Next Article in Journal
Modified 7-Chloro-11H-indeno[1,2-b]quinoxaline Heterocyclic System for Biological Activities
Next Article in Special Issue
Gas-Phase Selective Oxidation of Methane into Methane Oxygenates
Previous Article in Journal
Fabrication of Pt-Loaded Catalysts Supported on the Functionalized Pyrolytic Activated Carbon Derived from Waste Tires for the High Performance Dehydrogenation of Methylcyclohexane and Hydrogen Production
Previous Article in Special Issue
Photoelectrochemical Conversion of Methane into Value-Added Products
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

CO and CO2 Methanation over CeO2-Supported Cobalt Catalysts

1
Department of Energy System Research, Ajou University, 206, World-Cup-Ro, Yeongtong-Gu, Suwon 16499, Korea
2
Department of Chemical Engineering, Ajou University, 206, World-Cup-Ro, Yeongtong-Gu, Suwon 16499, Korea
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(2), 212; https://doi.org/10.3390/catal12020212
Submission received: 28 December 2021 / Revised: 31 January 2022 / Accepted: 9 February 2022 / Published: 11 February 2022

Abstract

:
CO2 methanation is a promising reaction for utilizing CO2 using hydrogen generated by renewable energy. In this study, CO and CO2 methanation were examined over ceria-supported cobalt catalysts with low cobalt contents. The catalysts were prepared using a wet impregnation and co-precipitation method and pretreated at different temperatures. These preparation variables affected the catalytic performance as well as the physicochemical properties. These properties were characterized using various techniques including N2 physisorption, X-ray diffraction, H2 chemisorption, temperature-programmed reduction with H2, and temperature-programmed desorption after CO2 chemisorption. Among the prepared catalysts, the ceria-supported cobalt catalyst that was prepared using a wet impregnation method calcined in air at 500 °C, and reduced in H2 at 500 °C, showed the best catalytic performance. It is closely related to the large catalytically active surface area, large surface area, and large number of basic sites. The in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) study revealed the presence of carbonate, bicarbonate, formate, and CO on metallic cobalt.

1. Introduction

CO2 capture, utilization, and storage have become increasingly important in terms of mitigating global climate change caused by escalation in CO2 atmospheric concentrations since the first industrial revolution. CO2 utilization can be categorized into physical and chemical methods. The latter is preferred over the former as CO2 can be converted into other chemicals through chemical transformations [1]. Among the various chemical utilization methods of CO2, CO2 hydrogenation has attracted significant attention as it can produce various platform chemicals, including methane, methanol, and formic acid. It can also utilize hydrogen generated from water with a water electrolysis system powered by renewable energy [2,3,4].
In particular, CO2 methanation is considered an energy storage alternative to address variability in renewable energy as the produced methane can be transported through a gas grid. Furthermore, CO2 methanation (Reaction 1) is thermodynamically more feasible than other CO2 hydrogenation reactions (e.g., reverse water-gas-shift reaction (Reaction 2) and methanol synthesis (Reaction 3)) [5,6,7].
Reaction 1:
CO 2 ( g ) + 4 H 2 ( g )     CH 4 ( g ) + 2 H 2 O ( g )   Δ G 298   K o = 113.2   kJ / mol ,   Δ H 298   K o = 165   kJ / mol
Reaction 2:
CO 2 ( g ) + H 2 ( g )     CO ( g ) + H 2 O ( g )   Δ G 298   K o = 28.6   kJ / mol ,   Δ H 298   K o = 41.2   kJ / mol
Reaction 3:
CO 2 ( g ) + 3 H 2 ( g )     CH 3 OH ( g ) + H 2 O ( g )   Δ G 298   K o = 3.5   kJ / mol ,   Δ H 298   K o = 49.5   kJ / mol
Since CO2 methanation is an exothermic and thermodynamically limited reaction, active catalysts at low temperatures should be developed. To date, some metal catalysts, including Ru [8,9,10,11], Ni [12,13,14,15,16,17,18,19], Co [20,21,22,23,24,25,26], and Fe [23,27,28,29] have been reported to be active in this reaction. However, Ni-based catalysts have been more thoroughly investigated due to their high activity and comparatively low cost [3,4]. In addition to the active metal, the nature of the support also affects the catalytic properties, such as the morphology of the active metal, metal dispersion, and reducibility of active metal/metal oxides [1,2,7]. Several metal oxide supports have been reported for CO2 methanation reactions, such as Al2O3 [30,31,32], CeO2 [33,34,35], SiO2 [36,37,38], ZrO2 [39,40,41,42], TiO2 [43,44,45], La2O3 [45,46,47,48], Y2O3 [31,49], and Sm2O3 [50,51]. Among them, CeO2 is a well-known support due to its redox properties and high surface oxygen vacancies. In particular, two oxidation states of cerium (e.g., Ce3+ and Ce4+) are interconvertible between CeO2 and CeO2−x under oxidized and reduced environments to release oxygen vacancy [52,53,54,55]. This is favorable for the dissociation of CO2 into CO and O species on the catalyst surface in the dissociative CO2 methanation mechanisms [52,53,54]. The other representative CO2 methanation mechanism is the associative one, in which the chemisorbed CO2 (CO2 ads) and chemisorbed H2 (H2 ads) species are found on the support surface and the active sites, respectively. In this CO2 associative reaction mechanism, several intermediates such as carbonate, bicarbonate, and formate are observed [1].
Additionally, catalytic performance is influenced by the catalyst preparation methods, such as precipitation [56], co-precipitation [57], wet impregnation [21,58], and co-wet impregnation [59]. The wet-impregnation method is a well-known traditional method that is simple and economical. However, this generally provides a weaker metal-support interaction than the co-precipitation method [60]. Pretreatment conditions are also key factors in controlling the structural properties of the catalysts [61,62].
In this study, cobalt and ceria were chosen as the active metal and support, respectively. Although cobalt is more expensive than nickel, cobalt is more stable under low-temperature reaction conditions than nickel as using nickel can result in the formation of unstable nickel carbonyl complexes in the presence of CO, especially at low temperatures. The cobalt content was fixed to low enough so that the catalyst price could be comparable with those of typical nickel-based catalysts with high nickel content (higher than 10 wt.%). Two preparation methods, wet impregnation and co-precipitation, were conducted in this study. The effect of pretreatment temperature on the catalytic performance was also investigated.

