Next Article in Journal
Aluminum Micropillar Surfaces with Hierarchical Micro- and Nanoscale Features for Enhancement of Boiling Heat Transfer Coefficient and Critical Heat Flux
Previous Article in Journal
Impact of Nanoparticle Addition on the Surface and Color Properties of Three-Dimensional (3D) Printed Polymer-Based Provisional Restorations
Previous Article in Special Issue
Density Functional Theory Study of Methanol Steam Reforming on Pt3Sn(111) and the Promotion Effect of a Surface Hydroxy Group
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

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

School of Materials Science and Engineering, China University of Petroleum (East China), Qingdao 266580, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2024, 14(8), 666; https://doi.org/10.3390/nano14080666
Submission received: 18 March 2024 / Revised: 6 April 2024 / Accepted: 9 April 2024 / Published: 11 April 2024

Abstract

:
Li-O2 batteries are considered a highly promising energy storage solution. However, their practical implementation is hindered by the sluggish kinetics of the oxygen reduction (ORR) and oxygen evolution (OER) reactions at cathodes during discharging and charging, respectively. In this work, we investigated the catalytic performance of Wn+1Cn and Wn+1CnO2 MXenes (n = 1, 2, and 3) as cathodes for Li-O2 batteries using first principles calculations. Both Wn+1Cn and Wn+1CnO2 MXenes show high conductivity, and their conductivity is further enhanced with increasing atomic layers, as reflected by the elevated density of states at the Fermi level. The oxygen functionalization can change the electronic properties of WC MXenes from the electrophilic W surface of Wn+1Cn to the nucleophilic O surface of Wn+1CnO2, which is beneficial for the activation of the Li-O bond, and thus promotes the Li+ deintercalation during the charge–discharge process. On both Wn+1Cn and Wn+1CnO2, the rate-determining step (RDS) of ORR is the formation of the (Li2O)2* product, while the RDS of OER is the LiO2* decomposition. The overpotentials of ORR and OER are positively linearly correlated with the adsorption energy of the RDS LixO2* intermediates. By lowering the energy band center, the oxygen functionalization and increasing atomic layers can effectively reduce the adsorption strength of the LixO2* intermediates, thereby reducing the ORR and OER overpotentials. The W4C3O2 MXene shows immense potential as a cathode catalyst for Li-O2 batteries due to its outstanding conductivity and super-low ORR, OER, and total overpotentials (0.25, 0.38, and 0.63 V).

Graphical Abstract

1. Introduction

Energy conversion and storage systems have become a critical component of the future energy sector as global energy demand continues to grow and energy transformation accelerates. Among these systems, Li-O2 batteries are considered one of the most promising solutions for energy storage due to their high energy density (up to 3500 Wh·kg−1), long lifespan, and environmental friendliness [1]. Li-O2 batteries typically are composed of lithium metal anodes, oxygen cathodes, and non-aqueous Li+ conductive electrolytes [2]. During discharge, atmospheric oxygen reacts with (Li+ + e-−) pairs to form (Li2O)2, which undergoes an oxygen reduction reaction (ORR) on the cathode:
4 L i + + 4 e + O 2 ( L i 2 O ) 2
Upon charging, (Li2O)2 is converted back to (Li+ + e) and O2. Therefore, the oxygen evolution reaction (OER) takes place at the cathode [3]. However, the sluggish kinetics of ORR and OER at the cathode lead to increased charge and discharge overpotentials, limited discharge capacity, and inadequate cycling performance of Li-O2 batteries, thereby limiting their practical applicability [4]. Furthermore, due to the poor conductivity of the discharge product Li4O2, its decomposition during charging requires a significant overpotential, resulting in the decomposition of the solvent at high voltage. Additionally, other components in the air, such as CO2 and H2O, may react with Li2O2 to form by-products, such as Li2CO3 and LiOH, which are more difficult to decompose. Finally, the battery reaches a fault state [5,6]. Consequently, the development and synthesis of efficient cathode catalysts play a pivotal role in enhancing the performance of Li-O2 batteries. To date, several categories of catalyst materials have been employed in Li-O2 batteries, including noble metals and their alloys (Pt [7] and Pt-Au alloys [8]), functional carbon materials (graphene [9] and carbon nanotube [10]), transition metal oxides (Mn3O4 [11] and Co3O4 [12]), transition metal carbides/nitrides (Mo2C [13] and MoN [1]), metal-organic frameworks (Ru-MOF-C [14] and Tz-Mg-MOF-74 [15]), and other composite structure electrocatalysts [16,17].
In recent years, MXenes have attracted considerable attention in the field of electrocatalysis due to their unique properties as two-dimensional (2D) layered nitride or carbide materials, including low resistivity, fast ion transport, and tunable interlayer structure [18]. MXenes are typically fabricated utilizing three layers of the MAX phase as a precursor, followed by etching the A layer via various techniques and modifying functional groups on the surface. The general formula for MAX phases is Mn+1AXn. The synthesized MXenes can be represented as Mn+1XnTx, where M is a transition metal (such as Cr, Ti, Mn, Mo, or W), A is usually a group 13 or 14 elements (such as Al, Si, Ga, or Ge), X is C or N, and T represents the end of surface (such as O, F, or OH). Nowadays, various MXenes have been used in Li-O2 batteries [19] and supercapacitors [20]. For example, Xu et al. reported that the O-terminated V2C MXene (V2CO2) can significantly reduce battery overpotential to 0.75 V, increase capacity to 8577 mAh g−1 at 100 mA g−1, and improve durability up to 302 cycles for Li-O2 batteries. The electrocatalytic activity of V2CO2 is enhanced by improving the affinity for the substrate with Li2O2 through O-termination. Moreover, the unique 2D V2CO2 structure exhibits remarkable conductivity and excellent mass transfer performance during battery operation, thereby effectively optimizing the dynamics of Li-O2 batteries [18]. The Nb2C MXene nano-sheets with uniform O-terminal surfaces were fabricated as a high-rate cathode for Li-O2 batteries by Li et al. [21]. This catalyst exhibits a large capacity of 19785.5 mAh g−1 and a high-rate stability of 130 cycles at 200 mA g−1 and 3 A g−1 [21]. Density functional theory (DFT) calculations indicate that the O-terminated Nb2C MXene can enhances its affinity with LiO2 and Li2O2, facilitating spatial orientation accumulation and stable decomposition of discharge products [21]. These findings highlight the significant potential of MXenes in Li-O2 batteries. However, further research is needed to elucidate the relationship between MXene structure and activity in Li-O2 batteries, especially regarding the influence of surface functional groups and the number of atomic layers of Mn+1XnTx MXene on catalytic performance.
In light of these formidable challenges, we systematically investigate the catalytic performance of WC-MXenes (i.e., Wn+1Cn and Wn+1CnO2, n = 1, 2, and 3) as cathode materials for Li-O2 batteries through first principles calculations. The WC-MXenes were selected for the following benefits: (a) W-based materials exhibit excellent mechanical strength, outstanding chemical stability, and high tolerance to acidic environments, and are naturally abundant and environmentally friendly [22,23]. (b) The WxCy materials have been proven Pt-like catalytic features in many electrochemical reactions, such as ORR and OER [24]. (c) WC MXenes have a metallic nature, high charge capacity, and low Li+ diffusion barrier, facilitating the rapid deintercalation of Li+ during the charge–discharge process [25,26,27].
Herein, the electrochemical catalytic models of the WC MXene cathode were established to simulate OER and ORR during the charge–discharge process. The relationship between structure and catalytic performance of WC MXenes was revealed. In particular, the effects of surface oxygen functionalization and atomic layer number on the electronic structure of WC MXenes were explored to modulate the ORR and OER activities. We focused solely on the intrinsic properties of the electrocatalysts. External factors such as discharge product morphology, electrolyte decomposition, side reactions with CO2 and H2O, lithium metal corrosion, and oxygen electrode polarization were not considered in this work. Initially, the geometric and electronic structures of pristine and functionalized WC MXenes were investigated, revealing that both pristine Wn+1Cn and O-functionalized Wn+1CnO2 (n = 1, 2, and 3) show excellent electrical conductivity. Furthermore, the influence of atomic layer number and oxygen functional groups on the electronic properties and catalytic activity of WC MXenes was examined. Then, the formation and reversible decomposition of LixO2 (x = 1, 2, and 4) were simulated during charge–discharge processes. Finally, the overpotentials of WC MXenes were quantitatively calculated to evaluate their catalytic performance for Li-O2 batteries. Our study demonstrates that oxygen functionalization can convert the electrophilic W surface of Wn+1Cn to the nucleophilic O surface of Wn+1CnO2, promoting the deintercalation of Li+ during the charge–discharge process. The ORR and OER overpotentials are positively linearly correlated with the adsorption energy of the LixO2* intermediates. Surface oxygen functionalization and increasing atomic layers can weaken the LixO2 adsorption by lowering the energy band center of WC MXenes, thereby reducing the ORR/OER overpotentials. Notably, the W4C3O2 MXene shows the superior catalytic performance, characterized by high conductivity and ultra-low ORR, OER, and total overpotentials (0.25, 0.38, and 0.63 V). This study bridges the gap left by WC MXenes as cathode catalysts for Li-O2 batteries and enriches the research of the MXene family in Li-O2 batteries.

