Next Article in Journal
Se-Doped Ni5P4 Nanocatalysts for High-Efficiency Hydrogen Evolution Reaction
Next Article in Special Issue
In Situ Formation of Z-Scheme Bi2WO6/WO3 Heterojunctions for Gas-Phase CO2 Photoreduction with H2O by Photohydrothermal Treatment
Previous Article in Journal
Design of a Multi-Tubular Catalytic Reactor Assisted by CFD Based on Free-Convection Heat-Management for Decentralised Synthetic Methane Production
Previous Article in Special Issue
Cu/Zn/Zr/Ga Catalyst for Utilisation of Carbon Dioxide to Methanol—Kinetic Equations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

CO2 Methanation of Biogas over Ni-Mg-Al: The Effects of Ni Content, Reduction Temperature, and Biogas Composition

1
Department of Environment-Energy Engineering, The University of Suwon, 17 Wauan-gil, Bongdam-eup, Hwaseong-si 18323, Korea
2
Bio Friends Inc., Yuseong-gu, Daejeon 34028, Korea
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(9), 1054; https://doi.org/10.3390/catal12091054
Submission received: 22 July 2022 / Revised: 5 September 2022 / Accepted: 13 September 2022 / Published: 16 September 2022
(This article belongs to the Special Issue CO2 Catalytic Conversion and Utilization)

Abstract

:
Biogas is mainly composed of CH4 and CO2, so it is used as an alternative energy to CH4 with high energy density by separating and removing CO2 from biogas. In addition, it can be utilized by producing synthesis gas (CO and H2) through thermal decomposition of biogas or by synthesizing CH4 by methanation of CO2. The technique of CO2 methanation is a method that can improve the CH4 concentration without CO2 separation. This study aims to produce more efficient methane through CO2 methanation of biogas over Ni-Mg-Al catalyst. So, the effect of Ni contents in catalyst, catalyst reduction temperature, CO2 concentration in biogas, and the initial concentration of CH4 on CO2 conversion rate and CH4 selectivity was investigated. In addition, the effect of increasing CO2 concentration, H2/CO2 ratio, and GHSV (gas space velocity per hour) on H2 conversion, CH4 productivity, and product was investigated. In particular, the durability and stability of CO2 methanation was tested over 60 wt% Ni-Mg-Al catalyst at 350 °C and 30,000/h for 130 h. From the long-term test results, the catalyst shows stability by maintaining a constant CO2 conversion rate of 72% and a CH4 selectivity of 95%.

1. Introduction

Because of the rapid development of the global economy, the severities of the energy crisis and environmental pollution are increasing. In 2020, fossil fuel consumption accounted for 10.3 Gt of CO2 emissions. At the current rate, the Earth’s temperature is predicted to rise by 6 °C before the end of 2050. To combat the threat of climate change, the Paris Climate agreement, in which countries around the world pledged carbon neutrality to limit the increase in the Earth’s temperature to 1.5 °C by 2050, was adopted. Measures to minimize CO2 in the atmosphere include the replacement of fossil fuels with alternative energy sources (e.g., hydrogen) and utilizing carbon capture, utilization, and storage (CCUS) technologies [1,2,3]. Among CCUS technologies, chemical conversion is the most convenient and effective [4] because the CO2 generated from the fuel is converted to fuel through the aid of a catalyst and recycled, thereby ensuring carbon neutrality [5]. As CO2 methanation, an important process that involves catalysis, has tremendous potential for commercial applications, it has received significant attention. The methanation reaction can reduce environmental pollution through the conversion of CO2 to CH4, which is cleaner for the environment, and it can alleviate the energy shortage problem. Therefore, methanation is among the most effective processes for mitigating CO2 emissions [6,7].
Catalytic processes involve a catalyst, a catalyst support, a catalytic promoter, and group VIII metals, such as Ru, Rh, Pd, Ni, Co, and Fe to provide active sites for the CO2 methanation reaction. Among these metals, Pd, Rh, and Ru are precious metals that are characterized by low-temperature catalytic activity and high methane selectivity [8]. However, because of their high cost and scarcity, these metals are unsuitable for large-scale applications [9]. Conversely, Co-based catalysts exhibit low-temperature activity and stability but low methane selectivity [10,11]. Further, in the presence of Fe-based catalysts, carbon accumulates on the surface of the catalyst and liquid hydrocarbons are easily formed; however, these catalysts are unsuitable for low-temperature reactions [12]. Ni-based catalysts have been extensively studied for application in CO2 methanation because of their excellent performance and low cost. Therefore, efforts have been devoted to identifying supports and promoters of Ni-based catalysts to address their low-temperature activity and sintering resistance limitations.
La3+, Pr3+, Sm3+ in to the crystal structure of cerium oxide created a higher population of oxygen vacant sites. Moreover, the co-presence of La3+, Mg2+ and La3+, Pr3+ into the CeO2 increased the methos of moderate basic sites. These physicochemical properties increased the rate of CO2 methanation reaction at a relatively low temperature. CO2 conversion was observed at 350 °C, Ni/La-Pr-Ce (55%) > Ni/La-Sm-Ce (53%) > Ni/La-Mg-Ce (49%) > Ni Ce(39%) [13].
In a series of 5%X-12%Ni/γ-Al2O3 (X = La, Ce, Pr, Eu & Gd) catalysts, 5%Pr-12%Ni/γ-Al2O3 catalysts showed the highest CO2 conversion of 98.2% with 100% CH4 selectivity at 300 °C for investigated reaction conditions. Additionally, the catalyst was tested for 48h, which showed the stability of the catalyst for CO2 conversion and CH4 selectivity. CO2 conversion was observed at 300 °C, 5%Pr-12%Ni/γ-Al2O3 > 5% Ce-12% Ni/γ-Al2O3 > Eu-12% Ni/γ-Al2O3 > 5% La-12%Ni/γ-Al2O3 12% Ni/γ-Al2O3 [14].
Types of bimetallic oxides act as catalysts and remarkably almost all of them present a catalytic activity superior to that of a commercial rhodium catalyst supported on alumina (5 wt% Rh/Al2O3) for the production of methane. CO2 conversion increases until 350–400 °C and stabilizes or decreases over all the nickel-lanthanide bimetallic oxides at higher temperatures. At temperature above 350–400 °C there is an increase in the formation of CO by product, due to reversed WGS reaction [15].
Considering that structural properties, dispersion, and metal-support interactions are major factors affecting the low-temperature catalytic activity and sintering resistance of Ni-based catalysts, designing and producing catalysts with adequate porosity, high potential for Ni dispersion, and elevated metal-support interaction is critical [1]. High Ni dispersion is generally achieved by using micro- or mesoporous supports characterized by excellent structural properties and the addition of promoters, such as Co, Zr, Sn, and Mg [16]. Supports can suppress or prevent sintering by enhancing the metal-support interaction and thus influence CO2 methanation by promoting the dissociation and adsorption of CO2 [17]. A catalytic support having a high specific surface area and an elevated pore volume can accommodate active Ni sites in a dispersed state [18]. Enhanced electron transfer between the support and the metal active sites increases the electron density of the metal and strengthens the Ni–C bond, which facilitates the dissociation of the C=O bond and the formation of CH4 [19]. In addition, other physicochemical properties of the support as well as the particle size and surface characteristics of Ni affect the reducibility and properties of the catalyst. In general, substances such as Al2O3, TiO2, SiO2, and ZrO2 are utilized as supports for Ni-based catalysts for CO2 methanation [15].
According to Romero-Sáez et al. [20], ZrO2 is unable to activate and dissociate H2, but it can activate CO2 to produce CO. Therefore, H2 molecules can be dissociated on the Ni surface, while CO2 molecules are activated on the ZrO2 surface [21]. Thus, interactions between hydrogen atoms and activated CO2 molecules are facilitated, and this expands the Ni-ZrO2 interface, thereby enhancing the CH4 selectivity and reaction rate [22]. However, owing to its excellent stability, elevated CO2 adsorption potential, and high oxygen storage capacity, CeO2 likely has a superior low-temperature activity to supported Ni-based catalysts [23,24,25,26]. Al2O3 is a low-cost porous support with a high specific area that enables uniform dispersion of Ni on the catalyst surface [27,28]. In addition, Al2O3 and Ni can react to form NiAl2O4 because the strong Ni–O bond prevents the reduction of Ni2+. The fine size of the Ni particles on the catalyst surface hinders carbon deposition [29].
Moghaddam et al. [30] indicated that an additive (promoter), such as Fe, Co, Zr, La, or Cu, can improve the conversion of CO2. Stangeland et al. [31] investigated CO2 methanation under the reaction conditions of the catalyst. Conversely, Aziz et al. [32] highlighted the importance of the physicochemical properties, catalytic activity, and reaction mechanism of a catalyst, and discussed the developments in heterogeneous catalysts. Doping catalysts with Zr, Co, La, and Mg alters the CO2 methanation reaction pathways by introducing alternative intermediates. Thus, CO formation can be prevented by effectively suppressing side reactions. MgO has been widely investigated as a basic modifier of Ni-based catalysts for CO2 methanation [33]. In particular, the addition of a small amount of MgO to a catalyst promotes the chemisorption and dissociation of CO2 [34]. MgO also improves the dispersion of Ni and modifies the oxidizing environment near its particles in the catalyst. Thus, severe aggregation at active Ni sites and carbon deposition on the catalyst surface can be prevented [35,36,37].
In this study, a Ni-Mg-Al catalyst with excellent redox properties was prepared. This Ni-based catalyst is inexpensive [38] and exhibits high activity, the Al support has high thermal stability and strong resistance to sintering and carbon deposition, and MgO enhances the adsorption behavior and dispersion of active Ni, thereby enhancing the catalytic activity [39]. In addition, MgO reduces catalyst deactivation, such as sintering and carbon formation, as well as negative effects on water on the catalyst. Since the reaction intermediate (CO) is converted to CH4, using Ni-Mg-Al catalysts with sufficient Ni content of 20, 40, and 60 wt% to provide many reaction sites, high activity and selectivity of the catalyst for CO2 methanation experiments improved [40]. In particular, the activation energy of the CO2 methanation reaction for a Ni-Mg-Al catalyst containing 60 wt% Ni was investigated. Based on the experimental results, the influence of CO2 concentration, GHSV (gas hourly space velocity) and H2/CO2 ratio for CH4 productivity and product production was investigated. Durability and stability for 130 h was performed for CO2 methanation over 60 wt% Ni-Mg-Al catalyst.

