods using CASTEP. Z. Kristallogr. Cryst. Mater. 2005, 220, 567–570. [Google Scholar]
  42. Payne, M.C.; Teter, M.P.; Allan, D.C.; Arias, T.A.; Joannopoulos, J.D. Iterative minimization techniques for ab initio total-energy calculations: Molecular dynamics and conjugate gradients. Rev. Mod. Phys. 1992, 64, 1045–1097. [Google Scholar]
  43. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [PubMed]
  44. Grimme, S. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar]
  45. Oschinski, H.; Morales-García, Á.; Illas, F. Interaction of first row transition metals with M2C (M = Ti, Zr, Hf, V, Nb, Ta, Cr, Mo, and W) MXenes: A quest for single-atom catalysts. J. Phys. Chem. C 2021, 125, 2477–2484. [Google Scholar]
  46. Bae, S.; Kang, Y.-G.; Khazaei, M.; Ohno, K.; Kim, Y.-H.; Han, M.J.; Chang, K.J.; Raebiger, H. Electronic and magnetic properties of carbide MXenes—The role of electron correlations. Mater. Today Adv. 2021, 9, 100118. [Google Scholar]
  47. Barman, S.C.; Jin, Y.; El-Demellawi, J.K.; Thomas, S.; Wehbe, N.; Lei, Y.; Hota, M.K.; Xu, X.; Hasan, E.A.; Mohammed, O.F.; et al. Antibody-functionalized MXene-based electrochemical biosensor for point-of-care detection of vitamin D deficiency. Commun. Mater. 2025, 6, 31. [Google Scholar]
  48. Saharan, S.; Ghanekar, U.; Meena, S. Sulphur-decorated Ti3C2 MXene structures as high-capacity electrode for Zn-ion batteries: A DFT study. Nanoscale 2025. [Google Scholar] [CrossRef]
  49. Meng, L.; Pokochueva, E.V.; Chen, Z.; Fedorov, A.; Viñes, F.; Illas, F.; Koptyug, I.V. Contrasting metallic (Rh0) and carbidic (2D-Mo2C MXene) surfaces in olefin hydrogenation provides insights on the origin of the pairwise hydrogen addition. ACS Catal. 2024, 14, 12500–12511. [Google Scholar]
  50. 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. 2010, 132, 154104. [Google Scholar]
  51. Hamann, D.R.; Schlüter, M.; Chiang, C. Norm-conserving pseudopotentials. Phys. Rev. Lett. 1979, 43, 1494–1497. [Google Scholar]
  52. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar]
Figure 1. (a,b) Top and (c,d) side views of (a,c) M2X and (b,d) M2XO2 MXenes (M = Ti, Zr, or Hf; X = C or N). Top, hcp, and fcc indicate possible adsorption sites.
Figure 1. (a,b) Top and (c,d) side views of (a,c) M2X and (b,d) M2XO2 MXenes (M = Ti, Zr, or Hf; X = C or N). Top, hcp, and fcc indicate possible adsorption sites.
Catalysts 15 00311 g001
Figure 2. Projected density of the states (PDOS) of surface O atoms for (a) Ti2CO2, (b) Zr2CO2, (c) Hf2CO2, (d) Ti2NO2, (e) Zr2NO2, and (f) Hf2NO2. The Fermi level marked by black dashed lines is set as the energy zero. The p-band center of surface O atoms is marked by the red dashed lines.
Figure 2. Projected density of the states (PDOS) of surface O atoms for (a) Ti2CO2, (b) Zr2CO2, (c) Hf2CO2, (d) Ti2NO2, (e) Zr2NO2, and (f) Hf2NO2. The Fermi level marked by black dashed lines is set as the energy zero. The p-band center of surface O atoms is marked by the red dashed lines.
Catalysts 15 00311 g002
Figure 3. Hirshfeld charge for M2XO2 MXenes: q(O), q(metal), and q(nonmetal) are the Hirshfeld charges of surface O groups, metal M atoms, and C or N atoms in M2XO2 MXenes, respectively.
Figure 3. Hirshfeld charge for M2XO2 MXenes: q(O), q(metal), and q(nonmetal) are the Hirshfeld charges of surface O groups, metal M atoms, and C or N atoms in M2XO2 MXenes, respectively.
Catalysts 15 00311 g003
Figure 4. The sliced deformation electron densities (SDED) of (a) Ti2CO2, (b) Zr2CO2, (c) Hf2CO2, (d) Ti2NO2, (e) Zr2NO2, and (f) Hf2NO2.
Figure 4. The sliced deformation electron densities (SDED) of (a) Ti2CO2, (b) Zr2CO2, (c) Hf2CO2, (d) Ti2NO2, (e) Zr2NO2, and (f) Hf2NO2.
Catalysts 15 00311 g004
