Next Article in Journal
Heat Treatments of Metastable β Titanium Alloy Ti-24Nb-4Zr-8Sn Processed by Laser Powder Bed Fusion
Next Article in Special Issue
Recent Advances of Indium Oxide-Based Catalysts for CO2 Hydrogenation to Methanol: Experimental and Theoretical
Previous Article in Journal
Analysis of Dispersion of Carbon Nanotubes in m-Cresol
Previous Article in Special Issue
Improving Anti-Coking Properties of Ni/Al2O3 Catalysts via Synergistic Effect of Metallic Nickel and Nickel Phosphides in Dry Methane Reforming
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Unveiling the Origin of Alkali Metal (Na, K, Rb, and Cs) Promotion in CO2 Dissociation over Mo2C Catalysts

1
School of Chemical Engineering, Sichuan University, Chengdu 610065, China
2
China-America Cancer Research Institute, Guangdong Provincial Key Laboratory of Medical Molecular Diagnostics, Guangdong Medical University, Dongguan 523808, China
3
National Supercomputing Center in Shenzhen (Shenzhen Cloud Computing Center), Shenzhen 518055, China
*
Authors to whom correspondence should be addressed.
Materials 2022, 15(11), 3775; https://doi.org/10.3390/ma15113775
Submission received: 9 April 2022 / Revised: 19 May 2022 / Accepted: 20 May 2022 / Published: 25 May 2022
(This article belongs to the Special Issue Nanocatalysts for CO2 Utilization)

Abstract

:
Molybdenum carbide (Mo2C) is a promising and low-cost catalyst for the reverse water−gas shift (RWGS) reaction. Doping the Mo2C surface with alkali metals can improve the activity of CO2 conversion, but the effect of these metals on CO2 conversion to CO remains poorly understood. In this study, the energies of CO2 dissociation and CO desorption on the Mo2C surface in the presence of different alkali metals (Na, K, Rb, and Cs) are calculated using density functional theory (DFT). Alkali metal doping results in increasing electron density on the Mo atoms and promotes the adsorption and activation of CO2 on Mo2C; the dissociation barrier of CO2 is decreased from 12.51 on Mo2C surfaces to 9.51–11.21 Kcal/mol on alkali metal-modified Mo2C surfaces. Energetic and electronic analyses reveal that although the alkali metals directly bond with oxygen atoms of the oxides, the reduction in the energy of CO2 dissociation can be attributed to the increased interaction between CO/O fragments and Mo in the transition states. The abilities of four alkali metals (Na, K, Rb, and Cs) to promote CO2 dissociation increase in the order Na (11.21 Kcal/mol) < Rb (10.54 Kcal/mol) < Cs (10.41 Kcal/mol) < K (9.51 Kcal/mol). Through electronic analysis, it is found that the increased electron density on the Mo atoms is a result of the alkali metal, and a greater negative charge on Mo results in a lower energy barrier for CO2 dissociation.

1. Introduction

Increasing atmospheric CO2 concentrations have resulted in global warming [1,2,3]. Therefore, CO2 capture, storage, and catalytic reduction have drawn attention to reduce this environmental burden [4]. In particular, the reverse water–gas shift (RWGS) reaction, which reduces CO2 to CO as an intermediate to generate methanol or other hydrocarbons, is promising [3]. The RWGS reaction is endothermic; thus, the RWGS reaction is thermodynamically favorable at high temperatures, as shown in Equation (1) [5].
C O 2 + H 2 C O + H 2 O           Δ H 298 K ° = 41   KJ / mol
Noble metal catalysts such as Pt [6,7], Rh [8,9], and Au [10] show reasonable activity and selectivity for the RWGS reaction but are costly. However, supported noble metal catalysts frequently suffer the problem of sintering under high temperature conditions. Moreover, noble metal catalysts are relatively expensive and scarce, which limits their ability to be widely used for CO2 hydrogenation. Transition metal carbides (TMCs) such as Mo2C [11,12], WC [13], and TiC [14] are considered as attractive candidates for the RWGS reaction because of their low cost and similar catalytic activity to platinum-based catalysts. TMCs have good performance in the reaction of CO2 conversion into CO [15], CH4 [16], CH3OH [17], and other hydrocarbons [18,19]. Among the carbide family members, molybdenum carbide (Mo2C) shows particularly high RWGS activity because of its favorable activity for C=O bond scission and H2 dissociation [20].
Alkali metals are very good promoters for CO2 conversion [21,22,23]. For example, CO2 activation was accelerated in K-modified CuxO/Cu (111) catalysts due to the geometric and electronic effects introduced by K [24]. The modification of Rh/Al2O3 with K changes the surroundings of the Rh particles, which influences the strength of CO adsorption and the activation ability of Rh for H2 dissociation [9]. The modification of Mo2C with alkali metals changes the structural and electronic properties of these catalysts and promotes the performances of Mo2C in CO2 conversions [25,26,27,28,29]. For example, the addition of 2 wt% K to Mo2C/γ-Al2O3 increases the CO selectivity to 95% from 73.5% [30], and the incorporation of K into Cu/Mo2C results in high CO2 dissociation activity (almost 1.5 times higher than Cu/Mo2C) but also reduces H2 adsorption, thus resulting in a low H2/COx ratio and low CH4 production [31]. The CO selectivity of Cs-Mo2C, which can reach 100% at low-temperatures (400 to 500 °C), is a result of increased electron transfer from Cs to Mo, thus favoring CO selectivity [28]. By the introduction of K into the single atom catalyst Rh0.2/β-Mo2C, the selectivity of hydrogenation of CO2 to ethanol is much-improved, and the catalysts exhibit up to 72.1% of ethanol selectivity at low temperature (150 °C) [32].
Although the promotion effects of alkali metals on TMCs have been observed experimentally, the structural and electronic effects of alkali metals on TMCs in the RWGS reaction remain unknown. Additionally, different alkali metals affect the WGS and RWGS reactions to various extents. For instance, when Na and K species are introduced into WC, they both promote the improvement of the selectivity of WC for the RWGS reaction at low temperatures (300–350 °C), and the highest CO yield is achieved using K-promoted WC [13]. Kowalik et al. [33] reported that the promotional effect of alkali metals on the WGS activity of Cu/ZnO/Al2O3 catalysts increases in the order of Li < Na < K < Cs, but the promotional effects of H2O and CO2 dissociation induced by alkali metals increase in the order of Na < K < Rb < Cs over Cu (111) catalysts [22]. However, the reasons underlying these observations remain unknown.
In this study, we investigated the effect of the modification of the surface of Mo2C catalysts with alkali metals (Na, K, Rb, and Cs) using density functional theory calculations and revealed the key electronic effects affecting the adsorption of the reactants and intermediate species, the desorption of the products, as well as the barrier of CO2 dissociation. Furthermore, the energy barrier of CO2 dissociation was correlated with the adsorption energy of surface species, thus revealing the origin of the promoting effects of various alkali metals on RWGS activity. Our findings further highlight the importance of modifying molybdenum carbide with alkali for carbon dioxide reduction.