2. Results and Discussion

2.1. Physicochemical Properties of the Prepared Catalysts

The textural properties of the prepared catalysts were determined using N2 physisorption. Figure S1 shows the N2 adsorption and desorption isotherms of each catalyst. The impregnated samples (Co/CeO2) show a Type IV physisorption isotherm with an H2(a) hysteresis loop. The co-precipitated catalysts (Co0.1Ce0.9Ox) exhibited a Type III physisorption isotherm with an H3 hysteresis loop. As listed in Table 1, the Brunauer-Emmett-Teller (BET) surface area of the catalyst decreased, however the average pore diameter of the catalyst increased with increasing calcination temperature, irrespective of the preparation method.
The cobalt contents of the impregnated and co-precipitated samples were determined to be 5.3 and 3.3 wt.%, respectively. The catalytically active surface area (CASA) of each catalyst was determined by H2 chemisorption. Interestingly, the CASA and cobalt dispersions show a volcano plot as a function of the calcination temperature, irrespective of the preparation method. Co0.1Ce0.9Ox(500) and Co/CeO2(500) have the largest CASA among the catalysts prepared using the co-precipitation and wet impregnation methods, respectively. The number of surface basic sites was measured using pulsed CO2 chemisorption. Of the co-precipitated catalysts, the largest number of surface basic sites was obtained for Co0.1Ce0.9Ox(500). The number of surface basic sites decreased with increasing calcination temperature for the Co/CeO2 catalysts prepared using the wet impregnation method.
The bulk crystalline structures of the ceria-supported cobalt catalysts before and after reduction were probed using X-ray diffraction (XRD). Figure S2 reveals that the XRD peaks due to CeO2 (JCPDS 34-0394) are dominant for the samples before reduction. The crystallite size of CeO2 for the co-precipitated sample increased from 6.5 to 17.7 nm with increasing calcination temperature from 300 to 700 °C, respectively. Similarly, the crystallite size of CeO2 for the impregnated samples also increased from 6.7 to 12.1 nm with increasing calcination temperature from 300 to 700 °C, respectively. As a result of the low cobalt content, the XRD peaks due to Co3O4 (JCPDS 42-1467) were found only for samples calcined at 700 °C, irrespective of the preparation method.
Temperature-programmed reduction with H2 (H2-TPR) was conducted to determine the reducibility of metal oxides in each catalyst. The H2-TPR peaks can be divided into two regions: low-temperature and high-temperature. The low-temperature H2-TPR peaks range from 100 to 600 °C. These H2-TPR peaks are related to the reduction in cobalt oxides with different sizes and interactions with ceria. Generally, smaller metal oxides with weak interactions with a support can be reduced more easily at lower temperatures. Conversely, metal oxides with a strong interaction with a support can be reduced at high temperatures, even though the size of metal oxides is small. In the case of Co3O4, a two-step reduction process from Co3O4 to cobalt through CoO is well known [63]. The high-temperature H2-TPR peak above 600 °C was ascribed to the reduction in surface ceria. As shown in Figure 1, the Co/CeO2 catalysts prepared using the wet impregnation method have H2-TPR peaks at lower temperatures than the Co0.1Ce0.9Ox catalysts prepared using a co-precipitation method. This indicates that the co-precipitation method can provide a stronger interaction between cobalt oxides and ceria than the wet impregnation method. It is noteworthy that the H2-TPR peak position moves toward a high temperature with increasing calcination temperature, irrespective of the preparation method. This reveals that a stronger interaction between cobalt oxides and ceria can be achieved at higher calcination temperatures.
Temperature-programmed desorption after CO2 chemisorption (CO2-TPD) was performed to probe the surface basicity of each catalyst. As shown in Figure 2, there are two desorption peaks. The first occurs at temperatures less than 200 °C, and the second occurs at temperatures ranging from 200–600 °C. The low-temperature CO2-TPD peaks originate from weakly adsorbed CO2 on the catalyst surface, while the high-temperature CO2-TPD peaks indicate the strong adsorption of CO2 onto the catalyst surface. The CO2-TPD peak observed at higher temperatures implies stronger chemisorption of CO2 onto the catalyst surface. The Co0.1Ce0.9Ox(300) catalyst presents no low-temperature CO2-TPD peak, although it has two CO2-TPD peaks at high temperatures. However, all other catalysts exhibited low-temperature and high-temperature CO2-TPD peaks. These two CO2-TPD peaks were quantified, and the amounts of desorbed CO2 for each catalyst are listed in Table S1. The amount of strongly chemisorbed CO2 decreased with increasing calcination temperature for the co-precipitated catalyst (Co0.1Ce0.9Ox). The smallest amount of strongly chemisorbed CO2 was obtained for the Co0.1Ce0.9Ox(700) catalyst. Similarly, the amount of strongly chemisorbed CO2 decreased with increasing calcination temperature for the impregnated catalysts (Co/CeO2). The smallest amount of strongly chemisorbed CO2 was obtained for the Co/CeO2(700) catalyst.

2.2. CO Methanation

The catalytic activity for CO methanation was evaluated using ceria-supported Co catalysts. As shown in Figure 3, the Co/CeO2(500) catalyst exhibited the highest CO conversion at low temperatures of the tested catalysts, whereas Co0.1Ce0.9Ox(700) showed the lowest CO conversion even at high temperatures. As shown in Figures S3 and S4, CH4 is the main product, and a small amount of C2H6 is also detected as a byproduct. Separately, the specific reaction rate of CO methanation was also obtained under a kinetically controlled regime. The specific reaction rate decreased in the following order: Co/CeO2(500) > Co/CeO2(300)~Co0.1Ce0.9Ox(300) > Co0.1Ce0.9Ox(500)~Co/CeO2(700) > Co0.1Ce0.9Ox(700). This ranking order in the specific reaction rate is closely related to that of CASA (Table 1) for Co/CeO2 catalysts. This implies that the catalytic activity for CO methanation is directly related to the catalytically active surface area.
The apparent activation energy (Ea) for CO methanation, calculated from the Arrhenius plot in Figure S5 over Co/CeO2(300), Co/CeO2(500), Co/CeO2(700), Co0.1Ce0.9Ox(300), Co0.1Ce0.9Ox(500), and Co0.1Ce0.9Ox(700) were 128, 107, 129, 128, 146, and 162 kJ/mol, respectively. It is worth mentioning that the lowest apparent activation energy is obtained for the catalyst with the largest CASA among Co/CeO2 catalysts. The fact that the co-precipitated Co0.1Ce0.9Ox catalysts have higher apparent activation energies than the impregnated Co/CeO2 catalysts seems to be related to the H2-TPR results that the co-precipitation method can provide a stronger interaction between cobalt oxides and ceria than the wet impregnation method. Castillo et al. [64] reported that the CO consumption turnover rate was significantly lower for smaller cobalt nanoparticles during CO methanation over Co/SiO2 catalysts with different cobalt particle sizes ranging from 4 to 33 nm. The apparent activation energy was lower for smaller cobalt nanoparticles. The apparent activation energies for CO methanation over 10 wt.% Ni/CeO2 and 58 wt.% Ni-CeO2 were reported to be 115 and 133 kJ/mol, respectively [65].

2.3. CO2 Methanation

The catalytic activity for CO2 methanation was examined using ceria-supported cobalt catalysts. As shown in Figure 4, the tested catalysts are classified into three groups based on their catalytic activity for CO2 methanation. The most active catalyst groups include Co/CeO2(300), Co/CeO2(500), and Co/CeO2(700). The Co0.1Ce0.9Ox(300) catalyst was inferior to the most active catalyst group in this reaction. However, it was better than the least active catalyst groups (Co0.1Ce0.9Ox(500) and Co0.1Ce0.9Ox(700)). Unlike CO methanation, the catalytic activity for CO2 methanation over ceria-supported cobalt catalysts tested in this study was not directly correlated with that of CASA (Table 1). In addition to CASA, the amount of moderate basic sites has been reported to be an additional factor affecting the catalytic activity for CO2 methanation [66]. It is worth mentioning that the least active catalyst (Co0.1Ce0.9Ox(700)) had the lowest CO2 uptake among the prepared catalysts. Methane was produced as the main product, and CO was also detected as a byproduct due to the reversed water-gas-shift side reaction (Figures S6 and S7). The apparent Ea for CO2 methanation calculated from the Arrhenius plot in Figure S8 over Co/CeO2(300), Co/CeO2(500), Co/CeO2(700), Co0.1Ce0.9Ox(300), Co0.1Ce0.9Ox(500), and Co0.1Ce0.9Ox(700) were 81, 77, 86, 81, 84, and 86 kJ/mol, respectively. Interestingly, all catalysts have similar apparent activation energies. The reported activation energies for CO2 methanation over nickel-based catalysts range between 55 and 106 kJ/mol [47,67]. The apparent activation energies for CO2 methanation over 58 wt.% Ni-CeO2 were reported to be 95 kJ/mol [65]. Xu et al. [41] reported a higher activation energy for CO2 methanation over Ni/ZrO2 (99.7 kJ/mol) than over Ni/Y0.1Zr0.9Ox (84.2 kJ/mol). Furthermore, the activation energy of the Co/Ce0.8Zr0.2O catalyst was reported to be less than that of Ni/Ce0.8Zr0.2O catalysts [68].

2.4. In Situ DRIFTS Study

To probe the surface species during CO2 adsorption and methanation, in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) study was conducted over the most active catalysts selected from co-precipitated and impregnated catalysts (e.g., Co0.1Ce0.9Ox(300) and Co/CeO2(500)). The molecularly absorbed CO2 was observed at ν = 2400-2200 cm−1 [69]. CO2 adsorption was initially performed at 40 °C (Figure 5). The typical IR peaks observed in this study are listed in Table 2. Bicarbonate and carbonate species were detected during CO2 adsorption. For the bicarbonate intermediate species, asymmetric and symmetric stretching of OCO (νasy(OCO) and νs(OCO)) were observed at ν = 1602 and 1408 cm−1 over Co0.1Ce0.9Ox(300). The same bands were observed at ν = 1654 and 1434 cm−1 over Co/CeO2(500). In addition, COH stretching for bicarbonate species (HCO3) clearly appeared at ν = 1217 cm−1 over Co0.1Ce0.9Ox(300), while it was not observed for Co/CeO2(500). For the carbonate species, monodentate and bidentate were confirmed over Co/CeO2(500) at νasy(OCO) = 1475 and 1577 cm−1 and νs(OCO) = 1380, 1331, and 1265 cm−1, while the symmetric OCO peak centered at ν = 1293 cm−1 was detected for Co0.1Ce0.9Ox(300). During CO2 adsorption at 300 °C (Figures S9 and S10), the bicarbonate and carbonate species also occurred at the same position as at 40 °C. However, the noticeable difference during CO2 adsorption at low and high temperatures is the weak and broad peaks of adsorbed CO* species on the metallic cobalt at ν = 2184, 2142, 2045, and 2119 cm−1 over Co0.1Ce0.9Ox(300) and Co/CeO2(500). This implies that the C=O double bond in the CO2 molecule can be broken down to generate CO* intermediates at high temperatures.
The in situ DRIFTS study was also conducted during CO2 methanation with increasing reaction temperature. Figure 6 and Figure 7 show the IR spectra of Co0.1Ce0.9Ox(300) and Co/CeO2(500), respectively. The observed IR bands are summarized in Table 2. The sharp IR peak at ν = 1216 cm−1, which can be ascribed to COH stretching at temperatures below 300 °C, is responsible for the bicarbonate species observed over both Co0.1Ce0.9Ox(300) and Co/CeO2(500). Furthermore, carbonate species were found in the monodentate and bidentate forms, whereas the OCO and CO vibrations were obtained in the range of ν = 1575 and ν = 1075 cm−1. In particular, the specific formate species, which are common intermediates during CO2 hydrogenation, were also found in the stretching of CH and OCO at the high-frequency adsorption band. The vibrations at ν = 2944–2713 cm−1 are ascribed to CH stretching. The weak vibration of adsorbed CO* species on metallic cobalt was also detected at ν = 2115 and 2075 cm−1 over Co0.1Ce0.9Ox(300), while Con+-CO bridged species were only observed at ν = 2115 cm−1 over Co/CeO2(500). In addition, methane adsorbed on the catalyst surface was detected at ν = 1305 and 3014 cm−1. Based on the IR study, it can be proposed that the CO2 associative pathway is the main route as a result of the predominant appearance of bicarbonate, carbonate, and formate intermediates. However, the CO2 dissociative route cannot be excluded to produce methane over ceria-supported cobalt catalysts.

3. Experimental Sections

3.1. Materials

All chemicals, including cobalt (II) nitrate hexahydrate (Junsei Chemical Co. Ltd., Tokyo, Japan), cerium (III) nitrate hexahydrate (Junsei Chemical Co. Ltd., Tokyo, Japan), and ceria (Rodia, SBET = 250 m2/g), were reagent grade and used as received.

3.2. Preparation of the Catalysts

Co0.1Ce0.9Ox catalysts were prepared using the co-precipitation method. The calculated amount of each precursor was dissolved in distilled (DI) water to achieve 0.1 M total metal precursor solution. The pH was then increased by adding 0.5 M NH4OH aqueous solution dropwise to reach a final pH of nine with vigorous stirring. The slurry was aged at room temperature for 12 h, filtered, and washed several times with DI water. The recovered solid cake was dried at 110 °C overnight under vacuum conditions. The dried sample was further calcined in air at different temperatures and finally reduced in H2 at 500 °C.
For comparison, the Co/CeO2 catalysts were prepared using a wet impregnation method. Cobalt precursor (1.30 g) was dissolved in 50.0 mL DI water in a round flask. Ceria (5.00 g) was added to the cobalt precursor solution. This slurry was mixed for 6 h at 60 °C in a rotary evaporator (BUCHI Labortechnik AG, Flawil, Switzerland), and then the excess water was evaporated while maintaining low pressure. The recovered powder was dried at 110 °C overnight under vacuum conditions. The dried sample was further calcined in air at different temperatures and finally reduced in H2 at 500 °C.
The calcination temperature was included in the sample name to differentiate each catalyst calcined at different temperatures. For example, Co0.1Ce0.9Ox(300) and Co/CeO2(300) are catalysts prepared by co-precipitation and wet impregnation, respectively. These catalysts were calcined in air at 300 °C.

3.3. Catalyst Characterizations

N2 physisorption was performed at −196 °C on a Micromeritics ASAP 2020 (Micromeritics Ltd.), surface area and pore size analyzer after degassing at 200 °C under vacuum for 6 h. The specific surface area and pore size distribution were determined using the BET and Barret-Joyner-Halenda desorption methods, respectively.
XRD patterns were measured using a Rigaku D/Max instrument with a Cu Kα source. The crystallite size of ceria was determined using the Scherrer Formula (1):
d = K λ β cos ( θ )
where d is the average particle size (nm), K is the dimensionless shape factor (0.9), λ is the wavelength of the X-ray radiation (0.15406 nm), β is the full width at half maximum (FWHM) of a peak in rad, and θ is the Bragg angle.
The temperature-programmed reduction with H2 (H2-TPR) was conducted using a Micromeritics 2920 Autochem instrument. A sample of 100 mg was loaded in a quartz tube, and the temperature was increased from room temperature (RT) to 900 °C at a heating rate of 10 °C/min under a flow of 10 mol% H2/Ar.
The H2-chemisorption was performed using a Micromeritics ASAP 2020 instrument (Micromeritics Ltd.), to determine the metal dispersion and CASA. A sample of 200 mg was placed in a quartz tube, reduced in H2 at 500 °C for 1 h, cooled down to room temperature, and typical H2-chemisorption was conducted.
Pulsed CO2 chemisorption and temperature-programmed desorption after CO2 chemisorption (CO2-TPD) were conducted to measure the amount of basic sites and adsorption strength of CO2 in each sample, respectively. They were performed using a Micromeritics 2920 Autochem instrument. A sample of 100 mg was loaded into a quartz tube, reduced under 10 mol% H2/Ar flow at 500 °C for 1 h, cooled to room temperature, and received pulsed injection of CO2. After saturation, CO2 desorption was conducted using a He steam flow rate of 30 mL/min. The mass ion signals recorded at m/e = 44, m/e = 18, and m/e = 16 were monitored to detect desorbed CO2, H2O, and CH4, respectively.
The metal content in each sample was determined by inductively coupled plasma-optical emission spectroscopy (ICP-OES) using a Thermo Scientific iCAP 6500 instrument.
The evolution of surface species during CO2 adsorption and CO2 hydrogenation was probed with in situ DRIFTS on a NICOLET 6700 (Thermal Scientific, Waltham, MA, USA) spectrometer with a ZnSe window at a resolution of 3.857 cm−1. The sample was initially reduced in situ in the DRIFTS cell at 500 °C for 1 h under a H2 flow rate of 20 mL/min and then cooled to the desired temperature. The cell was purged and stabilized with He flow for 20 min. The background was then collected under a He flow. The CO2 adsorptions at 40 and 300 °C were separately performed under a flow of 10% CO2/He, and the spectra were recorded simultaneously. In addition, CO2 hydrogenation was examined at different temperatures under the same operating conditions mentioned above, and the spectra were collected after 15 min.

3.4. Catalytic Performance

The catalyst (100 mg) was placed in a fixed-bed reactor (internal diameter = 3 mm and length = 345 mm) and reduced under H2 flow at a flow rate of 30 mL/min for 1 h. The reactor was then cooled to the desired temperature, and the catalyst was placed in contact with a feed composed of 1 mol% CO (or CO2), 49 mol% He, and 50 mol% H2 at a flow rate of 100 mL/min. A mass flow rate controller (MFC) (Brooks Instrument) was used to control the flow rate of each gas.
The kinetic experiment was conducted separately using 80 mg of the catalyst and 200 mg of α-alumina powder as a diluent. The CO (or CO2) conversion was controlled to be less than 20%, and the Ea was calculated using the Arrhenius Equation (2):
k = Aexp ( E a RT )
where k is the reaction rate constant, A is the frequency factor, Ea is the apparent activation energy, R is the gas constant, and T is the temperature.
The effluent gas was separated using a packed column filled with carbonate for thermal conductive detector (TCD) and a capillary Poralot Q column for the flame ionization detector (FID) in a gas chromatograph (YL Instrument 6100GC). The CO (or CO2) conversion ( X CO or X CO 2 , %) and the product yields, including CH4 and CO ( Y CH 4 and Y CO , %) were determined according to the carbon balance as follows:
X CO = C CO   input C CO   output C CO   input × 100 %
X CO 2 = C CO 2   input C CO 2   output C CO 2   input × 100 %
Y CH 4 = C CH 4   output C CO   input +   C CO 2   input × 100 %
Y CO = C CO   output C CO 2   input × 100 %
The input and output concentrations of each gas (i) are denoted as C i   input and C i   output , respectively.

4. Conclusions

The catalytic activity for CO and CO2 methanation was dependent on the preparation method for ceria-supported cobalt catalysts with low cobalt contents. Stronger interactions between cobalt oxides and ceria were observed for the co-precipitated catalysts than for wet-impregnated interactions. A moderate calcination temperature is plausible for achieving a high CASA. The total number of basic sites decreased significantly with an increasing calcination temperature of 500 to 700 °C for the co-precipitated and impregnated catalysts. CO methanation activity appears to be closely related to CASA, and the CASA and the number of strong basic sites affect the catalytic activity for CO2 methanation. According to the in situ DRIFTS results, CO2 methanation proceeds mainly through the CO2 associative pathway, in which carbonate, bicarbonate, and formate intermediates are detected. Additionally, the CO2 dissociative route worked simultaneously as we also observed the adsorbed CO species on the metallic cobalt surface sites.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12020212/s1, Figure S1: Nitrogen adsorption and desorption isotherms (A,B) and pore size distribution (C) of ceria-supported cobalt catalysts calcined in air at different temperatures. Figure S2: XRD patterns of ceria-supported cobalt catalysts calcined in air at different temperatures (A) and after reduction in H2 at 500 °C (B). Figure S3: Carbon selectivities for CH4 and C2H6 for CO methanation over ceria-supported cobalt catalysts. Figure S4: Product yields for CO methanation over ceria-supported cobalt catalysts. Figure S5: Arrhenius plot for CO methanation over ceria-supported cobalt catalysts. Figure S6: Carbon selectivities for CH4 and CO for CO2 methanation over ceria-supported cobalt catalysts. Figure S7: Product yields for CO2 methanation over ceria-supported cobalt catalysts. Figure S8: Arrhenius plot for CO2 methanation over ceria-supported cobalt catalysts. Figure S9. In situ DRIFTS spectra after adsorption of CO2 on the Co0.1Ce0.9Ox(300) catalyst at 300 °C. The feed was composed of 33 mol% CO2 and 67 mol% He. Figure S10. In situ DRIFTS spectra after adsorption of CO2 on the Co/CeO2(500) catalyst at 300 °C. The feed was composed of 10 mol% CO2 and 90 mol% He. Table S1: Quantitative analysis of the CO2-TPD data.