Figure 5. The top (left three columns) and side (right three columns) views of (a) Ti2CO2 and (b) Ti2NO2 Mxenes with the adsorbed NaxO2. * indicates that the intermediate is in an adsorbed state.
Figure 5. The top (left three columns) and side (right three columns) views of (a) Ti2CO2 and (b) Ti2NO2 Mxenes with the adsorbed NaxO2. * indicates that the intermediate is in an adsorbed state.
Catalysts 15 00311 g005
Figure 6. The adsorption energy of NaO2, Na2O2, and Na4O2 on (a) M2CO2 and (b) M2NO2 MXenes.
Figure 6. The adsorption energy of NaO2, Na2O2, and Na4O2 on (a) M2CO2 and (b) M2NO2 MXenes.
Catalysts 15 00311 g006
Figure 7. The linear fit of the adsorption energy of (a) NaO2, (b) Na2O2, and (c) Na4O2 with the p-band center of the O atoms (εp(O)) for M2XO2 MXenes.
Figure 7. The linear fit of the adsorption energy of (a) NaO2, (b) Na2O2, and (c) Na4O2 with the p-band center of the O atoms (εp(O)) for M2XO2 MXenes.
Catalysts 15 00311 g007
Figure 8. The calculated free energy profiles of the ORR and OER processes on (a) Ti2CO2, (b) Zr2CO2, (c) Hf2CO2, (d) Ti2NO2, (e) Zr2NO2, and (f) Hf2NO2. * indicates that the intermediate is in an adsorbed state.
Figure 8. The calculated free energy profiles of the ORR and OER processes on (a) Ti2CO2, (b) Zr2CO2, (c) Hf2CO2, (d) Ti2NO2, (e) Zr2NO2, and (f) Hf2NO2. * indicates that the intermediate is in an adsorbed state.
Catalysts 15 00311 g008
Figure 9. Comparison of overpotentials for the Ti2CO2 MXene with other materials. The catalyst marked in red indicates the minimum overpotential.
Figure 9. Comparison of overpotentials for the Ti2CO2 MXene with other materials. The catalyst marked in red indicates the minimum overpotential.
Catalysts 15 00311 g009
Figure 10. Projected density of states (PDOS) for (ad) NaO2* and (eh) Na2O2* adsorbed on (a,e) Ti2CO2, (b,f) Ti2NO2, (c,g) Hf2CO2, and (d,h) Hf2NO2 MXenes. The dashed line indicates the position of the PDOS peak. * indicates that the intermediate is in an adsorbed state.
Figure 10. Projected density of states (PDOS) for (ad) NaO2* and (eh) Na2O2* adsorbed on (a,e) Ti2CO2, (b,f) Ti2NO2, (c,g) Hf2CO2, and (d,h) Hf2NO2 MXenes. The dashed line indicates the position of the PDOS peak. * indicates that the intermediate is in an adsorbed state.
Catalysts 15 00311 g010
Table 1. The ORR ( η O R R ), OER ( η O E R ), and total ( η T O T ) overpotentials (in V) for M2XO2 MXenes.
Table 1. The ORR ( η O R R ), OER ( η O E R ), and total ( η T O T ) overpotentials (in V) for M2XO2 MXenes.
OverpotentialsTi2CO2Ti2NO2Zr2CO2Zr2NO2Hf2CO2Hf2NO2
η O R R 0.230.870.470.761.031.25
η O E R 0.320.920.881.052.682.85
η T O T 0.551.791.351.813.714.10
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhao, L.; Jiang, Z.; Ding, T.; Peng, Z.; Lin, M.; Ren, H.; Xu, J.; Xing, W. In-Depth First-Principles Study of High-Performance M2XO2 MXene Cathode Catalysts for Sodium-Oxygen Batteries. Catalysts 2025, 15, 311. https://doi.org/10.3390/catal15040311

AMA Style

Zhao L, Jiang Z, Ding T, Peng Z, Lin M, Ren H, Xu J, Xing W. In-Depth First-Principles Study of High-Performance M2XO2 MXene Cathode Catalysts for Sodium-Oxygen Batteries. Catalysts. 2025; 15(4):311. https://doi.org/10.3390/catal15040311

Chicago/Turabian Style

Zhao, Lianming, Zhumei Jiang, Tao Ding, Zeyue Peng, Meixin Lin, Hao Ren, Jing Xu, and Wei Xing. 2025. "In-Depth First-Principles Study of High-Performance M2XO2 MXene Cathode Catalysts for Sodium-Oxygen Batteries" Catalysts 15, no. 4: 311. https://doi.org/10.3390/catal15040311

APA Style

Zhao, L., Jiang, Z., Ding, T., Peng, Z., Lin, M., Ren, H., Xu, J., & Xing, W. (2025). In-Depth First-Principles Study of High-Performance M2XO2 MXene Cathode Catalysts for Sodium-Oxygen Batteries. Catalysts, 15(4), 311. https://doi.org/10.3390/catal15040311

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop