Next Article in Journal
Highly Active TiO2 Photocatalysts for Hydrogen Production through a Combination of Commercial TiO2 Material Selection and Platinum Co-Catalyst Deposition Using a Colloidal Approach with Green Reductants
Next Article in Special Issue
Efficient Cross-Coupling of Acetone with Linear Aliphatic Alcohols over Supported Copper on a Fluorite-Type Pr2Zr2O7
Previous Article in Journal
Solochrome Dark Blue Azo Dye Removal by Sonophotocatalysis Using Mn2+ Doped ZnS Quantum Dots
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

CO2 Hydrogenation on NixMg1−xAl2O4: A Comparative Study of MgAl2O4 and NiAl2O4

1
School of Chemical Engineering, Yeungnam University, 280 Daehak-ro, Gyeongsan 38541, Gyeongbuk, Korea
2
Institute of Clean Technology, Yeungnam University, 280 Daehak-ro, Gyeongsan 38541, Gyeongbuk, Korea
*
Authors to whom correspondence should be addressed.
B.S. and E.H.K. contributed equally.
Catalysts 2021, 11(9), 1026; https://doi.org/10.3390/catal11091026
Submission received: 3 August 2021 / Revised: 20 August 2021 / Accepted: 23 August 2021 / Published: 24 August 2021
(This article belongs to the Special Issue Catalysis for Environmentally Benign Production of Alternative Fuels)

Abstract

:
Due to the increasing attention focused on global warming, many studies on reducing CO2 emissions and developing sustainable energy strategies have recently been performed. One of the approaches is CO2 methanation, transforming CO2 into methane. Such transformation (CO2 + 4H2 → CH4 + 2H2O) provides advantages of carbon liquification, storage, etc. In this study, we investigated CO2 methanation on nickel–magnesium–alumina catalysts both experimentally and computationally. We synthesized the catalysts using a precipitation method, and performed X-ray diffraction, temperature-programmed reduction, and N2 adsorption–desorption tests to characterize their physical and chemical properties. NiAl2O4 and MgAl2O4 phases were clearly observed in the catalysts. In addition, we conducted CO2 hydrogenation experiments by varying with temperatures to understand the reaction. Our results showed that CO2 conversion increases with Ni concentration and that MgAl2O4 exhibits high selectivity for CO. Density functional theory calculations explained the origin of this selectivity. Simulations predicted that adsorbed CO on MgAl2O4(100) weakly binds to the surface and prefers to desorb from the surface than undergoing further hydrogenation. Electronic structure analysis showed that the absence of a d orbital in MgAl2O4(100) is responsible for the weak binding of CO to MgAl2O4. We believe that this finding regarding the origin of the CO selectivity of MgAl2O4 provides fundamental insight for the design methanation catalysts.

1. Introduction

The amount of fossil fuel used worldwide continues to increase, and the resulting greenhouse gas emissions are now known to represent a major global challenge. The minimization of CO2 emitted from fossil-based energy sources is a priority, and renewable energy production units such as solar, wind, and geothermal energy plants are actively being installed [1]. Although solar power is now considered as one of the most economically competitive energy resources, large-scale investment is still needed to replace fossil-based energy for massive-scale energy production [2]. Given this scenario, the use of fossil fuels coupled with CO2 capture and utilization offers a near-term solution that meets the need for economically sustainable energy [3].
CO2 capture, utilization, and sequestration (CCUS) technology can provide a means of sustainable fossil-based energy use [4]. In terms of the large-scale CO2 utilization, CO2 can be hydrogenated to produce CH4 (by methanation; CO2 + 4H2 → CH4 + 2H2O) [5] which can be readily liquified, transported, and stored. In addition, CO, a central intermediate for Fischer–Tropsch and methanol synthesis, can also be produced from CO2 hydrogenation (by reverse water gas shift (RWGS); CO2 + H2 → CO + H2O) [6].
Depending on reaction conditions used for CO2 hydrogenation, CO2 can be converted into CH4 or CO by methanation or RWGS, respectively [7], although based on considerations of the scope of CH4 utilization and existing natural gas infrastructure, methanation is preferred.
Group 8 to 10 metals such as Ni, Ru, Rh, Co, and Fe are catalytically active with respect to CO2 hydrogenation [8,9]. Among these metals, nickel is preferred as an active catalyst metal because it is cheaper than Ru or Rh. The particle size of nickel strongly determines its selectivity for CO2 hydrogenation products produced by methanation or RWGS [10]. For example, it has been reported that smaller nickel particles favor high CO selectivity, whereas larger particles are more selective toward CH4 [11]. The mechanism of CO2 hydrogenation is believed to proceed via the hydrogenation of a surface-adsorbed intermediate [12]. CO2 is adsorbed on nickel and dissociates into CO, which is strongly adsorbed and further dissociated into atomic carbon that is hydrogenated to form CH4 [13]. On the other hand, weakly adsorbed CO can desorb into the gas phase [14,15]. In addition to the direct dissociation of CO2 on nickel, the dissociation of H2 and the resulting atomic hydrogen can facilitate the dissociation of C–O bonds. This hydrogen-assisted mechanism has been reported by observing the formation of formates and carbon hydroxyl species on a nickel-based catalyst [7].
Alumina has widely been used as a support for nickel catalysts due to its moderate price, high surface area, and ability to resist high temperatures. Although nickel-on-alumina catalysts are highly efficient for CO2 methanation, nickel tends to be dissolved into the alumina to form nickel aluminate (NiAl2O4; spinel structured). The catalytic activity of nickel for CO2 methanation is greater than nickel aluminate; therefore, efforts have been made to suppress the formation of nickel aluminate [16]. MgO was added to nickel–alumina catalysts to minimize nickel aluminate formation, but magnesium aluminate was generated [17]. The addition of Mg can boost the basicity of the alumina support and suppress carbon deposition during CO2 hydrogenation, whereas higher Ni content promotes higher selectivity for CO2 methanation due to the dissociation of H2 on the Ni0 site [18]. When MgO is added to nickel–alumina catalysts, three possibly active structures for CO2 hydrogenation may be formed, viz., nickel, nickel aluminate, and magnesium aluminate.
The role played by nickel in CO2 hydrogenation is well understood in terms of particle size [10] and metal–reactant intermediate reactions [19]. It has also been reported that oxygen vacancies in NiAl2O4 promote the dissociation of CO2 to CO [20] and facilitate the hydrogenation of adsorbed CO to atomic carbon. However, little is known of the role of magnesium aluminate on CO2 hydrogenation. In this study, we explored the use of nickel and magnesium aluminates catalysts for CO2 hydrogenation, experimentally and computationally.
The following questions are addressed in this article: (1) What are properties of the nickel–magnesium–alumina catalysts synthesized by co-precipitation?; (2) Are nickel and magnesium aluminates active CO2 hydrogenation catalysts?; (3) What is the reaction mechanism responsible for CO2 hydrogenation on nickel and magnesium aluminate?

2. Experimental Section

2.1. Synthesis of NixMg1−xAl2O4 Catalysts

NixMg1−xAl2O4 catalysts were prepared by co-precipitation. Stoichiometric amounts of metal nitrate precursors (Ni(NO3)2·6H2O (Sigma-Aldrich ≥ 99.0%), Mg(NO3)2·6H2O (Sigma-Aldrich ≥ 99.0%), and Al(NO3)3·9H2O (Sigma-Aldrich ≥ 98.0%)) were dissolved in deionized water, stirred for 1 h, and treated with ammonium bicarbonate ((NH3)HCO3, (Sigma-Aldrich ≥ 99.0%) dropwise to a pH of 7. Solid precipitates were then aged overnight and filtered. The cake obtained was washed with deionized water 5 times, dried for 24 h at 100 °C, crushed, and calcined at 900 °C under atmospheric conditions overnight. The calcined powder was then pulverized in a ball mill using high-purity alumina balls. The final product was named NixMg1−xAl2O4, where the subscript ‘x’ is the molar ratio of Ni to Mg.

2.2. Catalytic Activity Tests

Catalytic activity for CO2 hydrogenation was measured using an experimental setup similar to one previously reported [16]. Catalyst powder (5.0 g) was vertically loaded with quartz wool into a fixed bed reactor (inner diameter 1/2 inch) mounted in a 3-zone furnace. Initially, the catalyst was heated to the reaction temperature at 10 °C/min under 800 sccm N2 flow. The temperature of each heating zone was precisely monitored, and the flow rate of the gas mixture was controlled using a mass flow controller (MFC). The reactant gas CO2:H2:N2 (ratio 1:4:6.5 vol/vol) was then supplied at 1013 sccm into the heated reactor. WHSV (weight hourly space velocity) was set to ~12,000 mL·g−1·h−1. Steam in the effluent gas was condensed out using a cold trap, and the dried product gas mixture was monitored online using a gas chromatograph (Donam, DS6200) equipped with a thermo conductivity detector and gas analyzer with an infrared sensor (ABB-AO2000). CO2 conversion, CH4 selectivity, and CO selectivity were calculated as follows.
CO 2   conversion = F C O 2   i n F C O 2   o u t F C O 2   i n
CH 4   selectivity = F C H 4   o u t F C O 2   i n F C O 2   o u t
CO   selectivity = F C O   o u t F C O 2   i n F C O 2   o u t
where Fi is the volumetric flow rate of species i (sccm).

2.3. Characterization of Catalysts

Crystalline structures of catalysts were analyzed by powder X-ray diffraction (XRD, nickel-filtered CuKα radiation, 40.0 kV and 15.0 mA, Miniflex, Rigaku, Japan) using a scanning range from 10° to 90° and a scanning rate of 10°/min. Temperature-programmed reduction (TPR) by H2 was measured by chemisorption (BELCAT-B, MICROTRAC, Japan). Before measurements, samples were degassed at 300 °C under flowing Ar for 30 min and cooled to 30 °C at the same Ar flow rate. For TPR, samples (50 mg) were heated from 100 to 900 °C at 10 °C/min under a flowing Ar/10 vol.% H2 mix. Effluent gas was monitored using a thermal conductivity detector.

2.4. Computation Details

All plane-wave DFT calculations were performed using projector-augmented wave pseudopotentials [21] provided in the Vienna ab initio simulation package (VASP) [22]. The Perdew–Burke–Ernzerhof (PBE) [23] exchange-correlation was used with a plane-wave expansion cutoff of 400 eV. We used the dispersion-corrected DFT-D3 [24] method for all calculations. Due to the magnetic moment of MgAl2O4, we performed spin-polarized calculations for MgAl2O4 but nonspin-polarized calculations for NiAl2O4. We employed the spinel structures of MgAl2O4 and NiAl2O4 which were dominantly observed by XRD. The PBE bulk lattice constants of MgAl2O4 (a = b = c = 8.16 ,) and NiAl2O4 (a = b = c = 8.14 ) were used to fix the lateral dimensions of MgAl2O4 and NiAl2O4 slabs, respectively. In recent theoretical study using molecular dynamic simulation, the spinel structure of MgAl2O4 favorably has a low index facet of (100). In addition, the (100) surface is predicted to easily undergo surface re-constructions to be more stabilized [25]. However, we only focused our computational studies on the pristine (100) surface for both slabs to explore the intrinsic reactivity of MgAl2O4(100) and NiAl2O4(100) surfaces. Simulated MgAl2O4 (100) and NiAl2O4(100) slabs consisted of 4 layers with two fixed bottom layers, but other layers were allowed to relax until the forces were less than 0.03 eV/Å. Both surfaces of MgAl2O4 (100) and NiAl2O4(100) had coordinatively unsaturated (cus) surface metal sites and cus oxygen atoms. The MgAl2O4(100) surface with Mgcus, Alcus, and Ocus surface atoms is shown in Figure 1a, and the NiAl2O4(100) surface with Nicus, Alcus and Ocus sites is shown in Figure 1b. All computational slab models included a vacuum spacing of~20 Å, which was sufficient to reduce periodic interaction in the surface normal direction. In terms of system size, 2 × 2 (MgAl2O4) and 2 × 2 (NiAl2O4) unit cells with corresponding 2 × 2 × 1 Monkhorst–Pack k-point meshes were employed. Unless otherwise noted, our DFT calculations were performed for a single molecule adsorbed within the surface models, and it corresponds to coverages equal to 25% and 50% of the total density of Alcus metal atoms and cus metal atoms of Mg and Ni, respectively. Although NiO and Ni were also observed from the XRD experiment, we did not investigate Ni and NiO computationally because it has been extensively studied and proposed to be very active toward CO2 hydrogenation [8,9,16].
We defined the adsorption energy between molecule and surface as Equation (1), where E s l a b , E i s o , and E s l a b + a d s are energies of the bare surface, an isolated molecule, and an adsorbed molecule on the bare surface, respectively. A larger positive adsorption energy value indicated high stability of the adsorbed molecule under consideration.
E a d s = E s l a b + E i s o E s l a b + a d s
We evaluated the activation energy barriers for initial steps of methanation on surfaces using the climbing nudged elastic band (cNEB) method [26], and confirmed that the resulting transition states had one imaginary vibrational frequency. All energies reported in this paper are corrections of zero-point vibrational energy.
Projected crystal orbital Hamilton population (pCOHP) has been used to characterize and quantify the interactions between orbitals of CO and Metalcus [27,28,29]. pCOHP provides a measure of the overlap between specific atomic orbitals, and therefore, a relative quantification of bonding. LOBSTER software was used to obtain pCOHP values from VASP outputs [30].