2. Details of the Calculation

In this study, all spin-unrestricted DFT calculations were performed using the DMol3 module [28,29]. To describe electron exchange correlation, we employed the Perdew–Burke–Ernzerhof (PBE) functional based on the generalized gradient approximation (GGA) [30], which has been widely used in research on WC Mxenes [31,32]. However, it should be noted that GGA function fails to accurately describe the long-range van der Waals (vdW) interaction. Therefore, the Grimme’s dispersion correction (DFT-D3) method was incorporated into our study to account for vdW interactions [4]. The Grimme’s correction method consistently describes all chemically relevant elements within a periodic system and exhibits equal efficacy for both molecules and solids. Furthermore, it achieves a CCSD (T) accuracy, with an error of within 10% [33]. Therefore, the Grimme’s method provides a reliable description of the surface chemistry of MXene systems including the WC family [27,34].
The classical Monkhorst–Pack scheme is employed to generate K-points [35]. In the convergence test, a 4 × 4 × 1 Monkhorst–Pack grid was utilized for K-point sampling [36], and the self-consistent convergence criterion (SCF tolerance) was set to 1 × 10−5 Ha. To eliminate interlayer interactions, a vacuum layer with a thickness of 20 Å was introduced along the z direction of the WC MXene cell. During geometric optimization, equilibrium geometry was achieved when energy, force, and displacement fall below the thresholds of 2 × 10−5 Ha, 4 × 10−3 Ha Å−1, and 5 × 10−3 Å, respectively [37]. The smearing value of 0.005 Ha expedites the convergence rate of electronic structure optimization.
The formation energy ( E f ) of Wn+1Cn MXenes was defined as:
E f = E W n + 1 C n n + 1 E W n E C / ( 2 n + 1 )
where E W n + 1 C n is the total energy of Wn+1Cn, E W is the chemical potential of one W atom in the bulk W, and E C is the chemical potential of one C atom in graphene.
The adsorption energy ( E a d s ) was defined as:
E a d s = E t o t a l E s u b s t r a t e E a d s o r b a t e
where E t o t a l represents the total energy of the adsorption system, E s u b s t r a t e is the total energy of the substrate, and E a d s o r b a t e is the total energy of the adsorbate.
During the charge and discharge processes, the free energy change (ΔG) of the intermediates at each step can be described as follows:
Δ G = E E 0 Δ n L i μ L i e U + Δ n O 2 μ O 2
where E represents the total energy of the adsorption system at a specific reaction step, E 0 is the total energy of the adsorption system at the initial reaction step, and the n L i and n O 2 are the numbers of Li+ and O2, respectively. The chemical potential ( μ L i ) is defined as the energy of a Li atom in the bulk phase, and the chemical potential ( μ O 2 ) is defined as the energy of an isolated O2 molecule in the gas phase. It has been observed that there is a computational error when calculating the binding energy of O2 molecules by using the DFT algorithm [38]. In this study, we determine the total energy of an O2 molecule in gas phase by combining experimental O2-binding energy (5.12 eV [39]) with DFT-calculated O atom energy. This is a widely adopted approach in the previous literature [40]. Moreover, any over-binding errors for oxygen molecules are expected to be compensated for free energy profiles of ORR and OER. Therefore, we can accurately determine the qualitative characteristics of free energy profiles of ORR and OER. The term -eU was included to describe changes in electron potential at potential U. Additionally, since formation and decomposition of LixO2 occur under low temperatures (T) and pressures (P), effects such as entropy (-TS) and volume (PV) are disregarded, which is frequently used in studies in Li-O2 batteries [41,42]. In this work, we focused on examining the thermodynamic process of the elementary steps of ORR and OER. We assumed that any barriers between these steps are sufficiently small to not impose additional dynamic constraints on starting current at a measurable level. This approach has been widely employed in investigating Li-O2 batteries [43,44].
The ORR ( η O R R ), OER ( η O E R ), and total ( η T O T ) potentials are defined as η O R R = U 0 U D C , η O E R = U C U 0 , and η T O T = η O R R + η O E R , respectively. In this definition, U D C represents the highest discharge potential that drives the energy downhill for all ORR steps, U C represents the lowest charge potential that drives the energy downhill for all OER steps, and U 0 denotes the equilibrium potential ( G 0 ) that facilitates the spontaneous occurrence of ORR/OER [45,46].
The d-band center ( ε d ) and p-band center ( ε p ) are calculated using formulas:
ε d = E ρ d E d E ρ d E d E
and
ε p = E ρ p E d E ρ p E d E
where ρ d E and ρ p E denote the densities of d-states and p-states at the energy level E, respectively.

3. Results and Discussion

3.1. Structural Properties of WC MXenes

The crystal structures of optimized W2C, W3C2 and W4C3 are depicted in Figure 1. W2C MXene exhibits a hexagonal structure similar to hexagonal MoS2 with two surface W layers and an intermediate C layer (Figure 1a). By altering the stacking sequence of the W and C layers according to ABA stacking, the W-C-W sandwich structure of W2C MXene serves as a foundation for constructing thicker MXenes such as W3C2 and W4C3 (Figure 1). Consequently, W3C2 MXene consists of three W layers and two C layers, while W4C3 MXene contains four W layers and three C layers. The lattice constants of W2C MXenes are calculated to be a = b = 2.83 Å, and the W-C bond length is found to be 2.126 Å, which is in excellent agreement with the previously reported results (Lattice constant a = b = 2.84 Å, and W-C bond length 2.130 Å) [27,34]. In addition, the formation energy of W2C, W3C2, and W4C3 MXenes is calculated to be −3.48, −3.35, and −3.33 eV (see Table S1), respectively, indicating their strong thermodynamic stability [25]. All of these indicate the reliability of the WC-MXene models and calculation methods.
The surface modification is commonly employed to introduce functional groups, such as -O, -F, and -OH, onto the MXene surfaces to improve the interaction between MXenes and adsorbents [47]. Previous studies have demonstrated that oxygen from air can replace -F and -OH groups, leading to more stable O-terminated MXenes. This suggests that WC MXenes are highly susceptible to being covered by -O groups [48,49,50]. Therefore, we selected W2C, W3C2, and W4C3 MXenes with full coverage of -O groups (i.e., W2CO2, W3C2O2 and W4C3O2, see Figure 1d–f) as probes for surface-functionalized WC MXenes.
The projected state density (PDOS) of the pristine Wn+1Cn and O-terminated Wn+1CnO2 (n = 1, 2, 3) MXenes were plotted to investigate the electronic properties of WC MXenes, as depicted in Figure 2. It can be observed that the total density of states (TDOSs) of both Wn+1Cn and Wn+1CnO2 intersect with the Fermi level, indicating the high electrical conductivity of Wn+1Cn and Wn+1CnO2. With the increasing number of atomic layers of WC MXenes, the intensity of TDOS at the Fermi level increase gradually, suggesting the progressively enhanced conductivity of Wn+1Cn and Wn+1CnO2 MXenes. Furthermore, the d-band centers of the surface W atoms for W2C, W3C2, and W4C3 are calculated to be −3.11, −3.73, and −3.90 eV with respect to the Fermi level, respectively, while the p-band centers of the surface O atoms for W2CO2, W3C2O2, and W4C3O2 are calculated to be −4.37, −4.57, and −4.61 eV, respectively. As the atomic layer increases, both the W d band center of Wn+1Cn and the O p band center of Wn+1CnO2 move downwards and away from the Fermi level, suggesting that the binding strength of the MXene surfaces to the LixO2 intermediate in the Li-O2 batteries gradually weakens [51,52]. Therefore, increasing atomic layers of Wn+1Cn/Wn+1CnO2 MXenes is not only beneficial for improving conductivity, but also weakens the adsorption of LixO2 intermediates, thereby preventing their accumulation on the electrode surfaces.
According to Hirshfeld’s charge population analyses (Figure 3), the surface W atom of Wn+1Cn MXenes carries a positive charge of 0.12–0.19 e, while the sublayer C atom has a negative charge of −0.24 e to −0.25 e, suggesting electron transfer from W to C. Therefore, the surface W atom in Wn+1Cn is a positive charge center, showing electrophilicity. When the -O groups are introduced on the surface of Wn+1Cn, the positive charge on the W atom of Wn+1CnO2 increases to 0.30 e~0.34 e, and the -O group has a negative charge of −0.20 e, which indicates that strong electron transfer from W to O occurs in Wn+1CnO2. Thus, the surface O atom in Wn+1CnO2 forms a negative charge center, which is nucleophilic. This situation can be further confirmed with the differential electron density maps. As shown in Figure 4, an electron depletion region (yellow) is observed on the surface W atoms of Wn+1Cn, while an electron accumulation region (blue) is located on the surface O atoms of Wn+1CnO2, due to the different electronegativity between non-metallic oxygen and metallic W. Consequently, oxygen functionalization transforms WC MXenes from an electrophilic surface of Wn+1Cn to a nucleophilic surface of Wn+1CnO2, thereby regulating the adsorption and activation of intermediates on the MXene surfaces.
The adsorption of the LixO2 (x = 1, 2, and 4) intermediates was examined on Wn+1Cn and Wn+1CnO2. As shown in Figure S1, the LixO2 intermediates exhibit similar adsorption configurations on the Wn+1Cn surfaces, where the surface W atom binds to the O atom in LixO2. This is because the electrophilic W atoms on the Wn+1Cn surface prefer to bind with the negatively charged O atoms in LixO2. The strong interaction between the oxygen atom and the surface W atom promotes the activation of the O-O bond in the adsorbed LixO2*. Consequently, the O-O bonds in the adsorbed LixO2* on Wn+1Cn are apparently longer than the Li-O bonds (see Table S3), which hinders the deintercalation of lithium during the charge–discharge processes. Alternatively, the surface O atoms of Wn+1CnO2 directly adsorb the Li atoms of LixO2 (see Figure S2), because the nucleophilic surface O atoms tend to bind with positively charged Li atoms. This results in a shortened O-O bond and an elongated Li-O bond after LixO2 adsorbs on the surface of Wn+1CnO2 (Table S4), thereby facilitating the deintercalation of lithium during the charge–discharge processes. Thus, oxygen functionalization can effectively improve the lithium deintercalation process on the WC MXene surfaces.
The adsorption energies of the LixO2 (x = 1, 2, and 4) intermediates on Wn+1Cn and Wn+1CnO2 are listed in Table S2. Compared with the W surface of Wn+1Cn, the O surface of Wn+1CnO2 exhibits weaker adsorption towards the LixO2 intermediates. This is in good accordance with previous theoretical reports, where the oxide layer behaves as a passivation layer on the TiC(111), ZrC(111), α-MoC(001), and Mo2C(001) systems upon Li2O2 adsorption [52]. Furthermore, with an increasing number of atomic layers, the adsorption energy of the LixO2 intermediates on both Wn+1Cn and Wn+1CnO2 MXenes is gradually weakened. The situation can be attributed to the downward shift of the d-band centers of surface W atoms in Wn+1Cn and the p-band centers of surface O atoms in Wn+1CnO2 with the increasing atomic layers. As shown in Figure 2, the d-band center of surface W atoms is shifted from −3.11 eV in W2C to −3.90 eV in W4C3, while the p-band center of surface O atoms is shifted from −4.37 eV in W2CO2 to −4.61 eV in W4C3O2. The downward shift of the band center increases the electron filling on the anti-bonding states between MXenes and the adsorbate, resulting in a weakened binding interaction. Hence, both oxygen functionalization and increasing atomic layers can weaken the interaction between WC MXenes and the LixO2 intermediates, avoiding their accumulation caused by their excessive adsorption.