2. Results and Discussion

2.1. Effects of Temperature Reduction and Catalyst Loading on CO2 Conversion

The effect of the Ni proportion in the Ni-Mg-Al catalyst at reduced temperatures of 450–700 °C on the CO2 conversion is shown in Figure 1. Obviously, CO2 conversion and CH4 selectivity increased as the reduced temperature of the catalyst increased to 650–700 °C, indicating similarity in the behavior of the catalysts. However, 60 wt% Ni catalyst exhibited no significant change as the reduction temperature varied. This is consistent with the H2-TPR analysis results (Figure 11), which revealed that the maximum reducibility of the catalysts occurred at 680–720 °C. Therefore, 700 °C was selected as the reduction temperature for the activity experiments under different conditions.
The 20, 40, and 60wt% Ni catalysts were reduced to 700 °C, and the results of the catalysts activity experiment with reaction temperature at a H2/CO2 ratio of 4 and GHSV of 30,000/h are shown in Table 1. The temperature at which CO2 conversion reached 50% was 340, 298, and 291 °C, respectively, for the catalysts with 20, 40, and 60 wt% Ni, which indicates that the temperature decreases as the proportion of Ni increases. This behavior is attributed to the increase in active Ni sites as the proportion of Ni increases. This is in line with the XRD data and the TOF values, which showed an increase in active Ni sites, reflecting reactivity. In fact, a higher Ni loading promotes the production of CH4. This behavior is likely associated with an increase in catalytic activity because a high Ni loading increases the number of active sites. According to Wierzbicki et al. [42], the metal loading determines the reduction properties of a catalyst. Zhang et al. [43] indicated that the RWGS side reaction competes with the CO2 methanation reaction for active sites at a low Ni loading and thus a high Ni loading (≥25%) promotes CO2 methanation. Quindimil et al. [44] noted that an increase in the number of active sites (NiO) enhances CO2 conversion, although increasing Ni loading beyond a certain limit can damage the structure of the catalyst through thermal agglomeration of active Ni sites.

2.2. Effect of Temperature Reaction on CO2 Conversion

The reaction temperature is the dominant factor affecting the activity of Ni-based catalysts. Activation energy higher than the dissociation energy of CO2 is required for CO2 conversion at low temperatures. Therefore, increasing the reaction temperature enables the CO2 molecules to attain the required activation energy and thus enhances the reactivity. The highest CO2 conversion and CH4 selectivity were achieved at temperatures ranging between 300 °C–400 °C. Beyond this temperature range, the CH4 selectivity decreased. Mutz et al. [45] noted that, based on thermodynamics, the high-temperature and exothermic characteristics of the CO2 methanation process affect the equilibrium and deactivation of catalysts. High temperatures have been reported to promote the RWGS reaction and impede the CO2 methanation reaction. In addition, as shown in the equilibrium curve in Figure 2, the limitation of the CO2 conversion rate with temperature is also the cause of the lower CO2 conversion rate. CO2 conversion and CH4 selectivity with reaction temperature at a GHSV of 30,000/h and an H2/CO2 ratio of 4 are shown in Figure 2. Evidently, CO2 conversion increased as the temperature increased and peaked at 400 °C, while CH4 selectivity and yield reached maximum values at 350 °C. This behavior is attributed to suppression of the methanation reaction and an increase in the RWGS reaction above 350 °C, which promotes the conversion of CO2 to CO. In line herewith, Mohammad et al. [46] reported that the CO concentration increased and methane selectivity decreased at 400 °C because the RWGS reaction is enhanced at this temperature. In addition, Jia et al. [47,48] reported the highest CO2 conversion and CH4 yield at 350 °C, and CO2 conversion decreased beyond this temperature because of the thermodynamic equilibrium limitation. In the present study, the activation energy of 60 wt% Ni-Mg-Al catalyst was reduced and reacted at 200, 250, and 300 °C, and based on the data shown in Figure 3 the activation energy was reduced to approximately 59.7 kJ/mol. This demonstrates the excellent activity of our catalyst and this value is substantially lower than the 75 kJ/mol achieved with a Ni/Al2O3 catalyst as reported by Garbarino et al. [22].