3. Experimental Results

3.1. Characterization of Catalysts

XRD was used to determine the crystalline structures of NixMg1−xAl2O4 catalysts (Figure 2a). For Ni1Mg0Al2O4, peaks were located at 37.3°, 43.3°, and 63°, which corresponded to the crystalline NiO phase. Other peaks near 37.5°, 45.3°, 60.2°, and 66.1° represented the crystalline NiAl2O4 phase. After adding Mg to the Ni1Mg0Al2O4 catalyst to produce Ni0.75Mg0.25Al2O4, peaks related to MgAl2O4 were observed in the XRD pattern and the intensities of NiO-related peaks diminished. Increasing the amount of Mg enhanced MgAl2O4 peak intensities until only MgAl2O4 peaks were observed. XRD patterns indicated that crystalline NiO and NiAl2O4 phases were predominantly formed in Ni-rich catalysts and that the addition of Mg into Ni1Mg0Al2O4 resulted in the formation of a crystalline MgAl2O4 phase rather than NiAl2O4.
The reduction characteristics of NixMg1−xAl2O4 catalysts were evaluated by H2-TPR (Figure 2b). The NiO phase of Ni-included catalysts started to be reduced around 250 °C, i.e., Ni was formed at the temperature used for CO2 hydrogenation experiments. This reduction may have been due to the reduction of NiO to Ni, whereas the band observed at ~600 °C was attributed to the reduction of NiAl2O4 in Ni1Mg0Al2O4 and Ni0.75Mg0.25Al2O4, which corresponded to that reported in a previous study [31], in which almost-complete reductions of NiO and NiAl2O4 were observed below 700 °C. Ni0.5Mg0.5Al2O4 and Ni0.25Mg0.75Al2O4 which had a MgAl2O4 spinel rather than NiAl2O4 showed weak reduction bands between 600 and 750 °C due to the minor NiAl2O4 phase. On the other hand, reduction bands were hardly observed in H2-TPR of Ni0Mg1Al2O4. Based on XRD patterns and TPR curves, we speculated that the crystalline NiAl2O4 and MgAl2O4 structures of NixMg1−xAl2O4 catalysts could not be fully reduced during CO2 hydrogenation at 450 °C.
Textural properties of NixMg1−xAl2O4 catalysts were evaluated using N2 adsorption–desorption isotherms (Figure 3). All catalysts had similar isotherm type V with a hysteresis curve around a relative pressure of 0.8 due to high metal loading and the high calcination temperature used (900 °C). Surface areas, total pore volumes, and average pore sizes were quantified from N2 isotherms (Table 1). Despite the different compositions of NixMg1−xAl2O4 catalysts, textural properties were not noticeably different except for Ni0.75Mg0.25Al2O4, presumably because the same alumina support was used and all were calcined at 900 °C. The larger surface area and total pore volume of Ni0.75Mg0.25Al2O4 might increase the catalytic activity for CO2 hydrogenation. However, for other catalysts, the observation of similar textural properties given different crystalline structures implies that different catalytic activities for CO2 hydrogenation were predominantly due to the different NiAl2O4 and MgAl2O4 crystalline structures of NixMg1−xAl2O4 catalysts.

3.2. CO2 Hydrogenation Activity

CO2 hydrogenation was conducted over NixMg1−xAl2O4 catalysts to examine the different catalytic characteristics of NiAl2O4 and MgAl2O4 spinel structures. More specifically, H2 and CO2 conversions were measured versus temperature over NixMg1−xAl2O4 catalysts (Figure 4a,b). Despite the exothermic nature of CO2 methanation, H2 and CO2 conversion increased with temperature for all catalysts because reactions were conducted using a kinetically controlled regime. Regarding the effects of catalyst compositions, H2 and CO2 conversion tended to increase with the Ni composition. For example, whereas Ni0Mg1Al2O4, which mainly consisted of MgAl2O4, achieved CO2 and H2 conversions of 3.6 and 11.2%, respectively, at 450 °C, Ni1Mg0Al2O4, which was composed of NiO and NiAl2O4, achieved 39.8 and 42.5%, respectively. It has been widely reported that the presence of the metallic Ni, which can be partially reduced from NiO, is an excellent catalyst for CO2 hydrogenation [32]. A number of studies have investigated the catalytic activities of NiAl2O4 and MgAl2O4 for CO2 hydrogenation [16].
To understand the mechanisms responsible for the catalytic effects of NiAl2O4 and MgAl2O4, we calculated molar ratios of H2 to CO2 consumed (Figure 4c). This molar ratio is theoretically four, based on the stoichiometry of CO2 methanation, but one for RWGS. As nickel content increased, the molar ratio of consumed H2 to CO approached four, suggesting that CH4 methanation predominated over Ni and NiAl2O4. In contrast, as the Ni content decreased, the molar ratio approached one, indicating that RWGS predominated over MgAl2O4.
We also measured CO selectivity versus the Mg content of NixMg1−xAl2O4 catalysts (Figure 4d) to confirm the different reaction behaviors of NiAl2O4 and MgAl2O4 (Figure 4c). As was expected, CO selectivity increased with the increasing Mg content. In particular, the CO selectivity of Ni0.5Mg0.5Al2O4 was much greater than that of Ni0.75Mg0.25Al. This might be due to their distinct crystalline structures, whereas in the XRD patterns of Ni0.75Mg0.25Al, NiO and NiAl2O4, MgAl2O4 peaks predominated in the XRD pattern of Ni0.5Mg0.5Al2O4 (Figure 2a). Considering the enhanced CO selectivity of Ni0.75Mg0.25Al2O4 and its distinct MgAl2O4 structure, we speculated that the MgAl2O4 structure promoted RWGS. The detailed reaction mechanisms of MgAl2O4 and NiAl2O4 were further investigated by computational study.

4. Computational Results

4.1. Stabilities of CO2 and CO on MgAl2O4(100) and NiAl2O4(100)

To understand the CO2 methanation mechanism, we evaluated adsorbed CO2 stability on MgAl2O4(100) and NiAl2O4(100). DFT predicted the most favorable configurations of adsorbed CO2 on both surfaces are shown in Figure 5; the corresponding adsorption energies are provided below each configuration in the figure. We also tested other sites and other configurations of CO2, but the configurations shown in Figure 5 were predicted to be the most energetically favorable. Our simulations predict that adsorbed CO2 on Mgcus site is more stable than the CO2 on the Alcus site on the MgAl2O4(100) surface. The linear configuration of CO2 on the Mgcus site transformed to the bent configuration during DFT relaxation. In contrast to the Mgcus site, the linear CO2 weakly bound with an adsorption energy of 5.3 kJ/mol on the Alcus site (not shown), for which the energy barrier of transformation from linear CO2 to bent CO2 was predicted to be negligible. This implies that because the linear CO2 easily transforms to bent CO2 on the Alcus site by overcoming the negligible barrier, the bent CO2 would predominate at these sites. In general, CO2 tends to bind weakly on catalytic surfaces [33,34,35,36,37,38], but our results show that CO2 binds strongly at both cus sites with large adsorption energies. Methanation includes CO formation; therefore, we also investigated CO stability on cus sites of MgAl2O4(100) and NiAl2O4(100). Our simulations predict that on MgAl2O4(100), CO adsorption on the Mgcus site is energetically more favorable than on other sites, but that adsorbed CO is much less stable than adsorbed CO2. These results suggest that adsorbed CO generated from initial C–O bond cleavage in CO2 readily desorbs from the surface of MgAl2O4, which supports its observed high selectivity toward CO during CO2 methanation.
In terms of NiAl2O4(100), our simulation predicts that CO2 binds weakly to its surface with adsorption energies of 15.5 kJ/mol and 36.5 kJ/mol on Alcus and Mgcus sites, respectively (Figure 5). In addition, we tested bent CO2 configurations on both sites, and found that bent CO2 transformed to linear CO2 during DFT relaxations, which is the opposite of that observed for MgAl2O4(100). These results imply that CO2 experiences strong repulsive interactions during adsorption on NiAl2O4(100), which cause weak binding of the adsorbed CO2. In contrast, adsorbed CO binds strongly to NiAl2O4(100). CO strongly and stably interacted with Nicus sites with an adsorption energy of 179.8 kJ/mol, but only interacted weakly on Alcus sites. Accordingly, the order of the predicted stabilities of CO2 and CO on NiAl2O4(100) contrasted with that on MgAl2O4. Overall, our computational results confirm that the experimentally observed higher selectivity toward CO for CO2 methanation on MgAl2O4 stems from the low stability of adsorbed CO, which leads to facile CO desorption during the methanation reaction
During catalytic reactions, surfaces are not pristine because of on-going oxidation and reduction reactions. Thus, catalytic surfaces have edges, kinks, and some reduced sites, and these features have been proposed to be highly active because they are coordinatively more saturated [39,40,41,42]. To consider these effects during CO2 methanation, we assume that the effects of oxygen vacancies (Ov) can be representative of the effects of reduced sites, which would provide initial insights for the effects of reduced sites. We first checked the Ov formation energy to evaluate the thermodynamic feasibility of Ov formation on the (100) surfaces [39]. The predicted Ov formation energies were 589 kJ/mol for MgAl2O4(100) and 414 kJ/mol for NiAl2O4(100). This implies that Ov rarely forms on both surfaces thermodynamically. Despite the less favorable Ov formation, the oxygen vacancies still can be kinetically generated by surface reactions with carbon species from the CO2 methanation process. Therefore, we tested the effects of adjacent oxygen vacancies on the stabilities of CO2 and CO on MgAl2O4(100) and NiAl2O4(100) surfaces by removing a surface oxygen adjacent to adsorbates.
Our simulations predicted that the presence of an oxygen vacancy would significantly stabilize adsorbed CO and CO2 on MgAl2O4(100) and NiAl2O4(100) for most cases (Table 2). On MgAl2O4(100), CO stability on the Alcus site is enhanced with an adjacent oxygen vacancy by ~ 57 kJ/mol, but this is still too low to effectively prevent CO desorption during methanation. In contrast to MgAl2O4, we found that CO on Alcus sites in the presence of Ov is not stable on NiAl2O4; CO on Alcus sites with Ov was found to migrate to Nicus sites during DFT relaxation. On Nicus sites, the presence of Ov improves CO stability by ~ 40 kJ/mol. Based on these results, we conclude that the presence of oxygen vacancies influences the surface stabilities of CO and CO2 on MgAl2O4(100) and NiAl2O4(100), but that the effect of Ov on the stability of CO on MgAl2O4(100) is not enough to secure adsorbed CO during methanation, which suggests that CO selectivity would not be considerably changed by the presence of surface defects.

4.2. Electronic Analysis of the Stability of Adsorbed CO

To determine the reasons for the different stabilities of CO on MgAl2O4, we conducted pCOHP analysis between CO molecules and the cus sites of Mg and Ni [27,28,29]. Integrated pCOHP (IpCOHP), obtained by integrating the pCOHP up to the Fermi level, allowed us to separately evaluate bonding and antibonding contributions to CO stability. Table 3 shows the IpCOHP results for CO–Mgcus, and CO–Nicus. Results show that CO and Nicus have much higher total bonding interactions than CO and Mgcus. For s and p orbitals, weak bonding and strong antibonding interactions, and thus, weak total bonding interactions, were predicted for CO–Mgcus. The CO–Nicus interaction was predicted to involve strong bonding and weak antibonding interactions. In addition, the presence of d orbitals of the Nicus site was found to strongly contribute to the bonding interaction (−3.58), whereas no d orbitals were present at the Mgcus site. We believe that the weaker bonding and stronger antibonding for s and p orbitals and the absence of d orbital is why adsorbed CO on MgAl2O4 tends to desorb from Mgcus site.