3.2. Evaluation of Catalytic Activity

Based on the reported experimental and theoretical results [2,3,47], we investigated three surface reaction steps (Equations (2)–(4)) to simulate the ORR/OER process on WC MXenes. During the discharge process, the O2* species on the surface of WC MXene cathodes undergoes initial metallization with ( L i + + e ) to form adsorbed LiO2*. Subsequently, LiO2* undergoes a second metallization with ( L i + + e ) to generate Li2O2*. Finally, Li2O2* further reacts with ( L i + + e ) to yield the final product (Li2O)2*. The OER at the cathode of a Li-O2 battery during charging is essentially the reverse process of the aforementioned ORR. In the OER process, the ultimate adsorption product (Li2O)2* gradually decomposes into O2*, which then dissociates from the surface of WC MXenes.
O 2   * + L i + + e L i O 2   *
L i O 2   * + L i + + e L i 2 O 2   *
L i 2 O 2   * + 2 L i + + 2 e L i 2 O 2   *
Subsequently, we constructed the free energy diagram to illustrate the ORR/OER process on Wn+1Cn and Wn+1CnO2 (Figure 5). The purple arrows in the graph represent the discharge-oriented ORR process from left to right, and the orange arrows indicate the charging-driven OER process from right to left. The nucleation of (Li2O)2 on WN MXenes follows three steps: O2* → LiO2*→ Li2O2*→ (Li2O)2*. Notably, at open circuit voltage (U = 0 V), both Wn+1Cn and Wn+1CnO2 (the black path in Figure 5) exhibit a downhill trend in their free energy profiles for all three metallization steps of ORR, suggesting the spontaneous nucleation of (Li2O)2 on the surface of WC MXenes. Conversely, during the reverse OER process, the decomposition of (Li2O)2 is endothermic on both Wn+1Cn and Wn+1CnO2 MXenes. Furthermore, the whole free energy change (ΔG(O2*→(Li2O)2*)) of the O2*→(Li2O)2* process decreases with the increasing atomic layers of Wn+1Cn and Wn+1CnO2. This is because the WC MXenes exhibit a weakened affinity towards LixO2 intermediates as the number of atomic layers increases. These findings suggest that as the number of atomic layers increases, it becomes easier for (Li2O)2 products to undergo decomposition on the WC MXene surfaces. Moreover, compared to the pristine Wn+1Cn, ΔG(O2→(Li2O)2*) for the oxygen-functionalized Wn+1CnO2 is significantly reduced, which is further beneficial for the (Li2O)2 decomposition. Therefore, both increasing atomic layers and oxygen functionalization can promote the de-lithiation of (Li2O)2 products, thereby accelerating the electrochemical process during charging.
In Figure 5, UDc represents the maximum discharge potential driving the energy of all ORR steps to exhibit a downward trend along the red path from left to right, while UC represents the minimum charging potential driving the energy of all OER steps to decrease along the green path from right to left. The equilibrium potential U0 applied in the blue path facilitates achieving equilibrium in the electrochemical ORR/OER process. With an increase in atomic layer number, the ΔG(O2*→(Li2O)2*) of Wn+1Cn MXenes gradually decreases, and therefore the required U0 decreases as well. The values of U0 are calculated to be 4.70, 4.12, and 3.70 V for W2C, W3C2, and W4C3, respectively (Figure 5a–c). After surface oxygen functionalization, the ΔG(O2*→(Li2O)2*) of Wn+1CnO2 further decreases significantly, and thus the corresponding U0 for W2CO2, W3C2O2, and W4C3O2 is decreased to be 3.60, 3.36, and 3.08 V, respectively (Figure 5d–f). Consequently, the W4C3O2 MXene has a minimum U0 among Wn+1Cn and Wn+1CnO2 MXenes.
When U0 is applied to the electrochemical processes on Wn+1Cn and Wn+1CnO2, the formation steps of Li2O2* and LiO2* along the ORR pathway are still downhill in the free energy profiles (the blue path in Figure 5). However, the last (Li2O)2* formation step shows an upward trend, suggesting it forms the rate-determining step (RDS) of the ORR pathway on both Wn+1Cn and Wn+1CnO2. Compared to the downhill step of the (Li2O)2* decomposition, the decomposition of Li2O2* and LiO2* is uphill along the OER pathway in the free energy profiles. Furthermore, the decomposition of LiO2* requires a higher energy input than that of Li2O2*, suggesting that it serves as the RDS of OER on Wn+1Cn and Wn+1CnO2.
The overpotential η is defined as the minimum U U 0 that makes all the electrochemical steps downhill in free energy and serves as a crucial indicator for assessing the catalytic performance of a catalyst [51]. The smaller the value of η , the lower the actual voltage required to achieve a target current density, resulting in reduced energy consumption and enhanced catalytic activity [45,53]. In this study, we calculated the ORR ( η O R R ), OER ( η O E R ), and total ( η T O T ) overpotentials by η O R R = U 0 U D c , η O E R = U C U 0 , and η T O T = η O R R + η O E R , respectively. Detailed data on overpotentials are presented in Figure 6. The overpotentials ( η O E R / η O R R / η T O T ) of the pristine Wn+1Cn follow the order: W4C3 (0.78 V/0.51 V/1.29 V) < W3C2 (0.93 V/0.60 V/1.53 V) < W2C (1.74 V/1.03 V/2.77 V), suggesting that the Wn+1Cn MXenes show a decrease trend in overpotentials with an increasing number of atomic layers. After the WC surfaces are covered with O groups, there is a significant decrease in overpotentials. The values of η O E R , η O R R , and η T O T are decreased in the order of W4C3O2 (0.38 V/0.25 V/0.63 V) < W3C2O2 (0.45 V/0.39 V/0.84 V) < W2CO2 (0.54 V/0.48 V/1.02 V). Moreover, the values of η O E R for Wn+1Cn and Wn+1CnO2 MXenes are higher than those of η O R R , indicating the slower kinetics of the OER during the charging process, which may lead to poor cyclic stability [54]. This is attributed to the strong adsorption of the LixO2 produced during the discharge process, which makes it difficult to reversibly decompose, resulting in continuous accumulation [55,56]. Among all considered WC MXenes, W4C3O2 exhibits the lowest OER, ORR, and total overpotentials (0.38, 0.25, and 0.63 V). Furthermore, it is worth noting that the overpotentials ( η O E R , η O R R , and η T O T ) of W4C3O2 are lower than those of Nb2CO2 MXene (0.81, 0.50, and 1.31 V) [57], Se@NiO/CC (0.32, 0.36, and 0.68 V) [58], SASe–Ti3C2 (0.59, 0.29, and 0.88 V) [59], CuCo2S4 (0.35, 0.30, and 0.65 V) [60], NiSA-Co3O4 (1.09, 0.21, and 1.30 V) [12], CoS2 (0.89, 0.47, and 1.36 V) [61], and other recently reported two-dimensional materials [45,62,63]. These suggest that the W4C3O2 MXene shows excellent catalytic performance as a cathode catalyst for Li-O2 batteries.
To further investigate the relationship between the adsorption property of LixO2 and the overpotentials, we plotted the correlation between the adsorption energy (Eads) of RDS intermediates and η O R R / η O E R in the ORR/OER process for WC MXenes (Figure 7). In the ORR process, the reduction of Li2O2* to (Li2O)2* serves as the RDS on both Wn+1Cn and Wn+1CnO2 MXenes, and thus the adsorption energy of Li2O2* (Eads(Li2O2*)) plays a key role in η O R R . As depicted in Figure 7a, there is a linear correlation between Eads(Li2O2*) and η O R R for WC MXenes. The values of η O R R decrease with the decrease of the adsorption energy of Li2O2* on Wn+1Cn and Wn+1CnO2 MXenes. Reducing the adsorption strength of Li2O2* is beneficial for further metallization to produce (Li2O)2*, resulting in a reduced value of the corresponding η O R R for WC MXenes. In the OER process, the decomposition of LiO2* into O 2   * acts as the RDS on WC MXenes. Thus, the adsorption energy of LiO2* (Eads(LiO2*)) is important in determining η O E R . As shown in Figure 7b, Eads(LiO2*) and η O E R exhibit a linear relationship for Wn+1Cn and Wn+1CnO2 MXenes. A weaker Eads(LiO2*) can promote the decomposition of LiO2*, leading to a lower value of the corresponding η O E R . These findings demonstrate that the reduced adsorption energy of the RDS intermediates (LiO2* and Li2O2*) has a positive effect on reducing overpotentials. By lowering the energy band center, oxygen functionalization and increasing the atomic layer can effectively reduce the adsorption strength of the RDS intermediates (Li2O2* and LiO2*), thereby reducing the ORR and OER overpotentials.

4. Conclusions

In this work, the models of pristine Wn+1Cn and oxygen-functionalized Wn+1CnO2 MXenes were constructed, and their catalytic performance as cathodes for Li-O2 batteries was evaluated using first principles calculations. The TDOS analyses at the Fermi level confirm the excellent electrical conductivity of both Wn+1Cn and Wn+1CnO2, which is further enhanced with increasing atomic layers. The oxygen functionalization alters the electronic properties of WC MXenes from the electrophilic W surface of Wn+1Cn to the nucleophilic O surface of Wn+1CnO2, which facilitates the Li-O bond activation and thus promotes Li deintercalation during the charge–discharge process. Compared to the pristine Wn+1Cn, the oxygen-functionalized Wn+1CnO2 has a significantly reduced adsorption energy towards LixO2, resulting in lower overpotentials of η O E R , η O R R , a n d   η T O T . As the number of atomic layers in WC MXenes increases, the adsorption energy of LixO2 is further decreased, leading to a reduction in η O E R , η O R R , and η T O T . The O-terminated W4C3O2 MXene shows superior electrical conductivity and remarkably low overpotentials (0.38 V for η O E R , 0.25 V for η O R R , and 0.63 V for η T O T ), highlighting its huge potential as a cathode catalyst for Li-O2 batteries. The study indicates that the WC MXenes can serve as cathode materials for Li-O2 batteries, and W4C3O2 is identified as a high-performance cathode catalyst material. This finding is of great importance for the design and manufacture of cathode catalysts used in Li-O2 batteries.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano14080666/s1, Figure S1: the top and side views of the adsorption configuration of LiO2 on WC MXenes; Figure S2: the top and side views of the adsorption configuration of Li2O2 on WC MXenes; Table S1: formation energy of W2C, W3C2, and W4C3; Table S2: the adsorption energy of LixO2 on Wn+1Cn and Wn+1CnO2; Table S3: the average length of the Li-O and O-O bonds in LiO2/Li2O2 and the adsorbed distance of LiO2/Li2O2 to Wn+1Cn; Table S4: the average length of the Li-O and O-O bonds in LiO2/Li2O2 and the adsorbed distance of LiO2/Li2O2 to Wn+1CnO2; Table S5: UDc, U0, UC, η O R R , η O E R , and η T O T for Wn+1Cn and Wn+1CnO2.

Author Contributions

L.Z. (Liwei Zhu)—Investigation, Writing; J.W.—Data curation; J.L.—Validation; R.W.—Visualization; M.L.—Formal analysis; T.W.—Investigation; Y.Z.—Validation, Visualization; J.X.—Funding acquisition, Resources, Supervision; L.Z. (Lianming Zhao)—Supervision, Funding acquisition, Project administration, Writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Shandong Provincial Natural Science Foundation (No. ZR2022MB094).

Data Availability Statement

Data will be made available on request.

Conflicts of Interest

The authors declare that they have no known competing financial interest or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Mu, X.; Xia, C.; Gao, B.; Guo, S.; Zhang, X.; He, J.; Wang, Y.; Dong, H.; He, P.; Zhou, H. Two-Dimensional Mo-Based Compounds for the Li-O2 Batteries: Catalytic Performance and Electronic Structure Studies. Energy Storage Mater. 2021, 41, 650–655. [Google Scholar] [CrossRef]
  2. Xia, Q.; Li, D.; Zhao, L.; Wang, J.; Long, Y.; Han, X.; Zhou, Z.; Liu, Y.; Zhang, Y.; Li, Y. Recent Advances in Heterostructured Cathodic Electrocatalysts for Non-Aqueous Li–O2 Batteries. Chem. Sci. 2022, 13, 2841–2856. [Google Scholar] [CrossRef] [PubMed]
  3. Du, D.; Zhu, Z.; Chan, K.-Y.; Li, F.; Chen, J. Photoelectrochemistry of Oxygen in Rechargeable Li–O2 Batteries. Chem. Soc. Rev. 2022, 51, 1846–1860. [Google Scholar] [CrossRef] [PubMed]
  4. Zhang, Y.; Zhang, S.; Li, H.; Lin, Y.; Yuan, M.; Nan, C.; Chen, C. Tunable Oxygen Vacancies of Cobalt Oxides in Lithium–Oxygen Batteries: Morphology Control of Discharge Product. Nano Lett. 2023, 23, 9119–9125. [Google Scholar] [CrossRef]
  5. Sun, Z.; Lin, X.; Wang, C.; Hu, A.; Hou, Q.; Tan, Y.; Dou, W.; Yuan, R.; Zheng, M.; Dong, Q. High-Performance Lithium–Oxygen Batteries Using a Urea-Based Electrolyte with Kinetically Favorable One-Electron Li2O2 Oxidation Pathways. Angew. Chem. 2022, 134, E202207570. [Google Scholar] [CrossRef]
  6. Xing, S.; Zhang, Z.; Dou, Y.; Li, M.; Wu, J.; Zhang, Z.; Zhou, Z. An Efficient Multifunctional Soluble Catalyst for Li-O2 Batteries. CCS Chem. 2023. [Google Scholar] [CrossRef]
  7. Zhao, W.; Wang, J.; Yin, R.; Li, B.; Huang, X.; Zhao, L.; Qian, L. Single-Atom Pt Supported on Holey Ultrathin G-C3N4 Nanosheets as Efficient Catalyst for Li-O2 Batteries. J. Colloid Interface Sci. 2020, 564, 28–36. [Google Scholar] [CrossRef]
  8. Zhou, Y.; Gu, Q.; Yin, K.; Li, Y.; Tao, L.; Tan, H.; Yang, Y.; Guo, S. Engineering Eg Orbital Occupancy of Pt with Au Alloying Enables Reversible Li− O2 Batteries. Angew. Chem. Int. Ed. 2022, 61, E202201416. [Google Scholar] [CrossRef] [PubMed]
  9. Cui, X.; Luo, Y.; Zhou, Y.; Dong, W.; Chen, W. Application of Functionalized Graphene in Li–O2 Batteries. Nanotechnology 2021, 32, 132003. [Google Scholar] [CrossRef]
  10. Dang, C.; Feng, P.; He, S.; Zhao, L.; Shan, A.; Li, M.; Kong, L.; Gao, L. Nicop Based Carbon Nanotube Heterostructure for Improved Oxygen Redox Reaction Kinetics in Li-O2 Batteries. Electrochim. Acta 2023, 462, 142771. [Google Scholar] [CrossRef]
  11. Li, Y.; Qin, J.; Ding, Y.; Ma, J.; Das, P.; Liu, H.; Wu, Z.-S.; Bao, X. Two-Dimensional Mn3O4 Nanosheets with Dominant (101) Crystal Planes on Graphene as Efficient Oxygen Catalysts for Ultrahigh Capacity and Long-Life Li–O2 Batteries. ACS Catal. 2022, 12, 12765–12773. [Google Scholar] [CrossRef]
  12. Lian, Z.; Lu, Y.; Ma, S.; Li, Z.; Liu, Q. Metal Atom-Doped Co3O4 Nanosheets for Li-O2 Battery Catalyst: Study on the Difference of Catalytic Activity. Chem. Eng. J. 2022, 445, 136852. [Google Scholar] [CrossRef]
  13. Zhang, L.; Zhao, C.; Kong, X.; Yu, S.; Zhang, D.; Liu, W. Construction of Co-NC@ Mo2C Hetero-Interfaces for Improving the Performance of Li-O2 Batteries. Electrochim. Acta 2023, 446, 142096. [Google Scholar] [CrossRef]
  14. Meng, X.; Liao, K.; Dai, J.; Zou, X.; She, S.; Zhou, W.; Ye, F.; Shao, Z. Ultralong Cycle Life Li–O2 Battery Enabled by a MOF-Derived Ruthenium–Carbon Composite Catalyst with a Durable Regenerative Surface. ACS Appl. Mater. Interfaces 2019, 11, 20091–20097. [Google Scholar] [CrossRef] [PubMed]
  15. Li, N.; Chang, Z.; Zhong, M.; Fu, Z.-X.; Luo, J.; Zhao, Y.-F.; Li, G.-B.; Bu, X.-H. Functionalizing MOF with Redox-Active Tetrazine Moiety for Improving the Performance as Cathode of Li–O2 Batteries. CCS Chem. 2021, 3, 1297–1305. [Google Scholar] [CrossRef]
  16. Zheng, X.; Yuan, M.; Huang, X.; Li, H.; Sun, G. In Situ Decoration of Cop/Ti3C2Tx Composite as Efficient Electrocatalyst for Li-Oxygen Battery. Chin. Chem. Lett. 2023, 34, 107152. [Google Scholar] [CrossRef]
  17. Teng, K.; Tang, W.; Qi, R.; Li, B.; Deng, Y.; Zhou, M.; Wu, M.; Zhang, J.; Liu, R.; Zhang, L. Nitrogen-Deficient G-C3N4 Compounded with NiCo2S4 (NiCo2S4@ ND-CN) as a Bifunctional Electrocatalyst for Boosting the Activity of Li-O2 Batteries. Catal. Today 2023, 409, 23–30. [Google Scholar] [CrossRef]
  18. Xu, H.; Zheng, R.; Du, D.; Ren, L.; Li, R.; Wen, X.; Zhao, C.; Shu, C. V2C Mxene Enriched with-O Termination as High-Efficiency Electrocatalyst for Lithium-Oxygen Battery. Appl. Mater. Today 2022, 27, 101464. [Google Scholar] [CrossRef]
  19. Tariq, H.A.; Nisar, U.; Abraham, J.J.; Ahmad, Z.; Alqaradawi, S.; Kahraman, R.; Shakoor, R. Tio2 Encrusted Mxene as a High-Performance Anode Material for Li-Ion Batteries. Appl. Surf. Sci. 2022, 583, 152441. [Google Scholar] [CrossRef]
  20. Zhang, P.; Li, J.; Yang, D.; Soomro, R.A.; Xu, B. Flexible Carbon Dots-Intercalated Mxene Film Electrode with Outstanding Volumetric Performance for Supercapacitors. Adv. Funct. Mater. 2023, 33, 2209918. [Google Scholar] [CrossRef]
  21. Tu, W.; Chen, K.; Zhu, L.; Zai, H.E.B.; Ke, X.; Chen, C.; Sui, M.; Chen, Q.; Li, Y. Tungsten-Doping-Induced Surface Reconstruction of Porous Ternary Pt-Based Alloy Electrocatalyst for Oxygen Reduction. Adv. Funct. Mater. 2019, 29, 1807070. [Google Scholar] [CrossRef]
  22. Du, Y.; Chen, W.; Zhou, L.; Hu, R.; Wang, S.; Li, X.; Xie, Y.; Yang, L.; Liu, Y.; Liu, Z. One-Step, in Situ Formation of WN-W2C Heterojunctions Implanted on N Doped Carbon Nanorods as Efficient Oxygen Reduction Catalyst for Metal-Air Battery. Nano Res. 2023, 16, 8773–8781. [Google Scholar] [CrossRef]
  23. Chen, J.; Ren, B.; Cui, H.; Wang, C. Constructing Pure Phase Tungsten-Based Bimetallic Carbide Nanosheet as An Efficient Bifunctional Electrocatalyst for Overall Water Splitting. Small 2020, 16, 1907556. [Google Scholar] [CrossRef] [PubMed]
  24. Guo, J.; Mao, Z.; Yan, X.; Su, R.; Guan, P.; Xu, B.; Zhang, X.; Qin, G.; Pennycook, S.J. Ultrasmall Tungsten Carbide Catalysts Stabilized in Graphitic Layers for High-Performance Oxygen Reduction Reaction. Nano Energy 2016, 28, 261–268. [Google Scholar] [CrossRef]