2.3. Effects of CO2 and H2 Concentrations on CO2 Conversion

Figure 4 shows the effects of the CO2 concentration (6, 10, 16, and 20 vol%) on CO2 conversion and CH4 selectivity for 60 wt% Ni catalyst with an H2/CO2 ratio of 4 and a GHSV of 30,000/h. CH4 selectivity essentially remained constant at approximately 98% as the CO2 concentration increased, while CO2 conversion increased because of the increase in reactants and the decrease in N2, which enhance the likelihood of reactions. At a CO2 concentration of approximately 15 vol%, the H2/CO2 ratio varied between 3.5 and 5 in increments of 0.5 under the same GHSV. The effects of the H2 concentration on CO2 conversion and CH4 selectivity are shown in Figure 5. As the H2 concentration increased, CH4 selectivity remained stable at approximately 97%, while the CO2 conversion increased at a level comparable to that observed for the effect of the CO2 concentration. Therefore, to enhance CH4 production, high CO2 and H2 concentrations are required.

2.4. Effect of CH4 Concentration on CO2 Conversion

The effect of the initial CH4 concentration on the production of CH4 from CO2 in biogas is shown in Figure 6. As the CH4 concentration increased from 0% to 16.8%, CO2 conversion and CH4 production decreased. According to Le Chatelier’s principle, if CH4 is present among the reactants, CO2 conversion to CH4 is hindered after equilibrium is attained, and thus an increase in the initial concentration of CH4 decreases CH4 production. Therefore, CH4 production decreases at a higher rate than the increase in the overall flow as the concentration of CH4 increases. Based on a simulation study, Jürgensen [49] found that CO2 conversion decreases as the initial CH4 concentration increases. H2 conversion showed minimal changes, indicating that it is essentially unaffected by the initial CH4 concentration.

2.5. Effect of CO2 Concentration, H2/CO2 Ratio and GHSV on CH4 Productivity and Product

Effect of CO2 concentration, H2/CO2 ratio, and GHSV on CH4 productivity and product at 350 °C are shown in Figure 7, Figure 8 and Figure 9, respectively. As mentioned before, the increase in the CO2 reactant not only increases the CO2 conversion rate but also increases the productivity and product due to the increase in reaction opportunity with a decrease in inert nitrogen. As shown in Figure 7, as the H2/CO2 ratio increased, H2 conversion decreased, while the CH4 product and productivity remained almost unchanged. It is important and necessary to find the optimization of the reaction from the above results. Figure 9 shows the effect of increasing GHSV on CH4 productivity and product. As shown in Figure 9, as GHSV increased, H2 conversion and CH4 productivity slightly decreased, but CH4 product increased. The CH4 productivity and product were different, depending on the presence or absence of a reaction in the balance of N2. In the reaction without N2, CH4 productivity and product decreased by 1.1% and increased by 1.5 mol, respectively, whereas in the reaction with N2, it decreased by 1.4% and increased by 0.37 mol.

2.6. Catalytic Activity and Stability Tests

The activity and stability of the catalyst were evaluated for approximately 130 h under a GHSV of 30,000/h, a temperature of 350 °C, and a CO2 concentration of 16 vol% at H2/CO2 ratios of 3.3, 3.6, and 4.0 (Figure 10). The maximum CO2 conversion and CH4 selectivity were 72% and 95%, respectively, during the 130 h. During the test period of 130 hours for CO2 methanation, almost no carbon was produced. These results confirm that the Ni-Mg-Al catalyst provides adequate activity and stability for CO2 methanation.

3. Characterization of the Catalysts

3.1. Surface Area Analysis

The surface area, which is a critical catalyst property, was measured by the Brunauer–Emmett–Teller (BET) method using an ASAP 2020 Plus Physisorption instrument (Micromeritics Instrument Corporation, Norcross, GA, USA). The surface area, pore volume, and pore size of fresh (20 and 40 wt% Ni-Mg-Al) and spent (20 wt% Ni-Mg-Al, 200 h of use) catalysts are presented in Table 2. Evidently, the BET surface area decreased as the Ni content increased, and the surface area, pore volume, and pore size of the spent catalysts were lower than those of the fresh catalysts. It is thought to be due to catalyst agglomeration and pore breakage through reduction and reaction.

3.2. H2-Temperature Programmed Reduction (TPR) and H2-Chemisorption Analyses

The results of H2-TPR analysis before and after reduction of the 20, 40, and 60 wt% Ni-Mg-Al catalysts at 600 °C using an Auto Chem II 2920 chemisorption analyzer are shown in Figure 11a and Figure 11b, respectively. Two peaks are observed in the low-temperature (<200 °C) and near-high-temperature (700 °C) regions, with the latter being substantially larger. The larger peak is attributed to a higher H2 concentration because of the higher Ni proportion, and it shifted toward lower temperatures. Conversely, in Figure 11b, only one peak is visible between 150–180 °C, demonstrating the complete reduction of the catalysts in the reactor. The peak observed for the spent catalyst is smaller than that for the same catalyst before the reduction. The temperature associated with peaks decreases as the proportion of Ni increases, which indicates improved reducibility. The low-temperature behavior observed during the reduction of the catalysts suggests the presence of some amorphous substances or decreased interaction because of the strongly reducing MgO. These observations were consistent with those obtained by X-ray diffraction (XRD) and X-ray photoelectron spectroscopy (XPS) (see below).
Figure 11. H2-TPR profiles of (a) fresh and (b) spent catalysts (* reference [41] was cited).
Figure 11. H2-TPR profiles of (a) fresh and (b) spent catalysts (* reference [41] was cited).
Catalysts 12 01054 g011
The Ni dispersion, particle size, and turnover frequency (TOF) of the 20, 40, and 60 wt% Ni-Mg-Al catalysts at 250 °C measured through H2-chemisorption analysis are presented in Table 3. As the Ni content increases, the Ni dispersion increased slightly and then decreased slightly. According to many researchers [50], increasing the catalyst content tends to significantly decrease the dispersion of Ni, but in this paper the dispersion of the catalyst is maintained uniformly. The strong interaction between MgO and Ni-species appears to maintain the formation of well-dispersed small Ni particles [51,52,53,54]. In addition, the increase in TOF value is consistent with the activity test results. The particle size remained quite stable (28.3–29.6 nm). The increases were the most pronounced at the highest proportion of Ni (60 wt%) because of the higher reactivity of the catalyst.

3.3. XRD Analysis

The elemental composition of the catalysts was determined with an X-ray diffractometer. The catalyst powder samples were heated at 250 °C for 5 h to remove moisture, and the catalyst crystals were analyzed. Cu-Kα radiation was used to fix the axis of the sample, and measurements were performed at 30 mA and 40 kV over a 2θ range of 10–80°. The XRD analysis results of the catalytic supports and Ni catalysts before and after the reaction are shown in Figure 12. The fresh catalysts (Figure 12a–c) produced no Ni peaks but rather were characterized by NiO peaks, whereas Ni peaks were evident in the patterns of the spent catalysts (Figure 12d–f). The Ni peaks (37.4°, 44.7°, 51.7°, and 76.3°), gamma Al2O3 diffraction peaks (39.5°, 46.1°, and 66.9°), and MgO were observed at 37.2°, 43.2°, and 62.4° (JCPDS No.04-0850). As the Ni loading was increased, the intensity of metallic Ni peaks increased (51.7° and 76.3°) [55]. The XRD MgO peak overlapped with the NiO and Al2O3 peaks, making it difficult to distinguish them, and no difference was found except for the intensity of the MgO peak before and after the reaction.