4.3. C–O Bond-Breaking Kinetics

In this section, we focus on C–O bond cleavage of CO2 and CO to generate the carbon that subsequently undergoes hydrogenation to form CH4. Energy diagrams of C–O bond cleavages on MgAl2O4(100) and NiAl2O4(100) are shown in Figure 6a. The proposed mechanism is based on the most favorable configurations of CO2 on both surfaces, meaning that CO2 is on the Mgcus site of MgAl2O4 and on the Nicus site of NiAl2O4. C–O bond cleavages of CO2 on the Alcus site were also examined (not shown) but were predicted to be energetically less favorable than cleavages on Mgcus and Nicus sites. Our results show that the overall C–O bond cleavage from CO2 to C is endothermic on pristine MgAl2O4(100) and NiAl2O4(100), and that cleavage requires 715 kJ/mol and 578 kJ/mol for MgAl2O4 and NiAl2O4, respectively. These results indicate that the overall C–O bond cleavage on both substrates requires high temperatures to overcome unfavorable thermodynamics. We also found that C–O bond cleavage of CO2 (CO) on MgAl2O4 had negligible reverse barriers, i.e., CO* and O* (C* and O*) are predicted to readily recombine. The C–O bond cleavage of CO2 on NiAl2O4(100) had a relatively small barrier of 127.2 kJ/mol and a reverse barrier of 26.6 kJ/mol, and C–O bond cleavage of CO also had a negligible reverse barrier, which is similar to that of MgAl2O4. It should be noted that the CO desorption energy is much smaller than the energy required to cleave the C–O bond of CO (30~60 kJ/mol vs. 622 kJ/mol) on MgAl2O4(100), which implies that the CO desorption rate would be much faster than the C–O cleavage rate regardless of prefactors. Again, these results agree well with the experimentally observed high CO selectivity of MgAl2O4. Overall, DFT-predicted large energy barriers for C–O cleavages and negligible reverse barriers on Mgcus and Nicus sites suggest that methanation kinetics via C–O bond cleavage on pristine surfaces are energetically and kinetically unfavorable, and thus, would require high temperatures.
We also evaluated the kinetics of C–O bond cleavage on reduced surfaces of MgAl2O4(100) and NiAl2O4(100) with an oxygen vacancy. Figure 6b provides energy diagrams of C–O bond cleavage on both surfaces. Again, results were obtained using the most stable configurations for CO2 and CO on Mgcus (MgAl2O4) and Nicus (NiAl2O4) sites. The kinetics of C–O bond cleavage in CO2 on the Alcus site was also investigated (details not shown), but the kinetics were found to be energetically less favorable than those at Mgcus and Nicus sites. From the simulations, we found that the oxygen vacancy was healed by an oxygen generated from C–O bond breaking from CO2 to CO on both surfaces and that bond breaking on MgAl2O4 requires much lower energy compared to the NiAl2O4. The CO2 bond cleavage on reduced MgAl2O4(100) was predicted to be exothermic (207.4 kJ/mol), but that on reduced NiAl2O4(100) was predicted to be endothermic (693.6 kJ/mol). Kinetic enhancement by adjacent oxygen vacancies has been reported for transition metal oxides [39,40,43]. In particular, on PdO(101), oxygen vacancies were found to impact CO oxidation and thermal reduction kinetics significantly. Metal atoms adjacent to oxygen vacancies can abstract electrons, which modifies their electronic structures and influences surface reaction kinetics. We found the similar enhancements for MgAl2O4(100) but not for NiAl2O4(100) (Figure 6b). Although a single oxygen vacancy did not enhance the kinetics of overall C–O bond cleavage on NiAl2O4(100), the kinetics were significantly affected. Initial bond cleavage in the presence of adjacent Ov required the more energy than cleavage on a pristine surface, but subsequent bond cleavage in the presence of adjacent Ov required the less energy than the C–O bond cleavage of CO on a pristine surface. Based on these results, we would expect that the presence of surface oxygen vacancies would strongly affect CO2 methanation kinetics on MgAl2O4 and NiAl2O4 surfaces.

4.4. C–O Bond Cleavage of CO vs. C–O Bond Cleavage of CHO

In addition to the C–O bond cleavage of CO2 and CO, many other potential C–O bond-breaking mechanisms may contribute to CH4 generation, and one such mechanism is H-assisted CO2 activations. On Ni-based catalysts, formate (CHOO) and carboxylate (COOH) pathways have been proposed, which provides kinetically and thermodynamically different preferences for the hydrogenation of CO2 [44]. However, in this study, we focus on another potential pathway of C–O bond breaking from CHO on MgAl2O4(100) and NiAl2O4(100). After initial bond cleavage of CO2, the generated CO reacts with adjacent H to form CHO, which undergoes C–O bond cleavage to produce CH. Such reactions would proceed if the kinetics and thermodynamics are more favorable than the other reactions paths. We evaluated the feasibilities of other potential mechanisms of C–O bond cleavage by focusing on the C–O bond cleavage of CHO. Simulations were performed on reduced surfaces because we had found earlier that the reduced surface enhances kinetics and adsorbate stabilities. Energy diagrams of C–O bond cleavage of CO vs. C–O bond cleavage of CHO are provided in Figure 7. Results are based on the most stable configurations of CO and CHO on Mgcus (MgAl2O4) and Nicus (NiAl2O4) sites. The simulation predicted that on NiAl2O4(100), CHO formation from CO and H requires 165.7 kJ/mol, but that the C–O bond cleavage of CHO has a large energy barrier of ~232.9 kJ/mol (Figure 7a). The activation energy required to cleave the C–O bond of CHO is similar to that required to cleave CO, which suggests that on NiAl2O4(100), C–O cleavage in CO and CHO would competitively occur due to the similar overall energies required. Energy diagrams of bond cleavages on reduced MgAl2O4(100) are provided in Figure 7b, which show the different behaviors of C–O bond cleavage. Our results show that CHO formation is exothermic, with 339.9 kJ/mol, and that C–O bond cleavage in CHO has an activation energy of 192.4 kJ/mol and a negligible reverse energy requirement. The overall energy required to active the bond breaking is exothermic, but the C–O bond breaking from CHO is not. These results indicate that CO bond-breaking mechanisms from CHO would occur more readily than direct C–O bond breaking when surface defects are present. Based on the results obtained, we cannot conclusively establish that the C–O bond cleavage of CHO is the dominant pathway on reduced MgAl2O4. Nevertheless, we believe that the alternative C–O cleavage pathways, including C–O bond cleavage in CHO, compete with direct C–O bond cleavage during methanation.

5. Conclusions

Mixtures of nickel–magnesium–alumina with different metal contents were investigated as CO2 hydrogenation catalysts. A range of NixMg1−xAl2O4 spinels were produced (from x = 0 to 1) by co-precipitation. According to our experimental results for CO2 hydrogenation over these catalysts, the reverse water gas shift reaction (CO2 → CO) was preferred at high magnesium atomic ratios, and CO2 methanation (CO2 → CH4) was favored by increasing nickel contents. To understand the kinetics for product selectivity of MgAl2O4 and NiAl2O4, we performed DFT calculations. Our simulation predicted that CO strongly adsorbs on the Nicus site of NiAl2O4; however, CO weakly adsorbs on the Mgcus site of MgAl2O4 due to the weak total binding interactions within s and p orbitals, and the absence of a d orbital. The low stability of adsorbed CO on MgAl2O4 triggers CO desorption rather than undergoing further hydrogenation or C–O dissociation. As a result, the CO selectivity is higher on MgAl2O4 than NiAl2O4. Based on our experimental and computational results, the different reaction mechanisms on NixMg1−xAl2O4 spinels for CO2 hydrogenation were revealed.