Author Contributions

Experimental investigation and data analysis, T.H.N.; formal analysis, H.B.K.; writing—original draft preparation, T.H.N.; writing—review and editing, supervision, project administration, and funding acquisition, E.D.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the C1 Gas Refinery Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Science and ICT (2015M3D3A1A01064899).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nguyen, T.T.H.; Park, E.D. A mini-review on power-to-methane catalysts. J. Energy Eng. 2021, 30, 10–25. [Google Scholar]
  2. Sreedhar, I.; Varun, Y.; Singh, S.A.; Venugopal, A.; Reddy, B.M. Developmental trends in CO2 methanation using various catalysts. Catal. Sci. Technol. 2019, 9, 4478–4504. [Google Scholar] [CrossRef]
  3. Younas, M.; Loong Kong, L.; Bashir, M.J.K.; Nadeem, H.; Shehzad, A.; Sethupathi, S. Recent advancements, fundamental challenges, and opportunities in catalytic methanation of CO2. Energy Fuels 2016, 30, 8815–8831. [Google Scholar] [CrossRef]
  4. Li, W.; Wang, H.; Jiang, X.; Zhu, J.; Liu, Z.; Guo, X.; Song, C. A short review of recent advances in CO2 hydrogenation to hydrocarbons over heterogeneous catalysts. RSC Adv. 2018, 8, 7651–7669. [Google Scholar] [CrossRef] [Green Version]
  5. Ashok, J.; Pati, S.; Hongmanorom, P.; Tianxi, Z.; Junmei, C.; Kawi, S. A review of recent catalyst advances in CO2 methanation processes. Catal. Today 2020, 356, 471–489. [Google Scholar] [CrossRef]
  6. Bailera, M.; Lisbona, P.; Romeo, L.M.; Espatolero, S. Power to gas projects review: Lab, pilot and demo plants for storing renewable energy and CO2. Renew. Sustain. Energy Rev. 2017, 69, 292–312. [Google Scholar] [CrossRef]
  7. Gao, J.; Liu, Q.; Gu, F.; Liu, B.; Zhong, Z.; Su, F. Recent advances in methanation catalysts for the production of synthetic natural gas. RSC Adv. 2015, 5, 22759–22776. [Google Scholar] [CrossRef]
  8. Garbarino, G.; Bellotti, D.; Riani, P.; Magistri, L.; Busca, G. Methanation of carbon dioxide on Ru/Al2O3 and Ni/Al2O3 catalysts at atmospheric pressure: Catalysts activation, behaviour and stability. Int. J. Hydrog. Energy 2015, 40, 9171–9182. [Google Scholar] [CrossRef]
  9. Kim, A.; Sanchez, C.; Haye, B.; Boissière, C.; Sassoye, C.; Debecker, D.P. Mesoporous TiO2 support materials for Ru-based CO2 methanation catalysts. ACS Appl. Nano Mater. 2019, 2, 3220–3230. [Google Scholar] [CrossRef]
  10. Chen, S.; Abdel-Mageed, A.M.; Dyballa, M.; Parlinska-Wojtan, M.; Bansmann, J.; Pollastri, S.; Olivi, L.; Aquilanti, G.; Behm, R.J. Raising the COx methanation activity of a Ru/gamma-Al2O3 catalyst by activated modification of metal-support interactions. Angew. Chem. Int. Ed. Eng. 2020, 59, 22763–22770. [Google Scholar] [CrossRef]
  11. Quindimil, A.; de la Torre, U.; Pereda-Ayo, B.; Davó-Quiñonero, A.; Bailón-García, E.; Lozano-Castelló, D.; González-Marcos, J.A.; Bueno-López, A.; González-Velasco, J.R. Effect of metal loading on the CO2 methanation: A comparison between alumina supported Ni and Ru catalysts. Catal. Today 2020, 356, 419–432. [Google Scholar] [CrossRef]
  12. Tan, J.; Wang, J.; Zhang, Z.; Ma, Z.; Wang, L.; Liu, Y. Highly dispersed and stable Ni nanoparticles confined by MgO on ZrO2 for CO2 methanation. Appl. Surf. Sci. 2019, 481, 1538–1548. [Google Scholar] [CrossRef]
  13. Bacariza, M.C.; Amjad, S.; Teixeira, P.; Lopes, J.M.; Henriques, C. Boosting Ni Dispersion on zeolite-supported catalysts for CO2 methanation: The influence of the impregnation solvent. Energy Fuels 2020, 34, 14656–14666. [Google Scholar] [CrossRef]
  14. Guilera, J.; del Valle, J.; Alarcón, A.; Díaz, J.A.; Andreu, T. Metal-oxide promoted Ni/Al2O3 as CO2 methanation micro-size catalysts. J. CO2 Util. 2019, 30, 11–17. [Google Scholar] [CrossRef]
  15. Song, F.; Zhong, Q.; Yu, Y.; Shi, M.; Wu, Y.; Hu, J.; Song, Y. Obtaining well-dispersed Ni/Al2O3 catalyst for CO2 methanation with a microwave-assisted method. Int. J. Hydrog. Energy 2017, 42, 4174–4183. [Google Scholar] [CrossRef]
  16. Daroughegi, R.; Meshkani, F.; Rezaei, M. Enhanced activity of CO2 methanation over mesoporous nanocrystalline Ni–Al2O3 catalysts prepared by ultrasound-assisted co-precipitation method. Int. J. Hydrog. Energy 2017, 42, 15115–15125. [Google Scholar] [CrossRef]
  17. Kesavan, J.K.; Luisetto, I.; Tuti, S.; Meneghini, C.; Iucci, G.; Battocchio, C.; Mobilio, S.; Casciardi, S.; Sisto, R. Nickel supported on YSZ: The effect of Ni particle size on the catalytic activity for CO2 methanation. J. CO2 Util. 2018, 23, 200–211. [Google Scholar] [CrossRef]
  18. Vogt, C.; Groeneveld, E.; Kamsma, G.; Nachtegaal, M.; Lu, L.; Kiely, C.J.; Berben, P.H.; Meirer, F.; Weckhuysen, B.M. Unravelling structure sensitivity in CO2 hydrogenation over nickel. Nat. Catal. 2018, 1, 127–134. [Google Scholar] [CrossRef]
  19. Danaci, S.; Protasova, L.; Lefevere, J.; Bedel, L.; Guilet, R.; Marty, P. Efficient CO2 methanation over Ni/Al2O3 coated structured catalysts. Catal. Today 2016, 273, 234–243. [Google Scholar] [CrossRef]
  20. Zhou, G.; Wu, T.; Xie, H.; Zheng, X. Effects of structure on the carbon dioxide methanation performance of Co-based catalysts. Int. J. Hydrog. Energy 2013, 38, 10012–10018. [Google Scholar] [CrossRef]
  21. Li, W.; Liu, Y.; Mu, M.; Ding, F.; Liu, Z.; Guo, X.; Song, C. Organic acid-assisted preparation of highly dispersed Co/ZrO2 catalysts with superior activity for CO2 methanation. Appl. Catal. B Environ. 2019, 254, 531–540. [Google Scholar] [CrossRef]
  22. Razzaq, R.; Li, C.; Usman, M.; Suzuki, K.; Zhang, S. A highly active and stable Co4N/γ-Al2O3 catalyst for CO and CO2 methanation to produce synthetic natural gas (SNG). Chem. Eng. J. 2015, 262, 1090–1098. [Google Scholar] [CrossRef]
  23. Yu, W.-Z.; Fu, X.-P.; Xu, K.; Ling, C.; Wang, W.-W.; Jia, C.-J. CO2 methanation catalyzed by a Fe-Co/Al2O3 catalyst. J. Environ. Chem. Eng. 2021, 9, 105594. [Google Scholar] [CrossRef]
  24. Li, W.; Nie, X.; Jiang, X.; Zhang, A.; Ding, F.; Liu, M.; Liu, Z.; Guo, X.; Song, C. ZrO2 support imparts superior activity and stability of Co catalysts for CO2 methanation. Appl. Catal. B Environ. 2018, 220, 397–408. [Google Scholar] [CrossRef]
  25. Li, W.; Zhang, G.; Jiang, X.; Liu, Y.; Zhu, J.; Ding, F.; Liu, Z.; Guo, X.; Song, C. CO2 hydrogenation on unpromoted and m-promoted Co/TiO2 catalysts (M = Zr, K, Cs): Effects of crystal phase of supports and metal-support interaction on tuning product distribution. ACS Catal. 2019, 9, 2739–2751. [Google Scholar] [CrossRef]
  26. Le, T.A.; Kim, M.S.; Lee, S.H.; Park, E.D. CO and CO2 methanation over supported cobalt catalysts. Top. Catal. 2017, 60, 714–720. [Google Scholar] [CrossRef]
  27. Satthawong, R.; Koizumi, N.; Song, C.; Prasassarakich, P. Comparative study on CO2 hydrogenation to higher hydrocarbons over Fe-based bimetallic catalysts. Top. Catal. 2013, 57, 588–594. [Google Scholar] [CrossRef]
  28. Serrer, M.-A.; Gaur, A.; Jelic, J.; Weber, S.; Fritsch, C.; Clark, A.H.; Saraçi, E.; Studt, F.; Grunwaldt, J.-D. Structural dynamics in Ni–Fe catalysts during CO2 methanation—Role of iron oxide clusters. Catal. Sci. Technol. 2020, 10, 7542–7554. [Google Scholar] [CrossRef]
  29. Pastor-Pérez, L.; Shah, M.; le Saché, E.; Ramirez Reina, T. Improving Fe/Al2O3 catalysts for the reverse water-gas shift reaction: On the effect of cs as activity/selectivity promoter. Catalysts 2018, 8, 608. [Google Scholar] [CrossRef] [Green Version]