3.4. XPS Analysis

An X-ray photoelectron spectrometer (Thermo Scientific, Waltham, MA, USA, K-Alpha plus model) was used to determine binding energy (BE) values, which were used to evaluate the oxidation states of the catalysts. BE values for fresh and spent catalysts without pre-treatment measured under vacuum using Al-Kα (50 mV) radiation are shown in Figure 13. The oxidation states of Ni can be determined from the BE of the XPS Ni2p3/2, which varies 855 to 856 eV for NiO, 852.3 to 852.6 eV for metallic Ni and 861.5 to 865.6 for Ni(OH)2. All of the catalyst were calcined in the presence of atmospheric air, and humidity could be formed Ni(OH)2. As shown in Figure 13, the BE values in the fresh catalysts appeared mainly NiO and Ni(OH)2. In the spent catalysts, in addition to NiO and Ni(OH)2, metallic Ni emerged at 852.3 eV [56,57]. Metallic Ni was formed during reduction; however, this Ni species is stable and it can be easily oxidized when it is exposed to air, generating NiO and Ni(OH)2.

3.5. Transmission Electron Microscopy (TEM) Characterization

The metallic Ni particle size of the fresh and spent catalyst with 20, 40, and 60 wt% Ni were analyzed using a TEM. As shown in Figure 14a–f, the metallic Ni particles were suitably disperesed over the supports. The size of the Ni particle were approximately 27~32 nm, Which matched with the metallic Ni particle size obtained using the H2 chemisorption. It can be confirmed that there is no change and no carbon deposition, when compared with the fresh catalyst.
SEM-energy-dispersive X-ray spectroscopy (EDX) was used to investigate the surface morphology of the catalysts and the dispersion of Ni. An APREO SEM system (FEI, Hillsboro, OR, USA) was used to acquire images of the catalyst samples prepared by removing powders and dust, drying at 120 °C for 1 h and coating with metal (Au) at a magnification of 20,000×. Because of the decomposition of Mg(NO3)2 · 6H2O to MgO at temperatures above 600 °C, spherical particles were uniformly dispersed, as has been previously reported [40].
The results of the elemental composition analysis of the catalysts using EDX are shown in Table 4. The contents of Ni, Mg, and Al metals in cross-sections of the catalysts are summarized. The EDX analysis data indicated that the actual Ni contents of the catalysts used in the experiment were close to the theoretical values of 20, 40, and 60 wt%.

4. Experimental Methods

4.1. Methanation Reaction

CO2 methanation, which was proposed by Sabatier in 1902, is a high-pressure, low-temperature process for efficiently converting CO2 to CH4. In this process, hydrogen produced using a renewable energy source (e.g., solar) and CO2 produced from biomass are reacted to produce CH4 [38,39]. Because of the significant heat released (−165 kJ/mol) during the methanation process, CO2 conversion and CH4 selectivity decrease at temperatures above 627 °C. This is because, under such conditions, the change in Gibbs free energy is > 0, and thus CO2 is produced through the reaction of CH4 and H2O [40]. In theory, to ensure high CO2 conversion and CH4 selectivity, the process requires a low temperature [58,59]. In addition, a catalyst with a high efficiency at a low temperature is required to ensure optimal catalytic activity.
The CO2 methanation process can be divided into two steps. The first step involves the formation of carbonaceous intermediates through the reaction of CO2 with the catalyst. In the second step, CH4 is formed through the reaction of the carbonaceous intermediates on the catalyst surface with hydrogen species. Various reaction intermediates produced during CO2 methanation have been reported [60]. CO, which is produced through the adsorption of CO2 on the catalyst surface and subsequent decomposition, is reportedly the most likely intermediate during the methanation process [61]. The CO is then dissociated into C and O species, and the former is hydrogenated to produce CH4 [62,63]. The CO2 methanation process can be summarized as equations (1)–(3), with the reverse water gas shift (RWGS) reaction producing CO, which is then converted to CH4 via CO methanation [64].
CO2 + 4 H2 → CH4 + H2O  ∆H = –165 kJ/mol
CO2 + H2 → CO + H2O  ∆H = 42.1 kJ/mol
CO + 3 H2 → CH4 + H2O  ∆H = –206 kJ/mol
In the CO2 methanation process, the H2/CO2 ratio required for CH4 production is 4, and the reaction occurs at temperatures above 250 °C. Although the RWGS reaction is endothermic, it occurs more readily than the methanation reaction at temperatures above 350 °C because of its low enthalpy. However, because the CO methanation reaction is exothermic and its enthalpy is very high, the overall CO2 methanation process is exothermic, which is disadvantageous for reaction at high temperatures. Therefore, a high-activity catalyst suitable for low-temperature reactions is desirable for CO2 methanation.

4.2. Catalyst Synthesis and Experimentation

In the present study, Ni-Mg-Al catalysts characterized by a high activity and CH4 affinity were utilized. The synthesis process is illustrated in Figure 15. The catalysts were synthesized by mixing Ni(NO3)2·6H2O, Al(NO3)2·9H2O and Mg(NO3)2·6H2O, followed by stirring at 60 °C to produce catalysts with 20, 40, and 60 wt% Ni loadings. A precipitant was added to the mixture while maintaining a constant pH, and the solution was stirred for 1 h to induce precipitation. The precipitated catalyst precursor was repeatedly washed with distilled water and filtered using a press until the pH was approximately 7.0. The catalyst precursor was dried in an oven at 150 °C for 12 h and then heated at 600 °C for 4 h to obtain the 20, 40, and 60 wt% Ni-Mg-Al catalysts. The prepared catalyst is used as a reaction catalyst after raising the reduction temperature within 2 h, while flowing a mixed gas of 20% H2 and 80% N2 at 100 mL/min and maintaining it for 4 h. The CO2 conversion significantly increased as the reduction temperature of the Ni-Mg-Al catalysts increased in the 450–700 °C range, while changes in the 600–700 °C range were minor [41]. Therefore, 700 °C was selected as the reduction temperature in the present study.
The plug-flow system utilized as a reactor for experiments in the present study is shown in Figure 16. A temperature sensor was installed in the reactor, and an isothermal experiment was performed, while maintaining the reaction temperature with an error of ± 1.0 °C by the PID controller under atmospheric conditions (1 bar). The catalyst (0.5 g) was layered on the bottom of the reactor and a mesh was installed to support this layer. A water trap at the outlet of the reactor served for the removal of water generated as a byproduct, and a check valve was installed to prevent the backflow of gas. The products from the reactor were characterized using a gas chromatography (GC) system (YL Instrument 6500, Youngin chromass, Anyang, Korea), which was equipped with an SS COL 10 ft 1/8″ Matrix Porapak N (model: 13052-U) column with a 13X 45/60 mesh. The H2, CH4, and CO concentrations were measured using a thermal conductivity detector, while the CO2 concentration was analyzed using a flame ionization detector with a CO2 methanizer. For analyses, the GC oven was maintained at 35 °C for 6 min and then the temperature was raised to 170 °C at a ramp rate of 15 °C /min. Approximately 35 mL/min of H2 and 300 mL/min of O2 were injected into the flame ionization detector for analysis at 250 °C, while 35 and 20 mL/min of H2 and Ar, respectively, were analyzed using the thermal conductivity detector at 150 °C.
Experiments were conducted with a reactant gas flow rate of 250 mL/min, a reaction temperature of 350 °C, a GHSV of 30,000/h, and an H2/CO2 ratio of 4. Other parameters were varied as indicated in Table 5. The reaction temperature was varied between 200 °C and 450 °C and the GHSV between 10,000 and 50,000/h at different H2/CO2 ratios to evaluate their effects on CO2 and H2 conversion. The H2/CO2 ratio was adjusted with the balance gas N2. The effect of the initial CH4 concentration (0%~16.8%) of biogas was investigated by CO2 and H2 conversion and CH4 production.
CO2 conversion (XCO2, %) [21], H2 conversion (XH2, %), CH4 selectivity (SCH4, %), CH4 productivity (PCH4, %), and product (mole) were calculated using Equations (4)–(8).
X C O 2 ( % ) = 1 C O 2 ,   O C H 4 , o + C O O + C O 2 , O × 100
X C O 2 ( % ) = 1 C O 2 ,   O C H 4 , o + C O O + C O 2 , O × 100
S C H 4 ( % ) = C H 4 C H 4 + C O × 100
P C H 4 ( % ) = C H 4 ,   p r o d u c t C O 2 + H 2 + N 2 + ( C H 4   o r   C O ) × 100
P r o d u c t ( m o l e ) = C O 2 ( m o l e ) × X C O 2 × S C H 4