2. Computation Detail and Models

All calculations were performed by using the DMol3 code within the Materials Studio 7.0 program [34]. The generalized gradient approximation (GGA) with the Perdew–Burke–Ernzerhof (PBE) functional [35] was selected to calculate the exchange-correlation energy. The wave functions were expanded by the utilization of the double numerical quality basis set with polarization functions (DNP) [36]. The energy, gradient, and displacement convergence criteria were 1 × 10−5 hartree, 2 × 10−3 hartree/Å, and 5 × 10−3 Å, respectively.
LST/QST was used to perform the transition states (TS) search [37,38]. The convergence criterion of the TS search was set to 0.002 Ha/Å on each atom. Only one virtual frequency could be considered as the real transition state.
The adsorption energy ( E ads ) of all intermediate species on the surface of catalyst was defined as:
E ads = E tot   -   E cat   -   E gas
where E tot is the total energy of the adsorbed species on the catalyst, E cat is the total energy of the clean catalyst, and E gas is the energy of the molecules in the gas phase. The activation barrier ( E Barrier ) and reaction energy (∆E) were calculated using the formulas:
E Barrier   = E TS   -   E IS
E = E FS   -   E IS
Here, E IS , E TS , and E FS represent the total energies of the initial state (IS), transition state (TS), and the final state (FS), respectively.
By applying geometry optimizations based on the minimization of the total energy of the unit cell, the DFT lattice parameters were found to be a = 6.00 Å, b = 5.78 Å, and c = 4.71 Å, which were in good agreement with the experimental results [39]. The slab model of the β-Mo2C (001) surface contained six atomic layers with a total of 24 C atoms and 48 Mo atoms in one unit cell (using a 2 × 2 supercell with size 12.00 × 11.57 × 4.71 Å, with a vacuum space of 20 Å). During the structural optimization, the bottom two layers were constrained in their bulk positions, whereas all the other atoms were allowed to relax. For alkali metal-modified β-Mo2C (001), one alkali atom (Na, K, Rb, Cs) was placed at different sites of the top layer of the molybdenum layer. After geometry optimization, the site that exhibited the strongest binding to K atom was selected for further calculations.

3. Results and Discussion

3.1. Optimized Structure of Alkali-Metal-Modified β-Mo2C (001)

The four alkali metal atoms at optimized structures of the X-Mo2C (X = Na, K, Rb, Cs) surfaces were all located on the 4F sites on the Mo2C surface, that is, between four Mo atoms (Figure 1). The alkali metal-promoted surfaces are shown in Figure 2a–d, and the key structural parameters of the X-Mo2C are listed in Table S1. The distances between alkali metals and Mo atoms increased with the increases in the atomic radii of the alkali metals: 3.26, 3.72, 3.89, and 4.04 Å on average for Na, K, Rb, and Cs, respectively. In addition, after the addition of the alkali metal, the Mo1–Mo3 and Mo3–M4 bonds increased in length by 0.04–0.06 Å, whereas the Mo1–Mo4 bond length was shortened by 0.03 Å for all X-Mo2C. The coverage of alkali metals on X-Mo2C surfaces was a 0.014 mono layer.
As shown by the charge analysis in Table 1, the Mulliken charge on Mo atoms in bare Mo2C was positive. In contrast, after the addition of the alkali metal, the charge on the adjacent Mo atoms became negative, suggesting the transfer of electrons from the alkali metal to Mo. Moreover, the closest Mo atoms to the alkali metal gained the most electrons. Because Rb and Cs atoms are less electronegative than the other alkali metals, they increased the electron density of the surface Mo atoms and subsurface C atoms. Therefore, in Rb- and Cs-Mo2C, the charges on Mo atoms were less negative than those of K and Na-Mo2C.
In addition, we calculated the changes in the d-band center [40] as a result of charge transfer between Mo and alkali metal atoms (Figure 2e). The d-band center before and after the addition of the alkali metal remained the same, consistent with previous findings [41,42].
Figure 2. The most stable structure of X−Mo2C (001): (a) Na−Mo2C, (b) K−Mo2C, (c) Rb−Mo2C, and (d) Cs−Mo2C; (e) d−band center of Mo atoms on Mo2C and X-Mo2C (X = Na, K, Rb, and Cs). Color codes: Mo—green, carbon—grey (shown in line model).
Figure 2. The most stable structure of X−Mo2C (001): (a) Na−Mo2C, (b) K−Mo2C, (c) Rb−Mo2C, and (d) Cs−Mo2C; (e) d−band center of Mo atoms on Mo2C and X-Mo2C (X = Na, K, Rb, and Cs). Color codes: Mo—green, carbon—grey (shown in line model).
Materials 15 03775 g002
Table 1. Mulliken charges (e) of key atoms in the clean and X-Mo2C (X = Na, K, Rb, Cs) *.
Table 1. Mulliken charges (e) of key atoms in the clean and X-Mo2C (X = Na, K, Rb, Cs) *.
CatalystsMulliken Charge (e)
XMo1Mo3Mo4C1C2C3
Mo2C/0.050.050.13−0.50−0.50−0.50
Na-Mo2C0.55−0.08−0.090.02−0.47−0.46−0.45
K-Mo2C0.72−0.10−0.090.03−0.52−0.51−0.49
Rb-Mo2C0.65−0.06−0.050.10−0.61−0.60−0.60
Cs-Mo2C0.66−0.07−0.060.09−0.54−0.53−0.53
* Atom labels are indicated in Figure 1.

3.2. Adsorption of Intermediate Species on Mo2C and X-Mo2C Surfaces

The schematic mechanism of the redox pathway is shown in Figure 3. We calculated the adsorption energy (Eads) and key parameters of CO2, CO, and O* species on bare and X-Mo2C surfaces involved in the redox pathway (shown in Figure 4 and Table S2).

3.2.1. Adsorption of CO2*

The most stable adsorption configuration of CO2 on Mo2C is shown in Figure 4, having an adsorption energy of −30.83 Kcal/mol, consistent with the literature value [30]. The C atom in CO2 was positioned at the bridge site between Mo3 and Mo5, and the two C–Mo bonds were 2.24 Å in length. The nearest distance between the O and Mo atoms was 2.12 Å for Oa–Mo3 and 2.43 Å for Ob–Mo5. The interaction of CO2 with Mo2C led to the elongation of the C–O bonds from 1.17 Å in the vacuum calculations to 1.26 Å in the adsorbed molecule, and the bond angle increased to 135°.
For the adsorption of CO2 on all X-Mo2C surfaces, the carbon atom of CO2 was attached to the top site of one Mo atom, and the oxygen atoms formed two Mo–O bonds with the adjacent Mo atoms. The Eads of CO2 was −41.17 Kcal/mol on Na-Mo2C, −38.74 Kcal/mol on K-Mo2C, −38.02 Kcal/mol on Rb-Mo2C, and −38.17 Kcal/mol on Cs-Mo2C. For CO2 adsorption on Na-Mo2C, the two O atoms in CO2 were equidistant from the Na atom (2.41 Å), CO2 was bent to 114°, and the C–O bonds were elongated to 2.41 Å. For the K-Mo2C catalyst, the K…Oa and K…Ob distances were found to be 2.74 and 2.97 Å, respectively, the CO2 bond was 115.6°, and the O–C bonds were elongated to 1.35 and 1.34 Å. For Rb-Mo2C, the Rb…O distances were 3.08 and 2.96 Å, and the C–Oa and C–Ob bond lengths were stretched to 1.34 and 1.36 Å, respectively. For Cs-Mo2, the Cs…Oa and Cs… Ob distances were 3.06 and 3.16 Å, respectively. Moreover, the C–Oa and C–Ob bond lengths were 1.33 and 1.36 Å, respectively.
Thus, the X–O distances were close to the sum of the X+ and O2− atomic radii, namely 2.41 Å for Na + +O2−, 2.77 Å for K+ + O2−, 2.91 Å for Rb+ + O2−, and 3.06 for Cs+ + O2−, as also observed for the X–O distances in crystalline metal oxides, for example, 2.40 Å in Na2O [43], 2.79 in K2O [43], 2.92 in Rb2O [43], and 3.26 Å in CsO2 [44]. These findings indicate that the electrostatic interaction between O and X was strong and similar to the ionic bonding between O and X in X2O. Wang et al. found that for oxygenate species adsorbed on K+-modified Cu (111) and Cu (110) surfaces, when the distance between K and O atoms was 3.00 Å, direct bonding between Oδ− and Kδ+ ions occurred [22]. Further, a short distance between X and O generated a longer C–O bond, indicating that stronger X–O interactions promote CO2 activation. Therefore, alkali metals could promote the adsorption and activation of CO2, consistent with theoretical and experimental findings [30].
Our charge analysis (Table S3) suggested that electrons are transferred from the alkali metal to Mo atoms and, thus, affect the surface charge of Mo2C [28]. Therefore, when CO2 was adsorbed on Mo2C, the Mo atoms lost electrons and CO2 gained 0.29 e. In contrast, when CO2 was adsorbed on X-Mo2C, the Mo atoms became more positive (lost more electrons); for example, CO2 gained 0.74 e on Na-Mo2C, 0.71 e on K-Mo2C, 0.73 e on Rb-Mo2C, and 0.69 e on Cs-Mo2C. The increase in the charge of CO2 also indicated enhanced charge transfer via alkali metal promotion. Moreover, in the X-Mo2C surfaces, alkali metal atoms lost electrons by 0.72 e for Na, 0.82 e for K, 0.76 e for Rb, and 0.74 e for Cs, respectively. Therefore, on one hand, alkali metal atoms donated electrons to CO2, but, on the other hand, they facilitated electron transfer from Mo to CO2 and thus promoted CO2 activation.

3.2.2. Adsorption of CO*

When CO adsorbed on the Mo2C surface, the carbon atom of CO2 was adsorbed on the bridge sites between Mo3 and Mo5 with an orientation tilted toward Mo4. The generated C–Mo4, C–Mo5, and O–Mo4 bond lengths were 2.27, 1.99, and 2.35 Å, respectively. Further, the C–O bond length was elongated to 1.23 Å, and the Eads of CO on Mo2C was −53.69 Kcal/mol.
The Eads values of CO on X-Mo2C (X = Na, K, Rb, and Cs) were around −56 Kcal/mol, approximately 1.75–2.67 Kcal/mol greater than that on Mo2C. The maximum differences in Eads on the X-Mo2C surfaces were within 1 Kcal/mol, i.e., negligible. In these systems, CO was adsorbed in a tilted orientation on the bridge sites of Mo3–Mo5 atoms with the oxygen atom oriented towards the Mo3 atom. The distances between alkali metal and O (in CO) were 2.42 Å for Na…O, 2.78Å for K…O, 2.93 Å for Rb…O, and 3.18 Å for Cs…O, suggesting that the alkali metals formed direct bonds with the O atom in CO in X-Mo2C systems. The Mulliken charge analysis (Table S4) showed that more electrons transferred to CO on X-Mo2C than that on bare Mo2C. Therefore, the addition of alkali metal atoms increased the adsorption of CO as compared with the clean surface and resulted in longer C–O bonds.

3.2.3. Adsorption of O*

Atomic oxygen adsorbed on the 3F sites between the Mo3–Mo4–Mo5 atoms, having average Mo–O bond lengths of 2.08 Å and Eads of −80.74 Kcal/mol. The Eads for O* on the surface of X-Mo2C (X = Na, K, Rb, and Cs) ranged from −81.64 to −82.10 Kcal/mol, slightly greater than that on bare Mo2C. The average distances between O* and Mo3 atoms were 2.46 Å for Na, 2.85Å for K, 3.03 Å for Rb, and 3.04 Å for Cs. On bare Mo2C, the atomic oxygen gained 0.66 e from the surface of Mo atoms, whereas for X-Mo2C, the O atom gained more electrons, namely 0.74 e from Na-Mo2C, 0.73 e from K-Mo2C, 0.72 e from Rb-Mo2C, and 0.73 e from Cs-Mo2C (see Table S5). Moreover, the Mo atoms in the X-Mo2C surfaces were more positive, suggesting that the addition of alkali metals promoted the loss of electrons from the Mo atoms around O*.

3.3. Energy Barriers for CO2 Dissociation on Mo2C and Alkali-Metal-Modified Mo2C Surfaces

Currently, the RWGS reaction mechanisms are classified into redox- (or direct-), carboxyl-, and formate-mediated routes. Chen et al. [45] performed ambient-pressure X-ray photoelectron spectroscopy (AP-XPS) measurements on the Mo2C catalyst, and they did not find intermediate species (carbonate, formate, carbonyl, etc.) under reaction conditions. Furthermore, they proved that CO2 was directly dissociated on Mo2C to produce CO and oxycarbide (Mo2C-O). Surface oxygen (Mo2C-O) was removed subsequently by hydrogen to produce H2O to complete the catalytic cycle [46]. Moreover, the elemental steps for CO2 dissociation to CO* and O* are known to be the rate limiting steps on both bare Mo2C and K-modified Mo2C catalysts [30]. Thus, the activation barriers for CO2 dissociation on the various alkali metal-modified Mo2C surfaces were studied and compared.
In Figure 5, we show the activation energy profiles of CO2 dissociation on the (a) bare, (b) Na-promoted, (c) K-promoted, (d) Rb-promoted, and (e) Cs-promoted Mo2C (001) surfaces. The activation barrier for CO2 dissociation on Mo2C surfaces was found to be 12.51 Kcal/mol, and the O–CO bond lengths of the transition states (TSs) were found to be 1.75 Å, indicating the cleavage of a C–O bond. The activation energies for this reaction were remarkably different on the four X-Mo2C surfaces, namely 11.21 Kcal/mol for Na-Mo2C, 9.51 Kcal/mol for K-Mo2C, 10.54 Kcal/mol for Rb-Mo2C, and 10.41 Kcal/mol for Cs-Mo2C, lower than that on bare Mo2C. For CO2 dissociation on the X-Mo2C surfaces, the bond length of C–Ob was elongated, ranging from 1.81 to 1.87 Å in the TS, respectively, longer than that on bare Mo2C. In addition, the distance between Na, K, Rb, and Cs and Oa in CO2 were 2.14, 2.94, 3.07, and 3.06 Å, respectively, suggesting the interaction between X and O throughout the reaction and suggesting the key role of the alkali metal in CO2 dissociation, consistent with experimental observations [47,48,49].
Figure 5 suggests that the transition state for CO2 dissociation is a late (product-like) transition state. Therefore, for CO2 dissociation, the stabilization of the final state should also stabilize the transition state, resulting in a lower activation barrier. Figure 5f shows that the CO2 dissociation barrier was strongly influenced by the E ads of CO and O fragments on the catalyst surfaces, and stronger CO or O binding resulted in lower CO2 dissociation barriers. Therefore, the identification of the role of the alkali metal atom on the stability of adsorbed CO and O during the reaction is necessary.

3.4. Energetic Analysis

As discussed earlier, alkali metals enhance the RWGS activity of Mo2C by decreasing the energy barriers for CO2 dissociation. Thus, to elucidate the effects of the alkali metals, the physical origin of the reaction barriers for CO2 dissociation on both bare and X-Mo2C surfaces were assessed using energy decomposition, as proposed by Hammer [50,51] (Equation (4)), and the results are listed in Table 2.
E Barrier = E bond CO 2   -   E CO 2 IS + E CO TS + E O TS + E int TS  
where E bond CO 2 represents the bonding energy of CO2 in gas. E CO 2 IS , E CO TS , E O TS , and E int TS   refer to binding energy of CO2* in the IS, binding energy of CO* (O*) in the TS, and the interaction of CO with O in the TS, respectively.
In the C–O bond scission of CO2 on Mo2C surfaces, alkali metals stabilize the binding of CO2 in the IS, which is unfavorable for reducing the energy barrier (Table 2). However, by strengthening CO and O binding in the TS ( E CO TS   and E O TS ) on X-Mo2C relative to those on clean surfaces, the alkali metal reduces the energy barrier and promote CO2 dissociation. In addition, all alkali metals can enhance the stability of CO or O fragments in the TS on Mo2C. However, Na- and K-modified surfaces effectively stabilize adsorbed CO compared to bare Mo2C ( Δ E CO TS > Δ E O TS ) , whereas Rb- and Cs-modified surfaces stabilize adsorbed O ( Δ E O TS > Δ E CO TS ) . Therefore, Max Δ E S TS , i.e., the maximum of Δ E CO TS and Δ E O TS was plotted against the energy barrier for CO2 dissociation. Figure 6 shows that the energy barrier for CO2 dissociation on X-Mo2C was linearly correlated to Max Δ E S TS (R2 = 0.90), and a greater value of Max Δ E S TS indicated a larger decrease of the barrier and a stronger promoting effect of the alkali metal. In other words, increasing the E ads of the CO and O fragments in the TS can effectively reduce the energy barrier, but the extent of the barrier reduction depends on the balance of stabilities of adsorbed CO and O resulting from alkali metal addition.
As displayed in Figure 5, the alkali metals interacted with CO and O on the X-Mo2C surfaces throughout the reaction. Hence, apart from the interaction between these adsorbates and surface Mo atoms, the interaction between them and the alkali metal adatom also made up the interaction of them with X-Mo2C.The interaction energies of the interactions between adsorbates and alkali metals were calculated with the following formulas, and the results are listed in Table 3.
E int A - X = E A / X TS   -   ( E A + Ex   -   E surf TS )
E int A - Mo = E A TS   -   E int A - X
E A / X TS , E A , and Ex represent the energies of the A−alkali metal complex, isolated A, and alkali metal at Mo2C surfaces, respectively; and E surf TS refers to the energy of the clean surface at the TS.
As shown in Table 3, the addition of alkali metal adatoms resulted in interactions between the alkali metal and oxygen species in the TS, but there was little correlation between E int CO - X / E int O - X and E Barrier . However, the addition of alkali metals increased the strength of the O−Mo or CO–Mo bonds, as shown by the greater E int CO - Mo and E int O - Mo on X-Mo2C compared to those on bare Mo2C. In addition, the increase in the CO–Mo and O–Mo interactions in X-Mo2C ( E int CO - Mo and E int O - Mo ) was linearly related to the increase in the bonding energies of CO ( Δ E CO TS )   and O ( Δ E O TS ) in the TS. Therefore, the increased bonding energies of CO and O in the TS were a result of the increased Mo–O and Mo–CO bond strength resulting from the addition of alkali metals.
Previous studies have demonstrated that the activation of CO2 requires electron transfer from the catalyst to CO2 [52,53,54,55]. The charge analysis in Table 3 shows that the enhancement in the Mo–O and Mo–CO interactions was due to the addition of alkali metals, which induced the accumulation of electron density at Mo and, thus, electron transfer from Mo to CO and O in the TS. Further, we observed a linear increase in the negative charge on the Mo atoms (increase in electron density) in the Na, Rb, Cs, and K-Mo2C surfaces, which was consistent with the reduction in E Barrier . This result indicated that the increase in charge at Mo resulted in a lower E Barrier . Additionally, although Rb and Cs are less electronegative than the other alkali metals, they transferred electrons to both Mo and C in the subsurface; thus, fewer electrons accumulated at Mo in Rb/Cs-Mo2C than in K-Mo2C.
In summary, the CO2 dissociation energy barrier was in the order of Mo2C (12.45 Kcal/mol) > Na (11.21 Kcal/mol) > Rb (10.54 Kcal/mol) > Cs (10.41 Kcal/mol) > K(9.51 Kcal/mol). This is because the K atom promoted the most electrons accumulated at the Mo atom and thereby generated the strongest Mo–CO interactions in the TS. The significantly improved stability of CO fragments in the TS led to the energy barrier of CO2 dissociation on K-Mo2C being the lowest.
Figure 6. Correlation between activation energy of the CO2 dissociation and MaxΔ E S TS (fitting model: y = 2.16x − 26.56; R2 = 0.90).
Figure 6. Correlation between activation energy of the CO2 dissociation and MaxΔ E S TS (fitting model: y = 2.16x − 26.56; R2 = 0.90).
Materials 15 03775 g006
Table 3. Mulliken charges (e) of key atoms in the calculated transition states.
Table 3. Mulliken charges (e) of key atoms in the calculated transition states.
MoCO2
Mo1 aMo2 bOaCObAlkali Metal
Mo2C0.190.30−0.300.15−0.50
Na-Mo2C0.120.17−0.410.19−0.610.74
K-Mo2C0.060.23−0.420.16−0.600.83
Rb-Mo2C0.200.19−0.390.13−0.580.77
Cs-Mo2C0.180.18−0.400.18−0.590.76
a Mo atom that bonded with Oa, b Mo atom that bonded with Ob.

3.5. CO Desorption on Mo2C and Alkali-Metal-Modified Mo2C Surfaces

Next, we calculated the energies of CO desorption on the surfaces of bare Mo2C and X-Mo2C (Figure 7). The CO desorption energy on bare Mo2C was endothermic by 63.65 Kcal/mol. The addition of alkali metals on Mo2C slightly increased the difficulty of CO desorption. Mpourmpakis et al. found that by pre-adsorption of low-coverage oxygen (<0.50 ML), the desorption of CO on K-modified Mo2C could be effectively promoted [56].

4. Conclusions

The activity of CO2 dissociation into CO on bare Mo2C and those on promoted surfaces of X-Mo2C (X = Na, K, Rb, and Cs) were studied using DFT calculations. The addition of alkali metal elements induced the accumulation of negative charges on the Mo atoms and thus promoted the adsorption and activation of CO2 on Mo2C. The CO2 dissociation energy barrier was in the order of Mo2C (12.45 Kcal/mol) > Na (11.21 Kcal/mol) > Rb (10.54 Kcal/mol) > Cs (10.41 Kcal/mol) > K (9.51 Kcal/mol). On the basis of energetic and electronic analysis, although the alkali metals directly bonded with oxygen atoms in the adsorbed oxygen species, the main reason for the reduction in the energy of CO2 dissociation was the stronger interaction between CO/O fragments and Mo in the TS. Through electronic analysis, the promoting effects of alkali metals were influenced by the difference in the increase of electron density at the Mo atoms. Specifically, the greater the negative charge on the Mo site, the lower the energy barrier for CO2 dissociation. In comparison, the K atom promoted the most electrons accumulated at the Mo atom and thereby generated the strongest Mo–CO interactions in the TS. The significantly improved stability of CO fragments in the TS led to the energy barrier of CO2 dissociation on K-Mo2C being the lowest.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma15113775/s1, Table S1: The key structural parameters of Mo2C and X-Mo2C; Table S2: Adsorption energies (Kcal/mol) and bond length (Å) of all possible intermediates on the X (Na, K, Rb, Cs)-Mo2C catalysts; Table S3: Mulliken charge; Table S4: Mulliken charge (e) of CO*-X-Mo2C; Table S5: Mulliken charge (e) of O*-X-Mo2C.

Author Contributions

Investigation, formal analysis and writing—original draft preparation, R.L.; Resources, software, C.C.; Formal analysis, investigation, writing-review-editing, W.C. Formal analysis, investigation, writing-review-editing, W.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Foundation of the State Key Laboratory of High-Efficiency Utilization of Coal and Green Chemical Engineering (2019-KF-09) and the Guangdong Basic and Applied Basic Research Foundation (2020A1515010490).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available from the corresponding author, upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pan, Y.-X.; Liu, C.-J.; Ge, Q. Effect of surface hydroxyls on selective CO2 hydrogenation over Ni4/γ-Al2O3: A density functional theory study. J. Catal. 2010, 272, 227–234. [Google Scholar] [CrossRef]
  2. Liu, C.; Liu, P. Mechanistic study of methanol synthesis from CO2 and H2 on a modified model Mo6S8 cluster. ACS Catal. 2015, 5, 1004–1012. [Google Scholar] [CrossRef]
  3. Wang, J.; Liu, C.-Y.; Senftle, T.P.; Zhu, J.; Zhang, G.; Guo, X.; Song, C. Variation in the In2O3 crystal phase alters catalytic performance toward the reverse water gas shift reaction. ACS Catal. 2019, 10, 3264–3273. [Google Scholar] [CrossRef]
  4. Aresta, M.; Dibenedetto, A.; Angelini, A. Catalysis for the Valorization of Exhaust Carbon: From CO2 to Chemicals, Materials, and Fuels. Technological Use of CO2. Chem. Rev. 2014, 114, 1709–1742. [Google Scholar] [CrossRef]
  5. Daza, Y.A.; Kuhn, J.N. CO2 conversion by reverse water gas shift catalysis: Comparison of catalysts, mechanisms and their consequences for CO2 conversion to liquid fuels. RSC Adv. 2016, 6, 49675–49691. [Google Scholar] [CrossRef]
  6. Zhao, Z.; Wang, M.; Ma, P.; Zheng, Y.; Chen, J.; Li, H.; Zhang, X.; Zheng, K.; Kuang, Q.; Xie, Z.-X. Atomically dispersed Pt/CeO2 catalyst with superior CO selectivity in reverse water gas shift reaction. Appl. Catal. B Environ. 2021, 291, 120101. [Google Scholar] [CrossRef]
  7. Kattel, S.; Yan, B.; Chen, J.G.; Liu, P. CO2 hydrogenation on Pt, Pt/SiO2 and Pt/TiO2: Importance of synergy between Pt and oxide support. J. Catal. 2016, 343, 115–126. [Google Scholar] [CrossRef] [Green Version]
  8. Tang, Y.; Asokan, C.; Xu, M.; Graham, G.W.; Pan, X.; Christopher, P.; Li, J.; Sautet, P. Rh single atoms on TiO2 dynamically respond to reaction conditions by adapting their site. Nat. Commun. 2019, 10, 4488. [Google Scholar] [CrossRef] [Green Version]
  9. Heyl, D.; Rodemerck, U.; Bentrup, U. Mechanistic study of low-temperature CO2 hydrogenation over modified Rh/Al2O3 catalysts. ACS Catal. 2016, 6, 6275–6284. [Google Scholar] [CrossRef]
  10. Aitbekova, A.; Wu, L.; Wrasman, C.J.; Boubnov, A.; Hoffman, A.S.; Goodman, E.D.; Bare, S.R.; Cargnello, M. Low-temperature restructuring of CeO2-supported Ru nanoparticles determines selectivity in CO2 catalytic reduction. J. Am. Chem. Soc. 2018, 140, 13736–13745. [Google Scholar] [CrossRef]
  11. Gao, J.; Wu, Y.; Jia, C.; Zhong, Z.; Gao, F.; Yang, Y.; Liu, B. Controllable synthesis of α-MoC1-x and β-Mo2C nanowires for highly selective CO2 reduction to CO. Catal. Commun. 2016, 84, 147–150. [Google Scholar] [CrossRef]
  12. Zhang, X.; Zhu, X.; Lin, L.; Yao, S.; Zhang, M.; Liu, X.; Wang, X.; Li, Y.-W.; Shi, C.; Ma, D. Highly dispersed copper over β-Mo2C as an efficient and stable catalyst for the reverse water gas shift (RWGS) reaction. ACS Catal. 2016, 7, 912–918. [Google Scholar] [CrossRef]
  13. Morse, J.R.; Juneau, M.; Baldwin, J.W.; Porosoff, M.D.; Willauer, H.D. Alkali promoted tungsten carbide as a selective catalyst for the reverse water gas shift reaction. J. CO2 Util. 2020, 35, 38–46. [Google Scholar] [CrossRef]
  14. Rodriguez, J.A.; Evans, J.; Feria, L.; Vidal, A.B.; Liu, P.; Nakamura, K.; Illas, F. CO2 hydrogenation on Au/TiC, Cu/TiC, and Ni/TiC catalysts: Production of CO, methanol, and methane. J. Catal. 2013, 307, 162–169. [Google Scholar] [CrossRef]
  15. Liu, X.; Kunkel, C.; Ramírez de la Piscina, P.; Homs, N.; Viñes, F.; Illas, F. Effective and highly selective CO generation from CO2 using a polycrystalline α-Mo2C catalyst. ACS Catal. 2017, 7, 4323–4335. [Google Scholar] [CrossRef] [Green Version]
  16. Xu, W.; Ramírez, P.J.; Stacchiola, D.; Brito, J.L.; Rodriguez, J.A. The Carburization of Transition Metal Molybdates (MxMoO4, M=Cu, Ni or Co) and the Generation of Highly Active Metal/Carbide Catalysts for CO2 Hydrogenation. Catal. Lett. 2015, 145, 1365–1373. [Google Scholar] [CrossRef]
  17. Posada-Pérez, S.; Ramírez, P.J.; Gutiérrez, R.A.; Stacchiola, D.J.; Viñes, F.; Liu, P.; Illas, F.; Rodriguez, J.A. The conversion of CO2 to methanol on orthorhombic β-Mo2C and Cu/β-Mo2C catalysts: Mechanism for admetal induced change in the selectivity and activity. Catal. Sci. Technol. 2016, 6, 6766–6777. [Google Scholar] [CrossRef]
  18. Li, N.; Chen, X.; Ong, W.-J.; MacFarlane, D.R.; Zhao, X.; Cheetham, A.K.; Sun, C. Understanding of Electrochemical Mechanisms for CO2 Capture and Conversion into Hydrocarbon Fuels in Transition-Metal Carbides (MXenes). ACS Nano 2017, 11, 10825–10833. [Google Scholar] [CrossRef]
  19. Shi, Z.; Yang, H.; Gao, P.; Chen, X.; Liu, H.; Zhong, L.; Wang, H.; Wei, W.; Sun, Y. Effect of alkali metals on the performance of CoCu/TiO2 catalysts for CO2 hydrogenation to long-chain hydrocarbons. Chin. J. Catal. 2018, 39, 1294–1302. [Google Scholar] [CrossRef]
  20. Posada-Pérez, S.; Viñes, F.; Ramirez, P.J.; Vidal, A.B.; Rodriguez, J.A.; Illas, F. The bending machine: CO2 activation and hydrogenation on δ-MoC(001) and β-Mo2C(001) surfaces. Phys. Chem. Chem. Phys. 2014, 16, 14912–14921. [Google Scholar] [CrossRef] [Green Version]