Author Contributions

Conceptualization, M.K., D.K. and N.-K.P.; Data curation, B.S. and E.H.K.; Investigation, B.S. and E.H.K.; Methodology, B.S., E.H.K. and J.B.; Supervision, M.K., D.K. and N.-K.P.; Validation B.S. and E.H.K.; Writing—original draft, B.S. and E.H.K.; Writing—review and editing; M.K., D.K. and N.-K.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Basic Science Research Program through the National Re-search Foundation of Korea (NRF) funded by the Ministry of Education (2020R1I1A3051997), and the 2020 Yeungnam University Research Grant.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gielen, D.; Boshell, F.; Saygin, D.; Bazilian, M.D.; Wagner, N.; Gorini, R. The role of renewable energy in the global energy transformation. Energy Strateg. Rev. 2019, 24, 38–50. [Google Scholar] [CrossRef]
  2. Shen, W.; Chen, X.; Qiu, J.; Hayward, J.A.; Sayeef, S.; Osman, P.; Meng, K.; Dong, Z.Y. A comprehensive review of variable renewable energy levelized cost of electricity. Renew. Sustain. Energy Rev. 2020, 133, 110301. [Google Scholar] [CrossRef]
  3. MacDowell, N.; Florin, N.; Buchard, A.; Hallett, J.; Galindo, A.; Jackson, G.; Adjiman, C.S.; Williams, C.K.; Shah, N.; Fennell, P. An overview of CO2 capture technologies. Energy Environ. Sci. 2010, 3, 1645–1669. [Google Scholar] [CrossRef] [Green Version]
  4. Tapia, J.F.D.; Lee, J.-Y.; Ooi, R.E.H.; Foo, D.C.Y.; Tan, R.R. A review of optimization and decision-making models for the planning of CO2 capture, utilization and storage (CCUS) systems. Sustain. Prod. Consum. 2018, 13, 1–15. [Google Scholar] [CrossRef]
  5. Rönsch, S.; Schneider, J.; Matthischke, S.; Schlüter, M.; Götz, M.; Lefebvre, J.; Prabhakaran, P.; Bajohr, S. Review on methanation—From fundamentals to current projects. Fuel 2016, 166, 276–296. [Google Scholar] [CrossRef]
  6. 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]
  7. Hernandez Lalinde, J.A.; Roongruangsree, P.; Ilsemann, J.; Bäumer, M.; Kopyscinski, J. CO2 methanation and reverse water gas shift reaction. Kinetic study based on in situ spatially-resolved measurements. Chem. Eng. J. 2020, 390, 124629. [Google Scholar] [CrossRef]
  8. Mills, G.A.; Steffgen, F.W. Catalytic Methanation. Catal. Rev. 1974, 8, 159–210. [Google Scholar] [CrossRef]
  9. Kirchner, J.; Baysal, Z.; Kureti, S. Activity and Structural Changes of Fe-based Catalysts during CO2 Hydrogenation towards CH4—A Mini Review. ChemCatChem 2020, 12, 981–988. [Google Scholar] [CrossRef] [Green Version]
  10. Wu, H.C.; Chang, Y.C.; Wu, J.H.; Lin, J.H.; Lin, I.K.; Chen, C.S. Methanation of CO2 and reverse water gas shift reactions on Ni/SiO2 catalysts: The influence of particle size on selectivity and reaction pathway. Catal. Sci. Technol. 2015, 5, 4154–4163. [Google Scholar] [CrossRef]
  11. Chen, C.-S.; Budi, C.S.; Wu, H.-C.; Saikia, D.; Kao, H.-M. Size-Tunable Ni Nanoparticles Supported on Surface-Modified, Cage-Type Mesoporous Silica as Highly Active Catalysts for CO2 Hydrogenation. ACS Catal. 2017, 7, 8367–8381. [Google Scholar] [CrossRef]
  12. Aziz, M.A.A.; Jalil, A.A.; Triwahyono, S.; Mukti, R.R.; Taufiq-Yap, Y.H.; Sazegar, M.R. Highly active Ni-promoted mesostructured silica nanoparticles for CO2 methanation. Appl. Catal. B Environ. 2014, 147, 359–368. [Google Scholar] [CrossRef]
  13. Solymosi, F.; Erdöhelyi, A.; Bánsági, T. Methanation of CO2 on supported rhodium catalyst. J. Catal. 1981, 68, 371–382. [Google Scholar] [CrossRef] [Green Version]
  14. Weatherbee, G.D.; Bartholomew, C.H. Hydrogenation of CO2 on group VIII metals: II. Kinetics and mechanism of CO2 hydrogenation on nickel. J. Catal. 1982, 77, 460–472. [Google Scholar] [CrossRef]
  15. Yang Lim, J.; McGregor, J.; Sederman, A.J.; Dennis, J.S. Kinetic studies of CO2 methanation over a Ni/γ-Al2O3 catalyst using a batch reactor. Chem. Eng. Sci. 2016, 141, 28–45. [Google Scholar] [CrossRef] [Green Version]
  16. Kwon, B.C.; Park, N.-K.; Kang, M.; Kang, D.; Seo, M.W.; Lee, D.; Jeon, S.G.; Ryu, H.-J. CO2 hydrogenation activity of Ni-Mg-Al2O3 catalysts: Reaction behavior on NiAl2O4 and MgAl2O4. Korean J. Chem. Eng. 2021, 38, 1188–1196. [Google Scholar] [CrossRef]
  17. Stangeland, K.; Kalai, D.; Li, H.; Yu, Z. CO2 Methanation: The Effect of Catalysts and Reaction Conditions. Energy Procedia 2017, 105, 2022–2027. [Google Scholar] [CrossRef]
  18. Karam, L.; Bacariza, M.C.; Lopes, J.M.; Henriques, C.; Massiani, P.; El Hassan, N. Assessing the potential of xNi-yMg-Al2O3 catalysts prepared by EISA-one-pot synthesis towards CO2 methanation: An overall study. Int. J. Hydrogen Energy 2020, 45, 28626–28639. [Google Scholar] [CrossRef]
  19. Galhardo, T.S.; Braga, A.H.; Arpini, B.H.; Szanyi, J.; Gonçalves, R.V.; Zornio, B.F.; Miranda, C.R.; Rossi, L.M. Optimizing Active Sites for High CO Selectivity during CO2 Hydrogenation over Supported Nickel Catalysts. J. Am. Chem. Soc. 2021, 143, 4268–4280. [Google Scholar] [CrossRef] [PubMed]