  30. Rahmani, S.; Rezaei, M.; Meshkani, F. Preparation of highly active nickel catalysts supported on mesoporous nanocrystalline γ-Al2O3 for CO2 methanation. J. Ind. Eng. Chem. 2014, 20, 1346–1352. [Google Scholar] [CrossRef]
  31. Italiano, C.; Llorca, J.; Pino, L.; Ferraro, M.; Antonucci, V.; Vita, A. CO and CO2 methanation over Ni catalysts supported on CeO2, Al2O3 and Y2O3 oxides. Appl. Catal. B Environ. 2020, 264, 118494. [Google Scholar] [CrossRef]
  32. Visconti, C.G.; Lietti, L.; Tronconi, E.; Forzatti, P.; Zennaro, R.; Finocchio, E. Fischer–Tropsch synthesis on a Co/Al2O3 catalyst with CO2 containing syngas. Appl. Catal. A Gen. 2009, 355, 61–68. [Google Scholar] [CrossRef]
  33. Vile, G.; Colussi, S.; Krumeich, F.; Trovarelli, A.; Perez-Ramirez, J. Opposite face sensitivity of CeO(2) in hydrogenation and oxidation catalysis. Angew. Chem. Int. Ed. Eng. 2014, 53, 12069–12072. [Google Scholar] [CrossRef] [PubMed]
  34. Tada, S.; Shimizu, T.; Kameyama, H.; Haneda, T.; Kikuchi, R. Ni/CeO2 catalysts with high CO2 methanation activity and high CH4 selectivity at low temperatures. Int. J. Hydrog. Energy 2012, 37, 5527–5531. [Google Scholar] [CrossRef]
  35. Zhou, G.; Liu, H.; Cui, K.; Xie, H.; Jiao, Z.; Zhang, G.; Xiong, K.; Zheng, X. Methanation of carbon dioxide over Ni/CeO2 catalysts: Effects of support CeO2 structure. Int. J. Hydrog. Energy 2017, 42, 16108–16117. [Google Scholar] [CrossRef]
  36. Dias, Y.R.; Perez-Lopez, O.W. Carbon dioxide methanation over Ni-Cu/SiO2 catalysts. Energy Convers. Manag. 2020, 203, 112214. [Google Scholar] [CrossRef]
  37. Ye, R.-P.; Gong, W.; Sun, Z.; Sheng, Q.; Shi, X.; Wang, T.; Yao, Y.; Razink, J.J.; Lin, L.; Zhou, Z.; et al. Enhanced stability of Ni/SiO2 catalyst for CO2 methanation: Derived from nickel phyllosilicate with strong metal-support interactions. Energy 2019, 188, 116059. [Google Scholar] [CrossRef]
  38. Park, J.-N.; McFarland, E.W. A highly dispersed Pd–Mg/SiO2 catalyst active for methanation of CO2. J. Catal. 2009, 266, 92–97. [Google Scholar] [CrossRef]
  39. Liu, C.; Wang, W.; Xu, Y.; Li, Z.; Wang, B.; Ma, X. Effect of zirconia morphology on sulfur-resistant methanation performance of MoO3/ZrO2 catalyst. Appl. Surf. Sci. 2018, 441, 482–490. [Google Scholar] [CrossRef]
  40. Zhao, K.; Wang, W.; Li, Z. Highly efficient Ni/ZrO2 catalysts prepared via combustion method for CO2 methanation. J. CO2 Util. 2016, 16, 236–244. [Google Scholar] [CrossRef]
  41. Xu, X.; Tong, Y.; Huang, J.; Zhu, J.; Fang, X.; Xu, J.; Wang, X. Insights into CO2 methanation mechanism on cubic ZrO2 supported Ni catalyst via a combination of experiments and DFT calculations. Fuel 2021, 283, 118867. [Google Scholar] [CrossRef]
  42. Ren, J.; Qin, X.; Yang, J.-Z.; Qin, Z.-F.; Guo, H.-L.; Lin, J.-Y.; Li, Z. Methanation of carbon dioxide over Ni–M/ZrO2 (M = Fe, Co, Cu) catalysts: Effect of addition of a second metal. Fuel Process. Technol. 2015, 137, 204–211. [Google Scholar] [CrossRef]
  43. Chai, S.; Men, Y.; Wang, J.; Liu, S.; Song, Q.; An, W.; Kolb, G. Boosting CO2 methanation activity on Ru/TiO2 catalysts by exposing (001) facets of anatase TiO2. J. CO2 Util. 2019, 33, 242–252. [Google Scholar] [CrossRef]
  44. Kim, A.; Debecker, D.P.; Devred, F.; Dubois, V.; Sanchez, C.; Sassoye, C. CO2 methanation on Ru/TiO2 catalysts: On the effect of mixing anatase and rutile TiO2 supports. Appl. Catal. B Environ. 2018, 220, 615–625. [Google Scholar] [CrossRef]
  45. Liu, J.; Li, C.; Wang, F.; He, S.; Chen, H.; Zhao, Y.; Wei, M.; Evans, D.G.; Duan, X. Enhanced low-temperature activity of CO2 methanation over highly-dispersed Ni/TiO2 catalyst. Catal. Sci. Technol. 2013, 3, 2627–2633. [Google Scholar] [CrossRef]
  46. Song, H.; Yang, J.; Zhao, J.; Chou, L. Methanation of carbon dioxide over a highly dispersed Ni/La2O3 catalyst. Chin. J. Catal. 2010, 31, 21–23. [Google Scholar] [CrossRef]
  47. Takami, K.; Takahash, T. Kinetics of the methanation of carbon dioxide over a supported La2O3. Can. J. Eng. 1988, 66, 343–347. [Google Scholar]
  48. Li, S.; Tang, H.; Gong, D.; Ma, Z.; Liu, Y. Loading Ni/La2O3 on SiO2 for CO methanation from syngas. Catal. Today 2017, 297, 298–307. [Google Scholar] [CrossRef]
  49. Kock, E.M.; Kogler, M.; Bielz, T.; Klotzer, B.; Penner, S. In situ FT-IR spectroscopic study of CO2 and CO adsorption on Y2O3, ZrO2, and yttria-stabilized ZrO2. J. Phy.s Chem. C Nanomater. Interfaces 2013, 117, 17666–17673. [Google Scholar] [CrossRef]
  50. Ayub, N.A.; Bahruji, H.; Mahadi, A.H. Barium promoted Ni/Sm2O3 catalysts for enhanced CO2 methanation. RSC Adv. 2021, 11, 31807–31816. [Google Scholar] [CrossRef]
  51. Ilsemann, J.; Sonström, A.; Gesing, T.M.; Anwander, R.; Bäumer, M. Highly active Sm2O3-Ni xerogel catalysts for CO2 methanation. ChemCatChem 2019, 11, 1732–1741. [Google Scholar] [CrossRef]
  52. Bian, Z.; Chan, Y.M.; Yu, Y.; Kawi, S. Morphology dependence of catalytic properties of Ni/CeO2 for CO2 methanation: A kinetic and mechanism study. Catal. Today 2020, 347, 31–38. [Google Scholar] [CrossRef]
  53. Guo, Y.; Mei, S.; Yuan, K.; Wang, D.-J.; Liu, H.-C.; Yan, C.-H.; Zhang, Y.-W. Low-temperature CO2 methanation over CeO2-supported Ru single atoms, nanoclusters, and nanoparticles competitively tuned by strong metal–support interactions and H-spillover effect. ACS Catal. 2018, 8, 6203–6215. [Google Scholar] [CrossRef]
  54. Ye, R.-P.; Li, Q.; Gong, W.; Wang, T.; Razink, J.J.; Lin, L.; Qin, Y.-Y.; Zhou, Z.; Adidharma, H.; Tang, J.; et al. High-performance of nanostructured Ni/CeO2 catalyst on CO2 methanation. Appl. Catal. B Environ. 2020, 268, 118474. [Google Scholar] [CrossRef]
  55. Sakpal, T.; Lefferts, L. Structure-dependent activity of CeO2 supported Ru catalysts for CO2 methanation. J. Catal. 2018, 367, 171–180. [Google Scholar] [CrossRef]
  56. Jiang, M.; Wang, B.; Yao, Y.; Li, Z.; Ma, X.; Qin, S.; Sun, Q. A comparative study of CeO2-Al2O3 support prepared with different methods and its application on MoO3/CeO2-Al2O3 catalyst for sulfur-resistant methanation. Appl. Surf. Sci. 2013, 285, 267–277. [Google Scholar] [CrossRef]
  57. Ho, P.H.; de Luna, G.S.; Angelucci, S.; Canciani, A.; Jones, W.; Decarolis, D.; Ospitali, F.; Aguado, E.R.; Rodríguez-Castellón, E.; Fornasari, G.; et al. Understanding structure-activity relationships in highly active La promoted Ni catalysts for CO2 methanation. Appl. Catal. B Environ. 2020, 278, 119256. [Google Scholar] [CrossRef]
  58. Liang, C.; Tian, H.; Gao, G.; Zhang, S.; Liu, Q.; Dong, D.; Hu, X. Methanation of CO2 over alumina supported nickel or cobalt catalysts: Effects of the coordination between metal and support on formation of the reaction intermediates. Int. J. Hydrog. Energy 2020, 45, 531–543. [Google Scholar] [CrossRef]
  59. Mihet, M.; Lazar, M.D. Methanation of CO2 on Ni/γ-Al2O3: Influence of Pt, Pd or Rh promotion. Catal. Today 2018, 306, 294–299. [Google Scholar] [CrossRef]
  60. Hongmanorom, P.; Ashok, J.; Zhang, G.; Bian, Z.; Wai, M.H.; Zeng, Y.; Xi, S.; Borgna, A.; Kawi, S. Enhanced performance and selectivity of CO2 methanation over phyllosilicate structure derived Ni-Mg/SBA-15 catalysts. Appl. Catal. B Environ. 2021, 282, 19564. [Google Scholar] [CrossRef]