5. Conclusions

In the present study, CO2 methanation experiments were conducted over a Ni-Mg-Al catalyst to produce CH4 from biogas. The following results were obtained through experiments on the effect of catalyst reduction temperature, catalyst content, initial CH4 concentration in biogas, CO2 concentration, GHSV, and H2/CO2 ratio on CH4 production.
(1) As the reaction temperature increased, the CO2 conversion rate increased and then decreased after 400 °C. The highest CH4 selectivity and yield were obtained at approximately 350 °C. CO2 conversion decreased at temperatures above 400 °C because of the thermodynamic equilibrium limitation, while CH4 selectivity decreased at temperatures above 350 °C because of the RWGS reaction. These processes suppressed the methanation reaction and thus increased the amount of CO generated. The activation energy obtained for the methanation process in the present study was 59.7 kJ/mol;
(2) As the Ni content increased, the CO2 conversion rate, H2 conversion rate, and CH4 product increased. A CO2 conversion of 72% and H2 conversion of 45% were obtained using the catalyst containing 60 wt% Ni. This is because the Ni active sites increased as the Ni content increased and the Ni dispersion was well maintained uniformly; this was consistent with the increase in active Ni sites revealed by XRD analysis and the increase in TOF values, which reflect the reactivity;
(3) An increase in the initial CH4 concentration in the biogas from 0 to 16.8 vol% decreased CO2 conversion by 20%. This behavior is explained by Le Chatelier’s principle, where the conversion to CH4 is suppressed because of the initial CH4 among the reactants;
(4) As CO2 concentration in the reaction gas increased, H2 conversion, CH4 productivity and product increased. As the H2/CO2 ratio increased, H2 conversion decreased, while CH4 productivity and product increased slightly. As GHSV increased, H2 and CH4 productivity decreased slightly, while CH4 product increased with or without N2;
(5) A stability test of CO2 methanation over the Ni-Mg-Al catalyst containing 60 wt% Ni conducted at 350 °C for 130 h revealed a constant CO2 conversion and CH4 selectivity of 71% and 95%, respectively. These results demonstrated that the catalyst has good stability during CO2 methanation.

Author Contributions

Data curation: D.H.; investigation and writing—original draft: D.H.; writing—review and editing: Y.B. and D.H.; funding acquisition: W.C., validation: W.C. and Y.B.; supervision: Y.B. All authors have read and agreed to the published version of the manuscript.

Funding

This study was conducted with the support of the Korea Institute of Planning and Evaluation for Technology in Food, Agriculture, and Forest (IPET) and the Korean government (Ministry of Agriculture, Feed and Rural Affairs, 2021, grant No. 421038-03, Development of energy high-efficiency circulation and CO2 from fuel cell and carbon capture utilization for smart farm research project). Further, this study received support from the Korea Institute of Energy Technology Evaluation and Planning (KETEP) and the Korean government (Ministry of Trade, Industry and Energy, 2021, grant No. 20213030040270, Development and demonstration of hydrogen production process based on waste plastic non-oxidative pyrolysis).

Data Availability Statement

Not applicable.