  20. Zhang, S.; Ying, M.; Yu, J.; Zhan, W.; Wang, L.; Guo, Y.; Guo, Y. NixAl1O2-δ mesoporous catalysts for dry reforming of methane: The special role of NiAl2O4 spinel phase and its reaction mechanism. Appl. Catal. B Environ. 2021, 291, 120074. [Google Scholar] [CrossRef]
  21. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef] [Green Version]
  22. Kresse, G.; Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B 1993, 47, 558. [Google Scholar] [CrossRef] [PubMed]
  23. Perdew, J.P.; Ernzerhof, M.; Burke, K. Rationale for mixing exact exchange with density functional approximations. J. Chem. Phys. 1996, 105, 9982–9985. [Google Scholar] [CrossRef]
  24. 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] [Green Version]
  25. Hasan, M.M.; Dholabhai, P.P.; Castro, R.H.R.; Uberuaga, B.P. Stabilization of MgAl2O4 spinel surfaces via doping. Surf. Sci. 2016, 649, 138–145. [Google Scholar] [CrossRef] [Green Version]
  26. Henkelman, G.; Uberuaga, B.P.; Jónsson, H. A climbing image nudged elastic band method for finding saddle points and minimum energy paths. J. Chem. Phys. 2000, 113, 9901–9904. [Google Scholar] [CrossRef] [Green Version]
  27. Deringer, V.L.; Tchougréeff, A.L.; Dronskowski, R. Crystal orbital Hamilton population (COHP) analysis as projected from plane-wave basis sets. J. Phys. Chem. A 2011, 115, 5461–5466. [Google Scholar] [CrossRef]
  28. Dronskowski, R.; Blöchl, P.E. Crystal orbital hamilton populations (COHP). Energy-resolved visualization of chemical bonding in solids based on density-functional calculations. J. Phys. Chem. 1993, 97, 8617–8624. [Google Scholar] [CrossRef]
  29. Maintz, S.; Deringer, V.L.; Tchougreeff, A.L.; Dronskowski, R. Analytic projection from plane-wave and PAW wavefunctions and application to chemical-bonding analysis in solids. J. Comput. Chem. 2013, 34, 2557–2567. [Google Scholar] [CrossRef]
  30. Maintz, S.; Deringer, V.L.; Tchougreeff, A.L.; Dronskowski, R. LOBSTER: A tool to extract chemical bonding from plane-wave based DFT. J. Comput. Chem. 2016, 37, 1030–1035. [Google Scholar] [CrossRef] [Green Version]
  31. Ashok, J.; Raju, G.; Reddy, P.S.; Subrahmanyam, M.; Venugopal, A. Catalytic decomposition of CH4 over NiO–Al2O3–SiO2 catalysts: Influence of catalyst preparation conditions on the production of H2. Int. J. Hydrogen Energy 2008, 33, 4809–4818. [Google Scholar] [CrossRef]
  32. Vesselli, E.; Schweicher, J.; Bundhoo, A.; Frennet, A.; Kruse, N. Catalytic CO2 Hydrogenation on Nickel: Novel Insight by Chemical Transient Kinetics. J. Phys. Chem. C 2011, 115, 1255–1260. [Google Scholar] [CrossRef]
  33. Hinojosa, J.A.; Antony, A.; Hakanoglu, C.; Asthagiri, A.; Weaver, J.F. Adsorption of CO2 on a PdO(101) Thin Film. J. Phys. Chem. C 2012, 116, 3007–3016. [Google Scholar] [CrossRef]
  34. Thompson, T.L.; Diwald, O.; Yates, J.T. CO2 as a Probe for Monitoring the Surface Defects on TiO2(110)Temperature-Programmed Desorption. J. Phys. Chem. B 2003, 107, 11700–11704. [Google Scholar] [CrossRef]
  35. Henderson, M.A.; Epling, W.S.; Perkins, C.L.; Peden, C.H.F.; Diebold, U. Interaction of Molecular Oxygen with the Vacuum-Annealed TiO2(110) Surface:  Molecular and Dissociative Channels. J. Phys. Chem. B 1999, 103, 5328–5337. [Google Scholar] [CrossRef]
  36. Wang, Y.; Lafosse, A.; Jacobi, K. Adsorption and Reaction of CO2 on the RuO2(110) Surface. J. Phys. Chem. B 2002, 106, 5476–5482. [Google Scholar] [CrossRef]
  37. Lafosse, A.; Wang, Y.; Jacobi, K. Carbonate formation on the O-enriched RuO2(110) surface. J. Chem. Phys. 2002, 117, 2823–2831. [Google Scholar] [CrossRef]
  38. Seiferth, O.; Wolter, K.; Dillmann, B.; Klivenyi, G.; Freund, H.J.; Scarano, D.; Zecchina, A. IR investigations of CO2 adsorption on chromia surfaces: Cr2O3 (0001)/Cr(110) versus polycrystalline α-Cr2O3. Surf. Sci. 1999, 421, 176–190. [Google Scholar] [CrossRef]
  39. Kim, M.; Pan, L.; Weaver, J.F.; Asthagiri, A. Initial Reduction of the PdO(101) Surface: Role of Oxygen Vacancy Formation Kinetics. J. Phys. Chem. C 2018, 122, 26007–26017. [Google Scholar] [CrossRef]
  40. Zhang, F.; Pan, L.; Li, T.; Diulus, J.T.; Asthagiri, A.; Weaver, J.F. CO Oxidation on PdO(101) during Temperature-Programmed Reaction Spectroscopy: Role of Oxygen Vacancies. J. Phys. Chem. C 2014, 118, 28647–28661. [Google Scholar] [CrossRef]
  41. Van Den Bossche, M.; Grönbeck, H. Methane Oxidation over PdO(101) Revealed by First-Principles Kinetic Modeling. J. Am. Chem. Soc. 2015, 137, 12035–12044. [Google Scholar] [CrossRef]
  42. Chin, Y.C.; Garcia-Dieguez, M.; Iglesia, E. Dynamics and Thermodynamics of Pd−PdO Phase Transitions: Effects of Pd Cluster Size and Kinetic Implications for Catalytic Methane Combustion. J. Phys. Chem. C 2016, 120, 1446–1460. [Google Scholar] [CrossRef]
  43. Schilling, A.C.; Groden, K.; Simonovis, J.P.; Hunt, A.; Hannagan, R.T.; Çlnar, V.; McEwen, J.S.; Sykes, E.C.H.; Waluyo, I. Accelerated Cu2O Reduction by Single Pt Atoms at the Metal-Oxide Interface. ACS Catal. 2020, 10, 4215–4226. [Google Scholar] [CrossRef]