  25. GarcÍA-Romeral, N.; Morales-GarcÍA, Á.; Viñes, F.; De PR Moreira, I.; Illas, F. The Nature of The Electronic Ground State of M2C (M= Ti, V, Cr, Zr, Nb, Mo, Hf, Ta, And W) Mxenes. Phys. Chem. Chem. Phys. 2023, 25, 31153–31164. [Google Scholar] [CrossRef]
  26. Qin, T.; Wang, Z.; Wang, Y.; Besenbacher, F.; Otyepka, M.; Dong, M. Recent Progress in Emerging Two-Dimensional Transition Metal Carbides. Nano-Micro Lett. 2021, 13, 183. [Google Scholar] [CrossRef] [PubMed]
  27. Wang, J.; Bai, L.; Yao, C.; Niu, L. A DFT Computational Prediction of 2H Phase W2C Monolayer and the Effect of O Functional Groups. Phys. Lett. A 2022, 424, 127842. [Google Scholar] [CrossRef]
  28. Chen, X.; Liu, Q.; Zhang, H.; Zhao, X. Exploring High-Efficiency Electrocatalysts of Metal-Doped Two-Dimensional C4N for Oxygen Reduction, Oxygen Evolution, and Hydrogen Evolution Reactions by First-Principles Screening. Phys. Chem. Chem. Phys. 2022, 24, 26061–26069. [Google Scholar] [CrossRef]
  29. Hu, H.; Zhang, P.; Xiao, B.-B.; Mi, J.-L. Theoretical Study of P-Block Metal–Nitrogen–Carbon Single-Atom Catalysts for the Oxygen Reduction Reaction. Catal. Sci. Technol. 2022, 12, 6751–6760. [Google Scholar] [CrossRef]
  30. Jing, T.; Liang, D.; Deng, M.; Cai, S.; Qi, X. Density Functional Theory Studies of Heteroatom-Doped Graphene-Like Gan Monolayers as Electrocatalysts for Oxygen Evolution and Reduction. ACS Appl. Nano Mater. 2021, 4, 7125–7133. [Google Scholar] [CrossRef]
  31. Gouveia, J.D.; Morales-Garcia, A.; Vines, F.; Gomes, J.R.; Illas, F. Facile Heterogeneously Catalyzed Nitrogen Fixation by Mxenes. ACS Catal. 2020, 10, 5049–5056. [Google Scholar] [CrossRef]
  32. Glechner, T.; Tomastik, C.; Badisch, E.; Polcik, P.; Riedl, H. Influence of WC/C Target Composition and Bias Potential on the Structure-Mechanical Properties of Non-Reactively Sputtered WC Coatings. Surf. Coat. Technol. 2022, 432, 128036. [Google Scholar] [CrossRef]
  33. 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] [CrossRef] [PubMed]
  34. Gouveia, J.D.; Viñes, F.; Illas, F.; Gomes, J.R. Mxenes Atomic Layer Stacking Phase Transitions and their Chemical Activity Consequences. Phys. Rev. Mater. 2020, 4, 054003. [Google Scholar] [CrossRef]
  35. Monkhorst, H.J.; Pack, J.D. Special Points for Brillouin-Zone Integrations. Phys. Rev. B 1976, 13, 5188. [Google Scholar] [CrossRef]
  36. Xiao, B.; Yang, L.; Liu, H.; Jiang, X.; Aleksandr, B.; Song, E.; Jiang, Q. Designing Fluorographene with Fen4 and Con4 Moieties for Oxygen Electrode Reaction: A Density Functional Theory Study. Appl. Surf. Sci. 2021, 537, 147846. [Google Scholar] [CrossRef]
  37. Halgren, T.A.; Lipscomb, W.N. The Synchronous-Transit Method for Determining Reaction Pathways and Locating Molecular Transition States. Chem. Phys. Lett. 1977, 49, 225–232. [Google Scholar] [CrossRef]
  38. Lin, L.; Yang, X.; Shi, P.; Yan, L.; Xie, K.; Deng, C.; Chen, Z. Probing the Origin of Transition Metal Carbide VC for Oxygen Reduction Reaction: A DFT Study. Surf. Interfaces 2023, 40, 103100. [Google Scholar] [CrossRef]
  39. Radin, M.D.; Rodriguez, J.F.; Tian, F.; Siegel, D.J. Lithium Peroxide Surfaces are Metallic, while Lithium Oxide Surfaces are Not. J. Am. Chem. Soc. 2012, 134, 1093–1103. [Google Scholar] [CrossRef]
  40. Yang, Y.; Wang, Y.; Yao, M.; Wang, X.; Huang, H. First-Principles Study of Rocksalt Early Transition-Metal Carbides as Potential Catalysts for Li–O2 Batteries. Phys. Chem. Chem. Phys. 2018, 20, 30231–30238. [Google Scholar] [CrossRef]
  41. Zhao, L.; Feng, J.; Abbas, A.; Wang, C.; Wang, H. MOF-Derived Mn2O3 Nanocage with Oxygen Vacancies as Efficient Cathode Catalysts for Li–O2 Batteries. Small 2023, 19, 2302953. [Google Scholar] [CrossRef] [PubMed]
  42. Li, J.; Shu, C.; Liu, C.; Chen, X.; Hu, A.; Long, J. Rationalizing the Effect of Oxygen Vacancy on Oxygen Electrocatalysis in Li–O2 Battery. Small 2020, 16, 2001812. [Google Scholar] [CrossRef] [PubMed]
  43. Hummelshøj, J.; Luntz, A.C.; Nørskov, J. Theoretical Evidence for Low Kinetic Overpotentials in Li-O2 Electrochemistry. J. Chem. Phys. 2013, 138, 034703. [Google Scholar] [CrossRef] [PubMed]
  44. Yang, Y.; Xue, X.; Qin, Y.; Wang, X.; Yao, M.; Qin, Z.; Huang, H. Oxygen Evolution Reaction on Pristine and Oxidized Tic (100) Surface in Li–O2 Battery. J. Phys. Chem. C 2018, 122, 12665–12672. [Google Scholar] [CrossRef]
  45. Ding, S.; Wu, L.; Yuan, X. Regulating D-Orbital Electronic Configuration of Ni-Based Chalcogenides to Enhance the Oxygen Electrode Reactions in Li-O2 Batteries. Chem. Eng. J. 2023, 478, 147473. [Google Scholar] [CrossRef]
  46. Lian, Z.; Lu, Y.; Zhao, S.; Li, Z.; Liu, Q. Engineering the Electronic Interaction Between Atomically Dispersed Fe and Ruo2 Attaining High Catalytic Activity and Durability Catalyst for Li-O2 Battery. Adv. Sci. 2023, 10, 2205975. [Google Scholar] [CrossRef] [PubMed]
  47. Zheng, X.; Yuan, M.; Zhao, Y.; Li, Z.; Shi, K.; Li, H.; Sun, G. Status and Prospects of Mxene-Based Lithium–Oxygen Batteries: Theoretical Prediction and Experimental Modulation. Adv. Energy Mater. 2023, 13, 2204019. [Google Scholar] [CrossRef]
  48. Bashir, T.; Ismail, S.A.; Wang, J.; Zhu, W.; Zhao, J.; Gao, L. Mxene Terminating Groups O,–F Or–OH,–F Or O,–OH,–F, or O,–OH,–Cl? J. Energy Chem. 2023, 76, 90–104. [Google Scholar] [CrossRef]
  49. Wang, J.; Xu, J.; Li, B.; Lin, M.; Wang, T.; Zhen, Y.; Huang, Z.; Xing, W.; Zhao, L. Theoretical Study of Catalytic Performance of WN Mxenes as Cathodes for Li-O2 Batteries: Effects of Surface Functionalization and Atomic Layers. Appl. Surf. Sci. 2023, 638, 158027. [Google Scholar] [CrossRef]
  50. Bai, X.; Guan, J. Applications of Mxene-Based Single-Atom Catalysts. Small Struct. 2023, 4, 2200354. [Google Scholar] [CrossRef]
  51. Tian, G.; Xu, H.; Wang, X.; Wen, X.; Zeng, T.; Liu, S.; Fan, F.; Xiang, W.; Shu, C. Optimizing D-Orbital Occupation on Interfacial Transition Metal Sites via Heterogeneous Interface Engineering to Accelerate Oxygen Electrode Reaction Kinetics in Lithium-Oxygen Batteries. Nano Energy 2023, 117, 108863. [Google Scholar] [CrossRef]