Acknowledgments

We are grateful to Eun-Duk Park, a professor of chemical engineering at Ajou University, for help with the catalyst analysis and to Myeong-won Seo of the Korea Institute of Energy Research (KIER) for providing the catalyst.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gac, W.; Zawadzki, W.; Słowik, G.; Sienkiewicz, A.; Kierys, A. Nickel catalysts supported on silica microspheres for CO2 methanation. Microporous Mesoporous Mater. 2018, 272, 79–91. [Google Scholar] [CrossRef]
  2. Zhang, G.; Liu, J.; Xu, Y.; Sun, Y. A review of CH4CO2 reforming to synthesis gas over Ni-based catalysts in recent years (2010–2017). Int. J. Hydrogen Energy 2018, 43, 15030–15054. [Google Scholar] [CrossRef]
  3. 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]
  4. 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]
  5. Benson, E.E.; Kubiak, C.P.; Sathrum, A.J.; Smieja, J.M. Electrocatalytic and homogeneous approaches to conversion of CO2 to liquid fuels. Chem. Soc. Rev. 2009, 38, 89–99. [Google Scholar] [CrossRef]
  6. Solis-Garcia, A.; Louvier-Hernandez, J.F.; Almendarez-Camarillo, A.; Fierro-Gonzalez, J. Participation of surface bicarbonate, formate and methoxy species in the carbon dioxide methanation catalyzed by ZrO2-supported Ni. Appl. Catal. B Environ. 2017, 218, 611–620. [Google Scholar] [CrossRef]
  7. Gnanakumar, E.S.; Chandran, N.; Kozhevnikov, I.V.; Grau-Atienza, A.; Fernández, E.V.R.; Sepulveda-Escribano, A.; Shiju, N.R. Highly efficient nickel-niobia composite catalysts for hydrogenation of CO2 to methane. Chem. Eng. Sci. 2019, 194, 2–9. [Google Scholar] [CrossRef]
  8. Panagiotopoulou, P. Hydrogenation of CO2 over supported noble metal catalysts. Appl. Catal. A Gen. 2017, 542, 63–70. [Google Scholar] [CrossRef]
  9. 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]
  10. 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]
  11. Liu, H.; Xu, S.; Zhou, G.; Xiong, K.; Jiao, Z.; Wang, S. CO2 hydrogenation to methane over Co/KIT-6 catalysts: Effect of Co content. Fuel 2018, 217, 570–576. [Google Scholar] [CrossRef]
  12. Veselovskaya, J.V.; Parunin, P.D.; Netskina, O.V.; Kibis, L.S.; Lysikov, A.I.; Okunev, A.G. Catalytic methanation of carbon dioxide captured from ambient air. Energy 2018, 159, 766–773. [Google Scholar] [CrossRef]
  13. Siakavelas, G.I.; Charisiou, N.D.; AlKhoori, A.; AlKhoori, S.; Sebastian, V.; Hinder, S.J.; Baker, M.A.; Yentekakis, I.V.; Polychronopoulou, K.; Goula, M.A. Highly selective and stable Ni/La-M (M=Sm, Pr, and Mg)-CeO2 catalysts for CO2 methanation. J. CO2 Util. 2021, 51, 101618. [Google Scholar] [CrossRef]
  14. Ahmad, W.; Younis, M.N.; Shawabkeh, R.; Ahmed, S. Synthesis of lanthanide series (La, Ce, Pr, Eu & Gd) promoted Ni/γ-Al2O3 catalysts for methanation of CO2 at low temperature under atmospheric pressure. Catal. Commun. 2017, 100, 121–126. [Google Scholar]
  15. Ferreira, A.C.; Branco, J.B. Methanation of CO2 over nanostructured nickel-4f block element bimetallic oxides. Int. J. Hydrogen Energy 2019, 44, 6505–6513. [Google Scholar] [CrossRef]
  16. Li, S.; Liu, G.; Zhang, S.; An, K.; Ma, Z.; Wang, L.; Liu, Y. Cerium-modified Ni-La2O3/ZrO2 for CO2 methanation. J. Energy Chem. 2020, 43, 155–164. [Google Scholar] [CrossRef]
  17. 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]
  18. Bacariza, M.C.; Graça, I.; Bebiano, S.S.; Lopes, J.M.; Henriques, C. Micro- and mesoporous supports for CO2 methanation catalysts: A comparison between SBA-15, MCM-41 and USY zeolite. Chem. Eng. Sci. 2018, 175, 72–83. [Google Scholar] [CrossRef]
  19. Liu, W.; Li, L.; Zhang, X.; Wang, Z.; Wang, X.; Peng, H. Design of Ni-ZrO2@SiO2 catalyst with ultra-high sintering and coking resistance for dry reforming of methane to prepare syngas. J. CO2 Util. 2018, 27, 297–307. [Google Scholar] [CrossRef]
  20. Romero-Sáez, M.; Dongil, A.; Benito, N.; Espinoza-González, R.; Escalona, N.; Gracia, F. CO2 methanation over nickel-ZrO2 catalyst supported on carbon nanotubes: A comparison between two impregnation strategies. Appl. Catal. B Environ. 2018, 237, 817–825. [Google Scholar] [CrossRef]
  21. Ocampo, F.; Louis, B.; Kiwi-Minsker, L.; Roger, A.-C. Effect of Ce/Zr composition and noble metal promotion on nickel based CexZr1−xO2 catalysts for carbon dioxide methanation. Appl. Catal. A Gen. 2011, 392, 36–44. [Google Scholar] [CrossRef]
  22. Garbarino, G.; Wang, C.; Cavattoni, T.; Finocchio, E.; Riani, P.; Flytzani-Stephanopoulos, M.; Busca, G. A study of Ni/La-Al2O3 catalysts: A competitive system for CO2 methanation. Appl. Catal. B Environ. 2019, 248, 286–297. [Google Scholar] [CrossRef]
  23. 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. Hydrogen Energy 2012, 37, 5527–5531. [Google Scholar] [CrossRef]
  24. Löfberg, A.; Guerrero-Caballero, J.; Kane, T.; Rubbens, A.; Jalowiecki-Duhamel, L. Ni/CeO2 based catalysts as oxygen vectors for the chemical looping dry reforming of methane for syngas production. Appl. Catal. B Environ. 2017, 212, 159–174. [Google Scholar] [CrossRef]
  25. Ratchahat, S.; Sudoh, M.; Suzuki, Y.; Kawasaki, W.; Watanabe, R.; Fukuhara, C. Development of a powerful CO2 methanation process using a structured Ni/CeO2 catalyst. J. CO2 Util. 2018, 24, 210–219. [Google Scholar] [CrossRef]
  26. Yan, X.; Hu, T.; Liu, P.; Li, S.; Zhao, B.; Zhang, Q.; Jiao, W.; Chen, S.; Wang, P.; Lu, J.; et al. Highly efficient and stable Ni/CeO2-SiO2 catalyst for dry reforming of methane: Effect of interfacial structure of Ni/CeO2 on SiO2. Appl. Catal. B Environ. 2019, 246, 221–231. [Google Scholar] [CrossRef]
  27. Akbari, E.; Alavi, S.M.; Rezaei, M. CeO2 Promoted Ni-MgO-Al2O3 nanocatalysts for carbon dioxide reforming of methane. J. CO2 Util. 2018, 24, 128–138. [Google Scholar] [CrossRef]
  28. Ray, K.; Bhardwaj, R.; Singh, B.; Deo, G. Developing descriptors for CO2 methanation and CO2 reforming of CH4 over Al2O3 supported Ni and low-cost Ni based alloy catalysts. Phys. Chem. Chem. Phys. 2018, 20, 15939–15950. [Google Scholar] [CrossRef]
  29. Ahn, J.Y.; Chang, S.W.; Lee, S.M.; Kim, S.S.; Chung, W.J.; Lee, J.C.; Cho, Y.J.; Shin, K.S.; Moon, D.H.; Nguyen, D.D. Developing Ni-based honeycomb-type catalysts using different binary oxide-supported species for synergistically enhanced CO2 methanation activity. Fuel 2019, 250, 277–284. [Google Scholar] [CrossRef]
  30. Moghaddam, S.V.; Rezaei, M.; Meshkani, F.; Daroughegi, R. Carbon dioxide methanation over Ni-M/Al2O3 (M: Fe, CO, Zr, La and Cu) catalysts synthesized using the one-pot sol-gel synthesis method. Int. J. Hydrogen Energy 2018, 43, 16522–16533. [Google Scholar] [CrossRef]
  31. 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]