  44. Vogt, C.; Monai, M.; Sterk, E.B.; Palle, J.; Melcherts, A.E.M.; Zijlstra, B.; Groeneveld, E.; Berben, P.H.; Boereboom, J.M.; Hensen, E.J.M.; et al. Understanding carbon dioxide activation and carbon–carbon coupling over nickel. Nat. Commun. 2019, 10, 5330. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Top views of the NiAl2O4(100) and MgAl2O4(100) surfaces. The cus on NiAl2O4(100) and MgAl2O4(100) represent coordinatively unsaturated sites. Red atoms are oxygens and metal atoms are labeled.
Figure 1. Top views of the NiAl2O4(100) and MgAl2O4(100) surfaces. The cus on NiAl2O4(100) and MgAl2O4(100) represent coordinatively unsaturated sites. Red atoms are oxygens and metal atoms are labeled.
Catalysts 11 01026 g001
Figure 2. (a) XRD patterns and (b) H2-TPR curves of NixMg1−xAl2O4 catalysts.
Figure 2. (a) XRD patterns and (b) H2-TPR curves of NixMg1−xAl2O4 catalysts.
Catalysts 11 01026 g002
Figure 3. BET results of NixMg1−xAl2O4 catalysts.
Figure 3. BET results of NixMg1−xAl2O4 catalysts.
Catalysts 11 01026 g003
Figure 4. Results of catalytic activity tests for CO2 hydrogenation over NixMg1−xAl2O4 catalysts, (a) H2 conversion versus temperature, (b) CO2 conversion versus temperature, (c) molar ratio of consumed H2/CO2 versus X for Ni1−xMgxAl catalysts, and (d) CO selectivity versus X for Ni1−xMgxAl catalysts.
Figure 4. Results of catalytic activity tests for CO2 hydrogenation over NixMg1−xAl2O4 catalysts, (a) H2 conversion versus temperature, (b) CO2 conversion versus temperature, (c) molar ratio of consumed H2/CO2 versus X for Ni1−xMgxAl catalysts, and (d) CO selectivity versus X for Ni1−xMgxAl catalysts.
Catalysts 11 01026 g004
Figure 5. Top and side views of adsorbed CO2 and CO on MgAl2O4(100) (ad) and NiAl2O4(100) (eh). DFT-predicted adsorption energies are provided below each figure. Dark grey and purple atoms are carbon of CO and oxygen of CO, respectively. The colors for other atoms are described in Figure 1.
Figure 5. Top and side views of adsorbed CO2 and CO on MgAl2O4(100) (ad) and NiAl2O4(100) (eh). DFT-predicted adsorption energies are provided below each figure. Dark grey and purple atoms are carbon of CO and oxygen of CO, respectively. The colors for other atoms are described in Figure 1.
Catalysts 11 01026 g005
Figure 6. Energy diagrams of C–O bond cleavages for CO2 and CO on (a) pristine and (b) defected surfaces. * represents adsorbed species on the surfaces (e.g., CO2*: adsorbed CO2 on the surface).
Figure 6. Energy diagrams of C–O bond cleavages for CO2 and CO on (a) pristine and (b) defected surfaces. * represents adsorbed species on the surfaces (e.g., CO2*: adsorbed CO2 on the surface).
Catalysts 11 01026 g006
Figure 7. Energy diagrams of the direct C–O bond breaking (CO* → C + O*) vs. indirect C–O bond breaking (CO* + H* → CHO* → CH* + O*) on (a) defected NiAl2O4(100) and (b) defected MgAl2O4(100).
Figure 7. Energy diagrams of the direct C–O bond breaking (CO* → C + O*) vs. indirect C–O bond breaking (CO* + H* → CHO* → CH* + O*) on (a) defected NiAl2O4(100) and (b) defected MgAl2O4(100).
Catalysts 11 01026 g007aCatalysts 11 01026 g007b
Table 1. Textural properties of NixMg1−xAl2O4 catalysts.
Table 1. Textural properties of NixMg1−xAl2O4 catalysts.
SampleSurface Area (m2/g)Total Pore Volume (mL/g)Average Pore Size (nm)
Ni1Mg0Al2O438.00.1710.8
Ni0.75Mg0.25Al2O464.20.2611.4
Ni0.5Mg0.5Al2O438.10.139.4
Ni0.25Mg0.75Al2O432.80.108.4
Ni0Mg1Al2O432.10.108.2
Table 2. Adsorption energies of CO2 and CO on MgAl2O4(100) and NiAl2O4(100) with and without adjacent oxygen vacancies.
Table 2. Adsorption energies of CO2 and CO on MgAl2O4(100) and NiAl2O4(100) with and without adjacent oxygen vacancies.
Adsorbed MoleculeMgAl2O4NiAl2O4
AlcusMgcusAlcusNicus
0 Ov1 Ov0 Ov1 Ov0 Ov1 Ov0 Ov1 Ov
CO30.887.959.757.816.9-179.8233.1
CO275.3187.1162.9138.37.968.336.5−4.6
Table 3. IpCOHP results for CO–Mgcus, and CO–Nicus.
Table 3. IpCOHP results for CO–Mgcus, and CO–Nicus.
Bonding TypesMgAl2O4NiAl2O4
MgcusNicus
s, pds, p, ds, pds, p, d
Bonding−2.08-−2.08−2.68−3.58−6.26
Anti-Bonding0.36-0.360.311.111.42
Total Bonding−1.72-−1.72−2.37−2.47−4.83
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Seo, B.; Ko, E.H.; Boo, J.; Kim, M.; Kang, D.; Park, N.-K. CO2 Hydrogenation on NixMg1−xAl2O4: A Comparative Study of MgAl2O4 and NiAl2O4. Catalysts 2021, 11, 1026. https://doi.org/10.3390/catal11091026

AMA Style

Seo B, Ko EH, Boo J, Kim M, Kang D, Park N-K. CO2 Hydrogenation on NixMg1−xAl2O4: A Comparative Study of MgAl2O4 and NiAl2O4. Catalysts. 2021; 11(9):1026. https://doi.org/10.3390/catal11091026

Chicago/Turabian Style

Seo, Boseok, Eun Hee Ko, Jinho Boo, Minkyu Kim, Dohyung Kang, and No-Kuk Park. 2021. "CO2 Hydrogenation on NixMg1−xAl2O4: A Comparative Study of MgAl2O4 and NiAl2O4" Catalysts 11, no. 9: 1026. https://doi.org/10.3390/catal11091026

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