  21. Yang, X.; Su, X.; Chen, X.; Duan, H.; Liang, B.; Liu, Q.; Liu, X.; Ren, Y.; Huang, Y.; Zhang, T. Promotion effects of potassium on the activity and selectivity of Pt/zeolite catalysts for reverse water gas shift reaction. Appl. Catal. B Environ. 2017, 216, 95–105. [Google Scholar] [CrossRef]
  22. Wang, Y.-X.; Wang, G.-C. A systematic theoretical study of water gas shift reaction on Cu(111) and Cu(110): Potassium effect. ACS Catal. 2019, 9, 2261–2274. [Google Scholar] [CrossRef]
  23. Wang, Y.-X.; Zhang, H.-L.; An, P.; Wu, H.-S.; Jia, J.-F. Effect of Potassium on Methanol Steam Reforming on the Cu(111) and Cu(110) Surfaces: A DFT Study. J. Phys. Chem. C 2021, 125, 20905–20918. [Google Scholar] [CrossRef]
  24. An, W.; Xu, F.; Stacchiola, D.; Liu, P. Potassium-Induced Effect on the Structure and Chemical Activity of the CuxO/Cu(1 1 1) (x ≤ 2) Surface: A Combined Scanning Tunneling Microscopy and Density Functional Theory Study. ChemCatChem 2015, 7, 3865–3872. [Google Scholar] [CrossRef]
  25. Feng, Z.; Su, G.; Ding, H.; Ma, Y.; Li, Y.; Tang, Y.; Dai, X. Atomic alkali metal anchoring on graphdiyne as single-atom catalysts for capture and conversion of CO2 to HCOOH. Mol. Catal. 2020, 494, 111142. [Google Scholar] [CrossRef]
  26. Gao, M.; Zhang, J.; Zhu, P.; Liu, X.; Zheng, Z. Unveiling the origin of alkali metal promotion in CO2 methanation over Ru/ZrO2. Appl. Catal. B Environ. 2022, 314, 121476. [Google Scholar] [CrossRef]
  27. Juneau, M.; Vonglis, M.; Hartvigsen, J.; Frost, L.; Bayerl, D.; Dixit, M.; Mpourmpakis, G.; Morse, J.R.; Baldwin, J.W.; Willauer, H.D.; et al. Assessing the viability of K-Mo2C for reverse water–gas shift scale-up: Molecular to laboratory to pilot scale. Energy Environ. Sci. 2020, 13, 2524–2539. [Google Scholar] [CrossRef]
  28. Zhang, Q.; Pastor-Pérez, L.; Jin, W.; Gu, S.; Reina, T.R. Understanding the promoter effect of Cu and Cs over highly effective β-Mo2C catalysts for the reverse water-gas shift reaction. Appl. Catal. B Environ. 2019, 244, 889–898. [Google Scholar] [CrossRef]
  29. Bugyi, L.; Oszkó, A.; Solymosi, F. Spectroscopic study on the formation of CO−2 on K-promoted Mo2C/Mo(100) surface. Surf. Sci. 2000, 461, 177–190. [Google Scholar] [CrossRef]
  30. Porosoff, M.D.; Baldwin, J.W.; Peng, X.; Mpourmpakis, G.; Willauer, H.D. Potassium-promoted molybdenum carbide as a highly active and selective catalyst for CO2 conversion to CO. ChemSusChem 2017, 10, 2408–2415. [Google Scholar] [CrossRef]
  31. Xu, J.; Gong, X.; Hu, R.; Liu, Z.-W.; Liu, Z.-T. Highly active K-promoted Cu/β-Mo2C catalysts for reverse water gas shift reaction: Effect of potassium. Mol. Catal. 2021, 516, 111954. [Google Scholar] [CrossRef]
  32. Ye, X.; Ma, J.; Yu, W.; Pan, X.; Yang, C.; Wang, C.; Liu, Q.; Huang, Y. Construction of bifunctional single-atom catalysts on the optimized β-Mo2C surface for highly selective hydrogenation of CO2 into ethanol. J. Energy Chem. 2022, 67, 184–192. [Google Scholar] [CrossRef]
  33. Kowalik, P.; Próchniak, W.; Borowiecki, T. The effect of alkali metals doping on properties of Cu/ZnO/Al2O3 catalyst for water gas shift. Catal. Today 2011, 176, 144–148. [Google Scholar] [CrossRef]
  34. Delley, B. From molecules to solids with the DMol3 approach. J. Chem. Phys. 2000, 113, 7756–7764. [Google Scholar] [CrossRef]
  35. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [Green Version]
  36. Delley, B. An all-electron numerical method for solving the local density functional for polyatomic molecules. J. Chem. Phys. 1990, 92, 508–517. [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. Henkelman, G.; Jónsson, H. Improved tangent estimate in the nudged elastic band method for finding minimum energy paths and saddle points. J. Chem. Phys. 2000, 113, 9978–9985. [Google Scholar] [CrossRef] [Green Version]
  39. Yang, C.; Guo, K.; Yuan, D.; Cheng, J.; Wang, B. Unraveling Reaction Mechanisms of Mo2C as Cathode Catalyst in a Li-CO2 Battery. J. Am. Chem. Soc. 2020, 142, 6983–6990. [Google Scholar] [CrossRef]
  40. Hammer, B.; Nørskov, J.K. Electronic factors determining the reactivity of metal surfaces. Surf. Sci. 1995, 343, 211–220. [Google Scholar] [CrossRef]
  41. Mortensen, J.J.; Hammer, B.; Nørskovi, J.K. A theoretical study of adsorbate–adsorbate interactions on Ru(0001). Surf. Sci. 1998, 414, 315–329. [Google Scholar] [CrossRef]
  42. Günther, S.; Esch, F.; del Turco, M.; Africh, C.; Comelli, G.; Kiskinova, M. K-stabilized high-oxygen-coverage states on Rh(110):  A low-pressure pathway to formation of surface oxide. J. Phys. Chem. B 2005, 109, 11980–11985. [Google Scholar] [CrossRef] [PubMed]
  43. McGuire, N.K.; O’Keeffe, M. Bond lengths in alkali metal oxides. J. Solid State Chem. 1984, 54, 49–53. [Google Scholar] [CrossRef]
  44. Tai, J.; Ge, Q.; Davis, R.J.; Neurock, M. Adsorption of CO2 on Model Surfaces of Cesium Oxides Determined from First Principles. J. Phys. Chem. B 2004, 108, 16798–16805. [Google Scholar] [CrossRef]
  45. Porosoff, M.D.; Yang, X.; Boscoboinik, J.A.; Chen, J.G. Molybdenum carbide as alternative catalysts to precious metals for highly selective reduction of CO2 to CO. Angew. Chem. Int. Ed. 2014, 53, 6705–6709. [Google Scholar] [CrossRef]
  46. Porosoff, M.D.; Kattel, S.; Li, W.; Liu, P.; Chen, J.G. Identifying trends and descriptors for selective CO2 conversion to CO over transition metal carbides. Chem. Commun. 2015, 51, 6988–6991. [Google Scholar] [CrossRef]