  32. Aziz, M.A.A.; Jalil, A.A.; Triwahyono, S.; Ahmad, A. CO2 methanation over heterogeneous catalysts: Recent progress and future prospects. Green Chem. 2015, 17, 2647–2663. [Google Scholar] [CrossRef]
  33. Guo, M.; Lu, G. The effect of impregnation strategy on structural characters and CO2 methanation properties over MgO modified Ni/SiO2 catalysts. Catal. Commun. 2014, 54, 55–60. [Google Scholar] [CrossRef]
  34. 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]
  35. Al-Fatesh, A.S.; Abu-Dahrieh, J.K.; Atia, H.; Armbruster, U.; Ibrahim, A.A.; Khan, W.U.; Abasaeed, A.E.; Fakeeha, A.H. Effect of pre-treatment and calcination temperature on Al2O3-ZrO2 supported Ni-Co catalysts for dry reforming of methane. Int. J. Hydrogen Energy 2019, 44, 21546–21558. [Google Scholar] [CrossRef]
  36. Feng, X.; Feng, J.; Li, W. Insight into MgO promoter with low concentration for the carbon-deposition resistance of Ni-based catalysts in the CO2 reforming of CH4. Chin. J. Catal. 2018, 39, 88–98. [Google Scholar] [CrossRef]
  37. Vázquez, F.V.; Kihlman, J.; Mylvaganam, A.; Simell, P.; Koskinen-Soivi, M.-L.; Alopaeus, V. Modeling of nickel-based hydrotalcite catalyst coated on heat exchanger reactors for CO2 methanation. Chem. Eng. J. 2018, 349, 694–707. [Google Scholar] [CrossRef]
  38. Su, X.; Xu, J.; Liang, B.; Duan, H.; Hou, B.; Huang, Y. Catalytic carbon dioxide hydrogenation to methane: A review of recent studies. J. Energy Chem. 2016, 25, 553–565. [Google Scholar] [CrossRef]
  39. Champon, I.; Bengaouer, A.; Chaise, A.; Thomas, S.; Roger, A.-C. Carbon dioxide methanation kinetic model on a commercial Ni/Al2O3 catalyst. J. CO2 Util. 2019, 34, 256–265. [Google Scholar] [CrossRef]
  40. Hu, F.; Tong, S.; Lu, K.; Chen, C.-M.; Su, F.-Y.; Zhou, J.; Lu, Z.-H.; Wang, X.; Feng, G.; Zhang, R. Reduced graphene oxide supported Ni-Ce catalysts for CO2 methanation: The support and ceria promotion effects. J. CO2 Util. 2019, 34, 676–687. [Google Scholar] [CrossRef]
  41. Han, D.; Kim, Y.; Byun, H.; Cho, W.; Baek, Y. CO2 Methanation of Biogas over 20 wt% Ni-Mg-Al Catalyst: On the Effect of N2, CH4, and O2 on CO2 Conversion Rate. Catalysts 2020, 10, 1201. [Google Scholar] [CrossRef]
  42. Wierzbicki, D.; Baran, R.; Dębek, R.; Motak, M.; Grzybek, T.; Gálvez, M.E.; Da Costa, P. The influence of nickel content on the performance of hydrotalcite-derived catalysts in CO2 methanation reaction. Int. J. Hydrogen Energy 2017, 42, 23548–23555. [Google Scholar] [CrossRef]
  43. Zhang, Z.; Tian, Y.; Zhang, L.; Hu, S.; Xiang, J.; Wang, Y.; Xu, L.; Liu, Q.; Zhang, S.; Hu, X. Impacts of nickel loading on properties, catalytic behaviors of Ni/γ–Al2O3 catalysts and the reaction intermediates formed in methanation of CO2. Int. J. Hydrogen Energy 2019, 44, 9291–9306. [Google Scholar] [CrossRef]
  44. 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 2019, 356, 419–432. [Google Scholar] [CrossRef]
  45. Mutz, B.; Carvalho, H.W.; Mangold, S.; Kleist, W.; Grunwaldt, J.-D. Methanation of CO2: Structural response of a Ni-based catalyst under fluctuating reaction conditions unraveled by operando spectroscopy. J. Catal. 2015, 327, 48–53. [Google Scholar] [CrossRef]
  46. Peymani, M.; Alavi, S.M.; Rezaei, M. Synthesis Gas Production by Catalytic Partial Oxidation of Propane on Mesoporous Nanocrystalline Ni/Al2O3Catalysts. Appl. Catal. A Gen. 2017, 529, 1–9. [Google Scholar] [CrossRef]
  47. Jia, X.; Zhang, X.; Rui, N.; Hu, X.; Liu, C. Structural effect of Ni/ZrO2 catalyst on CO2 methanation with enhanced activity. Appl. Catal. B Environ. 2019, 244, 159–169. [Google Scholar] [CrossRef]
  48. Dannesboe, C.; Hansen, J.B.; Johannsen, I. Catalytic methanation of CO2 in biogas: Experimental results from a reactor at full scale. React. Chem. Eng. 2020, 5, 183–189. [Google Scholar] [CrossRef]
  49. Jürgensen, L.; Ehimen, E.A.; Born, J.; BoHolm-Nielsena, J. Dynamic biogas upgrading based on the Sabatier process: Thermodynamic and dynamic process simulation. Bioresour. Technol. 2015, 178, 323–329. [Google Scholar] [CrossRef]
  50. Jaffar, M.M.; Nahil, M.A.; Williams, P.T. Parametric study of CO2 methanation for synthetic natural gas production. Energy Technol. 2019, 7, 1900795. [Google Scholar] [CrossRef]
  51. Erdöhelyi, A.; Pásztor, M.; Solymosi, F. Catalytic hydrogenation of CO2 over supported palladium. J. Catal. 1986, 98, 166–177. [Google Scholar] [CrossRef] [Green Version]
  52. Agnelli, M.; Kolb, M.; Mirodatos, C. Co Hydrogenation on a Nickel Catalyst: 1. Kinetics and Modeling of a Low-Temperature Sintering Process. J. Catal. 1994, 148, 9–21. [Google Scholar] [CrossRef]
  53. Tan, M.; Wang, X.; Wang, X.; Zou, X.; Ding, W.; Lu, X. Influence of calcination temperature on textural and structural properties, reducibility, and catalytic behavior of mesoporous γ-alumina-supported Ni–Mg oxides by one-pot template-free route. J. Catal. 2015, 329, 151–166. [Google Scholar] [CrossRef]
  54. Nakayama, T.; Ichikuni, N.; Sato, S.; Nozaki, F. Ni/Mgo catalyst prepared using citric acid for hydrogenation of carbon dioxide. Appl. Catal. A Gen. 1997, 158, 185–199. [Google Scholar] [CrossRef]
  55. Carbarino, G.; Riani, P.; Magistri, L.; Busca, G. A study of the methanation of carbon dioxide on Ni/Al2O3 catalysts at atmospheric pressure. Int. J. Hydrogen Energy 2014, 39, 11557–11565. [Google Scholar] [CrossRef]
  56. 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. Hydrogen Energy 2017, 42, 4174–4183. [Google Scholar] [CrossRef]
  57. Biesinger, M.C.; Payne, B.P.; Lau, L.W.M.; Gerson, A.; Smart, R.S.C. X-ray photoelectron spectroscopic chemical state quantification of mixed nickel metal, oxide and hydroxide systems. Surf. Interface Anal. 2009, 41, 324–332. [Google Scholar] [CrossRef]
  58. Jiang, H.; Gao, Q.; Wang, S.; Chen, Y.; Zhang, M. The synergistic effect of Pd NPs and UiO-66 for enhanced activity of carbon dioxide methanation. J. CO2 Util. 2019, 31, 167–172. [Google Scholar] [CrossRef]
  59. Ou, Z.; Qin, C.; Niu, J.; Zhang, L.; Ran, J. A comprehensive DFT study of CO2 catalytic conversion by H2 over Pt-doped Ni catalysts. Int. J. Hydrogen Energy 2018, 44, 819–834. [Google Scholar] [CrossRef]
  60. Lim, J.Y.; McGregor, J.; Sederman, A.; Dennis, J. 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]
  61. Westermann, A.; Azambre, B.; Bacariza, M.C.; Graça, I.; Ribeiro, M.F.; Lopes, J.M.; Henriques, C. Insight into CO2 methanation mechanism over NiUSY zeolites: An operando IR study. Appl. Catal. B Environ. 2015, 174–175, 120–125. [Google Scholar] [CrossRef]
  62. Miguel, C.; Mendes, A.; Madeira, L. Intrinsic kinetics of CO2 methanation over an industrial nickel-based catalyst. J. CO2 Util. 2018, 25, 128–136. [Google Scholar] [CrossRef]