  61. Li, T.; Wang, S.; Gao, D.-N.; Wang, S.-D. Effect of support calcination temperature on the catalytic properties of Ru/Ce0.8Zr0.2O2 for methanation of carbon dioxide. J. Fuel Chem. Technol. 2014, 42, 1440–1446. [Google Scholar] [CrossRef]
  62. Wu, H.; Zou, M.; Guo, L.; Ma, F.; Mo, W.; Yu, Y.; Mian, I.; Liu, J.; Yin, S.; Tsubaki, N. Effects of calcination temperatures on the structure–activity relationship of Ni–La/Al2O3 catalysts for syngas methanation. RSC Adv. 2020, 10, 4166–4174. [Google Scholar] [CrossRef] [Green Version]
  63. Sexton, B.A.; Hughes, A.E.; Turney, T.W. An XPS and TPR study of the reduction of promoted cobalt-kieselguhr Fischer-Tropsch catalysts. J. Catal. 1986, 97, 390–406. [Google Scholar] [CrossRef]
  64. Castillo, J.; Arteaga-Pérez, L.E.; Karelovic, A.; Jiménez, R. The consequences of surface heterogeneity of cobalt nanoparticles on the kinetics of CO methanation. Catal. Sci. Technol. 2019, 9, 6415–6427. [Google Scholar] [CrossRef]
  65. Le, T.A.; Kim, M.S.; Lee, S.H.; Kim, T.W.; Park, E.D. CO and CO2 methanation over supported Ni catalysts. Catal. Today 2017, 293–294, 89–96. [Google Scholar] [CrossRef]
  66. Liu, K.; Xu, X.; Xu, J.; Fang, X.; Liu, L.; Wang, X. The distributions of alkaline earth metal oxides and their promotional effects on Ni/CeO2 for CO2 methanation. J. CO2 Util. 2020, 38, 113–124. [Google Scholar] [CrossRef]
  67. Le, T.A.; Kim, J.; Kang, J.K.; Park, E.D. CO and CO methanation over Ni/Al@Al2O3 core–shell catalyst. Catal. Today 2020, 356, 622–630. [Google Scholar] [CrossRef]
  68. Xu, L.; Cui, Y.; Chen, M.; Wen, X.; Lv, C.; Wu, X.; Wu, C.-E.; Miao, Z.; Hu, X. Screening transition metals (Mn, Fe, Co, and Cu) promoted Ni-based CO2 methanation bimetal catalysts with advanced low-temperature activities. Ind. Eng. Chem. Res. 2021, 60, 8056–8072. [Google Scholar] [CrossRef]
  69. Yu, Y.; Mottaghi-Tabar, S.; Iqbal, M.W.; Yu, A.; Simakov, D.S.A. CO2 methanation over alumina-supported cobalt oxide and carbide synthesized by reverse microemulsion method. Catal. Today 2021, 379, 250–261. [Google Scholar] [CrossRef]
  70. Hadjiivanov, K.; Ivanova, E.; Dimitrov, L.; Knözinger, H. FTIR spectroscopic study of CO adsorption on Rh–ZSM-5: Detection of Rh+ −CO species. J. Mol. Struct. 2003, 661–662, 459–463. [Google Scholar] [CrossRef]
  71. Choi, J.G.; Rhee, H.K.; Moon, S.H. IR and TPD study of fresh and carbon-deposited Co/Al2O3 catalysts. Appl. Catal. A Gen. 1985, 13, 269–280. [Google Scholar] [CrossRef]
  72. Watson, C.D.; Martinelli, M.; Cronauer, D.C.; Kropf, A.J.; Jacobs, G. Low temperature water-gas shift: Enhancing stability through optimizing Rb loading on Pt/ZrO2. Catalysts 2021, 11, 210. [Google Scholar] [CrossRef]
  73. Das, T.; Deo, G. Synthesis, characterization and in situ DRIFTS during the CO2 hydrogenation reaction over supported cobalt catalysts. J. Mol. Catal. A Chem. 2011, 350, 75–82. [Google Scholar] [CrossRef]
  74. Baltrusaitis, J.; Schuttlefield, J.; Zeitler, E.; Grassian, V.H. Carbon dioxide adsorption on oxide nanoparticle surfaces. Chem. Eng. J. 2011, 170, 471–481. [Google Scholar] [CrossRef]
Figure 1. H2-TPR patterns of ceria-supported cobalt catalysts calcined in air at different temperatures.
Figure 1. H2-TPR patterns of ceria-supported cobalt catalysts calcined in air at different temperatures.
Catalysts 12 00212 g001
Figure 2. CO2-TPD profiles of ceria-supported cobalt catalysts calcined in air at different temperatures and reduced in H2 at 500 °C.
Figure 2. CO2-TPD profiles of ceria-supported cobalt catalysts calcined in air at different temperatures and reduced in H2 at 500 °C.
Catalysts 12 00212 g002
Figure 3. Catalytic activity for CO methanation over ceria-supported cobalt catalysts. All catalysts were reduced in H2 at 500 °C. The feed gas is composed of 1 mol% CO, 49 mol% He, and 50 mol% H2, and the total flow rate is 100 mL/min.
Figure 3. Catalytic activity for CO methanation over ceria-supported cobalt catalysts. All catalysts were reduced in H2 at 500 °C. The feed gas is composed of 1 mol% CO, 49 mol% He, and 50 mol% H2, and the total flow rate is 100 mL/min.
Catalysts 12 00212 g003
Figure 4. Catalytic activity for CO2 methanation over ceria-supported cobalt catalysts. All catalysts were reduced in H2 at 500 °C. The feed gas is composed of 1 mol% CO2, 49 mol% He, and 50 mol% H2, and the total flow rate is 100 mL/min.
Figure 4. Catalytic activity for CO2 methanation over ceria-supported cobalt catalysts. All catalysts were reduced in H2 at 500 °C. The feed gas is composed of 1 mol% CO2, 49 mol% He, and 50 mol% H2, and the total flow rate is 100 mL/min.
Catalysts 12 00212 g004
Figure 5. In situ DRIFTS spectra after adsorption of CO2 on Co0.1Ce0.9Ox(300) (A) and Co/CeO2(500) (B) at 40 °C. The feed gas is composed of 10 mol% CO2 and 90 mol% He.
Figure 5. In situ DRIFTS spectra after adsorption of CO2 on Co0.1Ce0.9Ox(300) (A) and Co/CeO2(500) (B) at 40 °C. The feed gas is composed of 10 mol% CO2 and 90 mol% He.
Catalysts 12 00212 g005
Figure 6. In situ DRIFTS spectra during CO2 methanation at temperatures of 100-450 °C over the Co0.1Ce0.9Ox(300) catalyst in the range of 2500–1000 cm−1 (A) and 2500-4000 cm−1 (B). The feed is composed of 1 mol% CO2, 49 mol% He, and 50 mol% H2.
Figure 6. In situ DRIFTS spectra during CO2 methanation at temperatures of 100-450 °C over the Co0.1Ce0.9Ox(300) catalyst in the range of 2500–1000 cm−1 (A) and 2500-4000 cm−1 (B). The feed is composed of 1 mol% CO2, 49 mol% He, and 50 mol% H2.
Catalysts 12 00212 g006
Figure 7. In situ DRIFTS spectra during CO2 methanation at temperatures of 100–450 °C over the Co/CeO2(500) catalyst in the range of 2500–1000 cm−1 (A) and 2500–4000 cm−1 (B). The feed is composed of 1 mol% CO2, 49 mol% He, and 50 mol% H2.
Figure 7. In situ DRIFTS spectra during CO2 methanation at temperatures of 100–450 °C over the Co/CeO2(500) catalyst in the range of 2500–1000 cm−1 (A) and 2500–4000 cm−1 (B). The feed is composed of 1 mol% CO2, 49 mol% He, and 50 mol% H2.
Catalysts 12 00212 g007
Table 1. Physicochemical properties of ceria-supported cobalt catalysts a.
Table 1. Physicochemical properties of ceria-supported cobalt catalysts a.
CatalystsBET Surface Area b
(m2/g)
Pore Volume b
(cm3/g)
Average Pore Diameter b
(nm)
Cobalt Dispersion c
(%)
CASA c
(m2/gcat)
CO2 Uptake d
(μmol/gcat)
Co0.1Ce0.9Ox(300)900.2310.24.650.74124
Co0.1Ce0.9Ox(500)720.2818.95.980.95146
Co0.1Ce0.9Ox(700)340.2226.33.270.5259
Co/CeO2(300)1300.123.93.870.94196
Co/CeO2(500)1280.144.34.651.13186
Co/CeO2(700)600.138.63.720.90122
a All catalysts were calcined in air at different temperatures and reduced in H2 at 500 °C. b The data were determined with N2 physisorption. c The data were calculated based on H2 chemisorption. d The data were quantified based on pulsed CO2 chemisorption.
Table 2. FT-IR peaks during CO2 adsorption and hydrogenation over Co0.1Ce0.9Ox(300) and Co/CeO2(500).
Table 2. FT-IR peaks during CO2 adsorption and hydrogenation over Co0.1Ce0.9Ox(300) and Co/CeO2(500).
Vibrational Mode AssignmentsIR Peak Position (cm−1)References
Co0.1Ce0.9Ox(300)Ce/CeO2(500)Literature
AdsorptionHydrogenationAdsorptionHydrogenation
40 °C300 °C 40 °C300 °C
Absorbed CO* species on metallic cobalt 2184
2142
2045
2115
2075
211921152181, 2168, 2118, 2084, 2048[70,71]
Absorbed CO2 2200–2400[69]
Bicarbonate (HCO3) [33,72,73,74]
ν(OH)
νasy(OCO)
νs(OCO)
ν(COH)