  47. Lee, J.S.; Kim, S.; Kim, Y.G. Electronic and geometric effects of alkali promoters in CO hydrogenation over K/Mo2C catalysts. Top. Catal. 1995, 2, 127–140. [Google Scholar] [CrossRef]
  48. Bugyi, L.; Solymosi, F. Effects of Potassium on the Chemisorption of CO on the Mo2C/Mo(100) Surface. J. Phys. Chem. B 2001, 105, 4337–4342. [Google Scholar] [CrossRef]
  49. Solymosi, F.; Bugyi, L. Effects of potassium on the chemisorption of CO2 and CO on the Mo2C/Mo (100) surface. Catal. Lett. 2000, 66, 227–230. [Google Scholar] [CrossRef]
  50. Hammer, B. Adsorption, diffusion, and dissociation of NO, N and O on flat and stepped Ru(0001). Surf. Sci. 2000, 459, 323–348. [Google Scholar] [CrossRef]
  51. Wang, Y.-X.; Wang, G.-C. A systematic theoretical study of the water gas shift reaction on the Pt/ZrO2 interface and Pt(111) face: Key role of a potassium additive. Catal. Sci. Technol. 2020, 10, 876–892. [Google Scholar] [CrossRef]
  52. Austin, N.; Ye, J.; Mpourmpakis, G. CO2 activation on Cu-based Zr-decorated nanoparticles. Catal. Sci. Technol. 2017, 7, 2245–2251. [Google Scholar] [CrossRef]
  53. Freund, H.-J.; Roberts, M.W. Surface chemistry of carbon dioxide. Surf. Sci. Rep. 1996, 25, 225–273. [Google Scholar] [CrossRef] [Green Version]
  54. Wurth, W.; Stöhr, J.; Feulner, P.; Pan, X.; Bauchspiess, K.R.; Baba, Y.; Hudel, E.; Rocker, G.; Menzel, D. Bonding, structure, and magnetism of physisorbed and chemisorbed O2 on Pt(111). Phy. Rev. Lett. 1990, 65, 2426–2429. [Google Scholar] [CrossRef] [PubMed]
  55. Austin, N.; Butina, B.; Mpourmpakis, G. CO2 activation on bimetallic CuNi nanoparticles. Prog. Nat. Sci. Mater. 2016, 26, 487–492. [Google Scholar] [CrossRef] [Green Version]
  56. Dixit, M.; Peng, X.; Porosoff, M.D.; Willauer, H.D.; Mpourmpakis, G. Elucidating the role of oxygen coverage in CO2 reduction on Mo2C. Catal. Sci. Technol. 2017, 7, 5521–5529. [Google Scholar] [CrossRef]
Figure 1. Position of the selected adsorption sites for alkali metal atoms. Color codes: Mo—green, carbon—grey, the first surface layer is shown using the ball and stick model, and the last five layers are shown using the line model.
Figure 1. Position of the selected adsorption sites for alkali metal atoms. Color codes: Mo—green, carbon—grey, the first surface layer is shown using the ball and stick model, and the last five layers are shown using the line model.
Materials 15 03775 g001
Figure 3. Schematic mechanism diagram of redox pathway in the RWGS reaction. The species with asterisks (*) represent adsorbed species.
Figure 3. Schematic mechanism diagram of redox pathway in the RWGS reaction. The species with asterisks (*) represent adsorbed species.
Materials 15 03775 g003
Figure 4. The most stable adsorption configurations of all possible surface intermediates as well as the adsorption energies (in Kcal/mol) on Mo2C and X−Mo2C (X = Na, K, Rb, and Cs). The species with asterisks (*) represent adsorbed species. Color codes: Mo—green, carbon—grey, O—red.
Figure 4. The most stable adsorption configurations of all possible surface intermediates as well as the adsorption energies (in Kcal/mol) on Mo2C and X−Mo2C (X = Na, K, Rb, and Cs). The species with asterisks (*) represent adsorbed species. Color codes: Mo—green, carbon—grey, O—red.
Materials 15 03775 g004
Figure 5. CO2 dissociation profiles on (a) Mo2C, (b) Na−Mo2C, (c) K−Mo2C, (d) Rb−Mo2C, and (e) Cs−Mo2C. (f) Correlation between activation energy of the CO2 dissociation and CO/O bonding energy.
Figure 5. CO2 dissociation profiles on (a) Mo2C, (b) Na−Mo2C, (c) K−Mo2C, (d) Rb−Mo2C, and (e) Cs−Mo2C. (f) Correlation between activation energy of the CO2 dissociation and CO/O bonding energy.
Materials 15 03775 g005
Figure 7. The desorption energy of CO on bare and X-Mo2C, where X = Na, K, Rb, Cs.
Figure 7. The desorption energy of CO on bare and X-Mo2C, where X = Na, K, Rb, Cs.
Materials 15 03775 g007
Table 2. Energy decomposition of the calculated activation barriers for the dissociation of CO2 on bare Mo2C and X-Mo2C.
Table 2. Energy decomposition of the calculated activation barriers for the dissociation of CO2 on bare Mo2C and X-Mo2C.
Catalysts E Barrier E CO 2 IS E CO TS E O TS E int TS E int CO - X   a E int CO - Mo E int O - X   a E int O - Mo Δ E C O T S   b Δ E O T S   b
Mo2C12.45−50.78−44.89−72.33−68.640.00−44.890.00−72.330.000.00
Na-Mo2C11.30−60.28−47.39−74.22−75.04−0.09−47.29−0.28−73.94−2.50−1.89
K-Mo2C9.45−60.44−51.82−73.31−73.38−0.12−51.71−0.32−73.00−6.930.99
Rb-Mo2C10.61−61.34−47.14−75.08−76.17−0.12−47.03−0.25−74.82−2.25−2.75
Cs-Mo2C10.38−60.11−45.36−76.04−75.89−5.77−45.32−0.25−75.77−0.47−3.71
a The interaction energies of A-X as well as between A and Mo atoms at the TSs (A refers to adsorption species). The C−O bonding energy of CO2 in the gas phase is calculated to be 147.59 Kcal/mol. b Δ E A TS = E A ,   X - Mo 2 C   TS E A ,   Mo 2 C   TS .
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, R.; Chen, C.; Chu, W.; Sun, W. Unveiling the Origin of Alkali Metal (Na, K, Rb, and Cs) Promotion in CO2 Dissociation over Mo2C Catalysts. Materials 2022, 15, 3775. https://doi.org/10.3390/ma15113775

AMA Style

Liu R, Chen C, Chu W, Sun W. Unveiling the Origin of Alkali Metal (Na, K, Rb, and Cs) Promotion in CO2 Dissociation over Mo2C Catalysts. Materials. 2022; 15(11):3775. https://doi.org/10.3390/ma15113775

Chicago/Turabian Style

Liu, Renmin, Congmei Chen, Wei Chu, and Wenjing Sun. 2022. "Unveiling the Origin of Alkali Metal (Na, K, Rb, and Cs) Promotion in CO2 Dissociation over Mo2C Catalysts" Materials 15, no. 11: 3775. https://doi.org/10.3390/ma15113775

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