  63. Li, Y.; Zhang, H.; Zhang, L.; Zhang, H. Bimetallic Ni–Pd/SBA-15 alloy as an effective catalyst for selective hydrogenation of CO2 to methane. Int. J. Hydrogen Energy 2019, 44, 13354–13363. [Google Scholar] [CrossRef]
  64. 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]
Figure 1. Effect of reduction temperature on CO2 conversion and CH4 selectivity at 350 °C (for 20 wt%, reference [41] was cited).
Figure 1. Effect of reduction temperature on CO2 conversion and CH4 selectivity at 350 °C (for 20 wt%, reference [41] was cited).
Catalysts 12 01054 g001
Figure 2. Effect of reaction temperature on CO2 conversion and CH4 selectivity.
Figure 2. Effect of reaction temperature on CO2 conversion and CH4 selectivity.
Catalysts 12 01054 g002
Figure 3. Arrhenius plot for CO2 methanation over 60 wt% Ni-Mg-Al catalyst.
Figure 3. Arrhenius plot for CO2 methanation over 60 wt% Ni-Mg-Al catalyst.
Catalysts 12 01054 g003
Figure 4. Effect of CO2 concentration on CO2 conversion and CH4 selectivity.
Figure 4. Effect of CO2 concentration on CO2 conversion and CH4 selectivity.
Catalysts 12 01054 g004
Figure 5. Effect of H2/CO2 ratio on CO2 conversion and CH4 selectivity.
Figure 5. Effect of H2/CO2 ratio on CO2 conversion and CH4 selectivity.
Catalysts 12 01054 g005
Figure 6. Effect of initial CH4 concentration on CO2 and H2 conversion and CH4 selectivity.
Figure 6. Effect of initial CH4 concentration on CO2 and H2 conversion and CH4 selectivity.
Catalysts 12 01054 g006
Figure 7. Effect of CO2 concentrations on the H2 conversion and CH4 production.(GHSV = 30,000/h, H2/CO2 = 4).
Figure 7. Effect of CO2 concentrations on the H2 conversion and CH4 production.(GHSV = 30,000/h, H2/CO2 = 4).
Catalysts 12 01054 g007
Figure 8. Effect of H2/CO2 ratio on the H2 conversion and CH4 production.(GHSV = 30,000/h).
Figure 8. Effect of H2/CO2 ratio on the H2 conversion and CH4 production.(GHSV = 30,000/h).
Catalysts 12 01054 g008
Figure 9. Effect of GHSV on the CH4 productivity and CH4 products (H2/CO2 = 4).
Figure 9. Effect of GHSV on the CH4 productivity and CH4 products (H2/CO2 = 4).
Catalysts 12 01054 g009
Figure 10. The durability and activity test of the catalyst for 130 h.
Figure 10. The durability and activity test of the catalyst for 130 h.
Catalysts 12 01054 g010
Figure 12. XRD patterns of fresh Ni-Mg-Al catalysts (a) 20 wt% Ni (b) 40 wt% Ni, or (c) 60 wt% Ni and spent catalysts (d) 20 wt% Ni (e) 40 wt% Ni, or (f) 60 wt% Ni ((A) fresh, (B) spent)(* reference [41] was cited).
Figure 12. XRD patterns of fresh Ni-Mg-Al catalysts (a) 20 wt% Ni (b) 40 wt% Ni, or (c) 60 wt% Ni and spent catalysts (d) 20 wt% Ni (e) 40 wt% Ni, or (f) 60 wt% Ni ((A) fresh, (B) spent)(* reference [41] was cited).
Catalysts 12 01054 g012
Figure 13. XPS patterns of fresh Ni-Mg-Al catalysts (a) 20 wt% Ni, (b) 40 wt% Ni, (c) 60 wt% Ni, (d) 60wt% Mg and spent catalysts (e) 20 wt% Ni, (f) 40 wt% Ni, and (g) 60 wt% Ni, (h) 60wt% Mg(* for 20 wt%, Reference [41] was cited.).
Figure 13. XPS patterns of fresh Ni-Mg-Al catalysts (a) 20 wt% Ni, (b) 40 wt% Ni, (c) 60 wt% Ni, (d) 60wt% Mg and spent catalysts (e) 20 wt% Ni, (f) 40 wt% Ni, and (g) 60 wt% Ni, (h) 60wt% Mg(* for 20 wt%, Reference [41] was cited.).
Catalysts 12 01054 g013aCatalysts 12 01054 g013b
Figure 14. TEM images of reduced Ni-Mg-Al catalysts (a) 20 wt%, (b) 40 wt%, (c) 60 wt% Ni and spent Ni-Mg-Al catalyst (d) 20 wt%, (e) 40 wt%, (f) 60 wt%.
Figure 14. TEM images of reduced Ni-Mg-Al catalysts (a) 20 wt%, (b) 40 wt%, (c) 60 wt% Ni and spent Ni-Mg-Al catalyst (d) 20 wt%, (e) 40 wt%, (f) 60 wt%.
Catalysts 12 01054 g014aCatalysts 12 01054 g014bCatalysts 12 01054 g014c
Figure 15. Block diagram of the process for synthesizing the Ni-catalyst used for CO2 methanation.
Figure 15. Block diagram of the process for synthesizing the Ni-catalyst used for CO2 methanation.
Catalysts 12 01054 g015
Figure 16. Schematic representation of the plug-flow system used for CO2 methanation.
Figure 16. Schematic representation of the plug-flow system used for CO2 methanation.
Catalysts 12 01054 g016
Table 1. Reaction temperatures to reach 50% CO2 conversion.
Table 1. Reaction temperatures to reach 50% CO2 conversion.
Ni Catalyst20 wt% Ni-Mg-Al40 wt% Ni-Mg-Al60 wt% Ni-Mg-Al
Reaction temperature (°C)340298291
Table 2. BET specific surface area, pore volume, and pore size of Ni-Mg-Al catalysts containing different proportions of Ni before (fresh) and after (spent) experiments.
Table 2. BET specific surface area, pore volume, and pore size of Ni-Mg-Al catalysts containing different proportions of Ni before (fresh) and after (spent) experiments.
Ni CatalystBET (m2/g)Total Pore Volume (m3/g)Pore Size (Å)
* 20
(wt%)
fresh180.30.3681.5
spent148.90.3081.1
40
(wt%)
fresh155.80.3177.4
spent107.80.3282.4
60
(wt%)
fresh140.90.3291.5
spent110.80.2385.9
* 20 wt%, 40 wt% data was cited reference [41].
Table 3. Ni dispersion, particle size, and TOF for the 20, 40, and 60 wt% Ni-Mg-Al catalysts.
Table 3. Ni dispersion, particle size, and TOF for the 20, 40, and 60 wt% Ni-Mg-Al catalysts.
Ni CatalystNi Dispersion (%)Ni Particle Size (nm)TOF (s−1)
20 wt% *3.5728.30.01093
40 wt% *3.6827.50.02366
60 wt%3.4229.60.02593
* Reference [41] was cited.
Table 4. SEM-EDX analysis results for fresh Ni-Mg-Al catalysts.
Table 4. SEM-EDX analysis results for fresh Ni-Mg-Al catalysts.
ItemsNi (wt%)Mg (wt%)Al2O3(wt%)
20 wt%15.9 (20)3.2 (5)79.8 (75)
40 wt%36.8 (40)2.0 (5)61.2 (55)
60 wt%41.5 (60)1.0 (5)57.5 (35)
( ): theoretical amount
Table 5. Experimental conditions used for the methanation reaction.
Table 5. Experimental conditions used for the methanation reaction.
ParameterConditions
Temperature (°C)200~450
Pressure(bar)1
GHSV (/h)10,000~50,000
N2 (vol%)Balance gas
CO2 (vol%)6, 10, 16, 20
H2/CO23.5, 4, 4.5, 5
CH4 (vol%)0, 6.3, 10.8, 16.8
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Han, D.; Cho, W.; Baek, Y. CO2 Methanation of Biogas over Ni-Mg-Al: The Effects of Ni Content, Reduction Temperature, and Biogas Composition. Catalysts 2022, 12, 1054. https://doi.org/10.3390/catal12091054

AMA Style

Han D, Cho W, Baek Y. CO2 Methanation of Biogas over Ni-Mg-Al: The Effects of Ni Content, Reduction Temperature, and Biogas Composition. Catalysts. 2022; 12(9):1054. https://doi.org/10.3390/catal12091054

Chicago/Turabian Style

Han, Danbee, Wonjun Cho, and Youngsoon Baek. 2022. "CO2 Methanation of Biogas over Ni-Mg-Al: The Effects of Ni Content, Reduction Temperature, and Biogas Composition" Catalysts 12, no. 9: 1054. https://doi.org/10.3390/catal12091054

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