1602
1408
1217



1217



1216

1654
1434
3636, 3667


1216
3600–3627, 3620
1655, 1652, 1650
1440, 1424, 1435
1225, 1220, 1228
Carbonate (CO3) 1550 and 1380[33,69,74]
Monodentate
νasy(OCO)
νs(OCO)
ν(CO)


1450
1385



1394


1475
1380


1457
1354

1452
1349, 1396
1075

1446–1590
1380–1395
1040
Bidentate
νasy(OCO)
νs(OCO)


1293


1288

1575
1286
1545
1265
1331

1577
1279


1284

1535–1670
1243–1355
Formate
ν(CH)
ν(OCO)


2850

2713, 2849
2944

2848
1523

2838, 2719
2940

2756–2866, 2905
1540, 1510, 1450
[69,72,73]
Methane 3014 1305, 30141305 and 3015[73]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nguyen, T.H.; Kim, H.B.; Park, E.D. CO and CO2 Methanation over CeO2-Supported Cobalt Catalysts. Catalysts 2022, 12, 212. https://doi.org/10.3390/catal12020212

AMA Style

Nguyen TH, Kim HB, Park ED. CO and CO2 Methanation over CeO2-Supported Cobalt Catalysts. Catalysts. 2022; 12(2):212. https://doi.org/10.3390/catal12020212

Chicago/Turabian Style

Nguyen, Thuy Ha, Han Bom Kim, and Eun Duck Park. 2022. "CO and CO2 Methanation over CeO2-Supported Cobalt Catalysts" Catalysts 12, no. 2: 212. https://doi.org/10.3390/catal12020212

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