  52. Tereshchuk, P.; Golodnitsky, D.; Natan, A. Trends in the Adsorption of Oxygen and Li2O2 on Transition-Metal Carbide Surfaces: A Theoretical Study. J. Phys. Chem. C 2020, 124, 7716–7724. [Google Scholar] [CrossRef]
  53. Jiang, Z.; Wen, B.; Huang, Y.; Guo, Y.; Wang, Y.; Li, F. New Reaction Pathway of Superoxide Disproportionation Induced by a Soluble Catalyst in Li-O2 Batteries. Angew. Chem. 2024, 136, E202315314. [Google Scholar] [CrossRef]
  54. Zheng, J.; Zhang, W.; Wang, R.; Wang, J.; Zhai, Y.; Liu, X. Single-Atom Pd-N4 Catalysis for Stable Low-Overpotential Lithium-Oxygen Battery. Small 2023, 19, 2204559. [Google Scholar] [CrossRef] [PubMed]
  55. Zheng, L.J.; Song, L.N.; Wang, X.X.; Liang, S.; Wang, H.F.; Du, X.Y.; Xu, J.J. Intrinsic Stress-Strain in Barium Titanate Piezocatalysts Enabling Lithium− Oxygen Batteries with Low Overpotential and Long Life. Angew. Chem. Int. Ed. 2023, 62, E202311739. [Google Scholar] [CrossRef] [PubMed]
  56. Ding, Y.; Huang, Y.; Li, Y.; Zhang, T.; Wu, Z.S. Regulating Surface Electron Structure of PtNi Nanoalloy Via Boron Doping for High-Current-Density Li-O2 Batteries with Low Overpotential and Long-Life Cyclability. Smartmat 2024, 5, E1150. [Google Scholar] [CrossRef]
  57. Li, G.; Li, N.; Peng, S.; He, B.; Wang, J.; Du, Y.; Zhang, W.; Han, K.; Dang, F. Highly Efficient Nb2C Mxene Cathode Catalyst with Uniform O-Terminated Surface for Lithium–Oxygen Batteries. Adv. Energy Mater. 2021, 11, 2002721. [Google Scholar] [CrossRef]
  58. Wang, L.; Lu, Y.; Xie, M.; Zhao, S.; Li, Z.; Liu, Q. Interfacially Engineered Induced Nickel-Based Heterostructures as Efficient Catalysts for Li-O2 Batteries. Electrochim. Acta 2023, 437, 141476. [Google Scholar] [CrossRef]
  59. Zhao, D.; Wang, P.; Di, H.; Zhang, P.; Hui, X.; Yin, L. Single Semi-Metallic Selenium Atoms on Ti3C2 Mxene Nanosheets as Excellent Cathode for Lithium–Oxygen Batteries. Adv. Funct. Mater. 2021, 31, 2010544. [Google Scholar] [CrossRef]
  60. Tian, G.; Ren, L.; Xu, H.; Zeng, T.; Wang, X.; Wen, X.; Du, D.; Yan, Y.; Liu, S.; Shu, C. Metal Sulfide Heterojunction with Tunable Interfacial Electronic Structure as An Efficient Catalyst for Lithium-Oxygen Batteries. Sci. China Mater. 2023, 66, 1341–1351. [Google Scholar] [CrossRef]
  61. Ding, S.; Wu, L.; Zhang, F.; Yuan, X. Modulating Electronic Structure with Copper Doping to Promote the Electrocatalytic Performance of Cobalt Disulfide in Li–O2 Batteries. Small 2023, 19, 2300602. [Google Scholar] [CrossRef]
  62. Feng, J.; Wang, H.; Guo, L.; Su, W.; Zhao, L.; Li, G.; Chen, T.; Wang, C.; Dang, F. Stacking Surface Derived Catalytic Capability and By-Product Prevention for High Efficient Two Dimensional Bi2Te3 Cathode Catalyst in Li-Oxygen Batteries. Appl. Catal. B Environ. 2022, 318, 121844. [Google Scholar] [CrossRef]
  63. Jiang, H.; Zhao, T.; Shi, L.; Tan, P.; An, L. First-Principles Study of Nitrogen-, Boron-Doped Graphene and Co-Doped Graphene as the Potential Catalysts in Nonaqueous Li–O2 Batteries. J. Phys. Chem. C 2016, 120, 6612–6618. [Google Scholar] [CrossRef]
Figure 1. Top and side views of (a) W2C, (b) W3C2, (c) W4C3, (d) W2CO2, (e) W3C2O2, and (f) W4C3O2. Top, hcp, and fcc indicate possible adsorption sites.
Figure 1. Top and side views of (a) W2C, (b) W3C2, (c) W4C3, (d) W2CO2, (e) W3C2O2, and (f) W4C3O2. Top, hcp, and fcc indicate possible adsorption sites.
Nanomaterials 14 00666 g001
Figure 2. Projected density of states (PDOS) of (ac) Wn+1Cn and (df) Wn+1CnO2. The Fermi level marked by the black dashed line is set as energy zero.
Figure 2. Projected density of states (PDOS) of (ac) Wn+1Cn and (df) Wn+1CnO2. The Fermi level marked by the black dashed line is set as energy zero.
Nanomaterials 14 00666 g002
Figure 3. The Hirshfeld charge (in e) of surface W atoms, sublayer C atoms, and surface O groups of Wn+1Cn and Wn+1CnO2.
Figure 3. The Hirshfeld charge (in e) of surface W atoms, sublayer C atoms, and surface O groups of Wn+1Cn and Wn+1CnO2.
Nanomaterials 14 00666 g003
Figure 4. Differential electron density maps of (a) W2C, (b) W3C2, (c) W4C3, (d) W2CO2, (e) W3C2O2, and (f) W4C3O2.
Figure 4. Differential electron density maps of (a) W2C, (b) W3C2, (c) W4C3, (d) W2CO2, (e) W3C2O2, and (f) W4C3O2.
Nanomaterials 14 00666 g004
Figure 5. Free energy diagram of the ORR/OER process for LixO2 intermediates on (a) W2C, (b) W3C2, (c) W4C3, (d) W2CO2, (e) W3C2O2, and (f) W4C3O2. * indicates that the intermediate is in an adsorbed state.
Figure 5. Free energy diagram of the ORR/OER process for LixO2 intermediates on (a) W2C, (b) W3C2, (c) W4C3, (d) W2CO2, (e) W3C2O2, and (f) W4C3O2. * indicates that the intermediate is in an adsorbed state.
Nanomaterials 14 00666 g005
Figure 6. (a) The ORR (ηORR), OER (ηOER), and total (ηTOT) overpotentials for WC MXenes. (b) Comparison of overpotentials for W4C3O2 MXene with other materials.
Figure 6. (a) The ORR (ηORR), OER (ηOER), and total (ηTOT) overpotentials for WC MXenes. (b) Comparison of overpotentials for W4C3O2 MXene with other materials.
Nanomaterials 14 00666 g006
Figure 7. (a) The ORR overpotential (ƞORR) as the function of the adsorption energy of Li2O2 (Eads(Li2O2)). (b) The OER overpotential (ƞOER) as the function of the adsorption energy of LiO2 (Eads(LiO2)).
Figure 7. (a) The ORR overpotential (ƞORR) as the function of the adsorption energy of Li2O2 (Eads(Li2O2)). (b) The OER overpotential (ƞOER) as the function of the adsorption energy of LiO2 (Eads(LiO2)).
Nanomaterials 14 00666 g007
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

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. Nanomaterials 2024, 14, 666. https://doi.org/10.3390/nano14080666

AMA Style

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. Nanomaterials. 2024; 14(8):666. https://doi.org/10.3390/nano14080666

Chicago/Turabian Style

Zhu, Liwei, Jiajun Wang, Jie Liu, Ruxin Wang, Meixin Lin, Tao Wang, Yuchao Zhen, Jing Xu, and Lianming Zhao. 2024. "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" Nanomaterials 14, no. 8: 666. https://doi.org/10.3390/nano14080666

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