Next Article in Journal
Regioselective Hydroxylation of Rhododendrol by CYP102A1 and Tyrosinase
Next Article in Special Issue
Unravelling the Structural Modification (Meso-Nano-) of Cu/ZnO-Al2O3 Catalysts for Methanol Synthesis by the Residual NaNO3 in Hydroxycarbonate Precursors
Previous Article in Journal
Recent Advancements and Future Prospects in Ultrathin 2D Semiconductor-Based Photocatalysts for Water Splitting
Previous Article in Special Issue
Direct Synthesis of Dimethyl Ether from Syngas on Bifunctional Hybrid Catalysts Based on Supported H3PW12O40 and Cu-ZnO(Al): Effect of Heteropolyacid Loading on Hybrid Structure and Catalytic Activity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Experimental Study on CO2 Methanation over Ni/Al2O3, Ru/Al2O3, and Ru-Ni/Al2O3 Catalysts

Department of Mechanical Engineering, National Chung Hsing University, Taichung 40227, Taiwan
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(10), 1112; https://doi.org/10.3390/catal10101112
Submission received: 18 July 2020 / Revised: 22 September 2020 / Accepted: 22 September 2020 / Published: 25 September 2020
(This article belongs to the Special Issue Catalysts for Production and Conversion of Syngas)

Abstract

:
CO2 methanation is recognized as one of the best technologies for storing intermittent renewable energy in the form of CH4. In this study, CO2 methanation performance is investigated using Ni/Al2O3, Ru/Al2O3, and Ru-Ni/Al2O3 as the catalysts under conditions of atmospheric pressure, a molar ratio of H2/CO2 = 5, and a space velocity of 5835 h−1. For reaction temperatures ranging from 250 to 550 °C, it was found that the optimum reaction temperature is 400 °C for all catalysts studied. At this temperature, the maximum values of CO2 conversion, H2 efficiency, and CH4 yield and lowest CO yield can be obtained. With temperatures higher than 400 °C, reverse CO2 methanation results in CO2 conversion and CH4 yield decreases with increased temperature, while CO is formed due to reverse water-gas shift reaction. The experimental results showed that CO2 methanation performance at low temperatures can be enhanced greatly using the bimetallic Ru-Ni catalyst compared with the monometallic Ru or Ni catalyst. Under ascending-descending temperature changes between 250 °C and 550 °C, good thermal stability is obtained from Ru-Ni/Al2O3 catalyst. About a 3% decrease in CO2 conversion is found after three continuous cycles (74 h) test.

1. Introduction

The efficient utilization of renewable energy resources from wind and sun is an ongoing challenge. Electrical power from wind and solar devices is intermittent and produced at locations where it is not directly consumed. Thus, solutions for long-range renewable energy transportation and storage are currently of great interest [1,2,3]. Power-to-Gas (PtG) technology might contribute to solving this issue [4,5]. PtG links the power grid with the gas grid by converting surplus power into a grid compatible gas via a two-step process: H2 production by water electrolysis and H2 conversion with captured CO2 from combustion or other sources into CH4 via methanation [6]. The resulting CH4 is known as substitute natural gas (SNG). It can be either injected into the existing gas distribution grid, stored, or utilized in other well established natural gas facilities.
The stoichiometric CO2 methanation reaction is:
CO 2 + 4 H 2 CH 4 + 2 H 2 O ;   Δ H 298 K = 165   kJ / mole
CO2 methanation is composed of two successive reactions: reverse water-gas shift reaction (RWGS) and CO methanation:
RWGS: CO 2 + H 2 CO + H 2 O ,   Δ H 298 K = 42   kJ / mole
CO methanation: CO + 3 H 2 CH 4 + H 2 O ,   Δ H 298 K = 206   kJ / mole
Note also that CO2 methanation is exothermic. Thus, lower reaction temperature is desirable for higher CH4 selectivity. It is therefore important to enhance the RWGS reaction rate and CO methanation at low temperatures. Different mechanisms have been proposed for CO2 methanation. Eckle et al. [7] and Fisher et al. [8] proposed a CO2 methanation mechanism that involves direct dissociation of CO2 to CO(ads) and O(ads) on the metal surface, with CO(ads) subsequently hydrogenated into CH4. CO2 dissociation has been generally recognized as the rate-determining step of the reaction [9,10]. The CO2 methanation reaction is also known as the Sabatier reaction and has been studied extensively using different types of metals and supports [11,12,13]. Ni-, Ru-, and Rh-based catalysts have been proven to have good catalytic activity in CO2 methanation [14,15,16]. Among these metals, Ni-based catalysts are the most commonly used catalysts because of their high activity and CH4 selectivity. It was also found that Ni-based catalysts could maintain very good activity over a long reaction time with high CH4 selectivity. However, high CO2 conversion is difficult to achieve at low temperatures on Ni catalysts because high activation energy is required [17]. Compared with Ni, Ru-based catalysts are more active in CO2 methanation and produce CH4 almost exclusively. Moreover, it has also been shown that Ru-based catalysts have good activity at low temperatures [18,19,20].
Some efforts have been made to develop effective Ni-based catalysts with satisfactory catalytic activity for CO2 methanation at low temperatures. Rahmani et al. [21] pointed out that Ni dispersion can be enhanced by the addition of Ce. In the study by Lee et al. [22], a Ni-based catalyst prepared using dielectric barrier discharge (DBD) plasma produced higher metal dispersion and fewer defect sites than the impregnation method. Both studies pointed out that high Ni dispersion can enhance Ni catalyst activity at low temperatures. In a study by Xu et al. [23], a series of rare-earth metals (La, Ce, Sm, and Pr) were doped onto Ni-based catalysts and used in CO2 methanation. It was found that the apparent activation energy could be decreased using rare-earth dopants and that low-temperature catalytic activity could be greatly intensified over these rare-earth element catalysts. Additionally, some studies showed that Ni support is one of the factors that affect Ni activity at low temperatures. The catalytic activity for CO2 methanation of Ni with various metal oxides as supports was evaluated in a study by Muroyama et al. [24]. They found that the order of CH4 yield at 250 °C was Ni/Y2O3 > Ni/Sm2O3 > Ni/ZrO2 > Ni/CeO2 > Ni/Al2O3 > Ni/La2O3. In a study by Abate et al. [25], CO2 methanation performance was studied using Ni supported by γ-Al2O3-ZrO2-TiO2-CeO2 composite oxide. Their experimental results revealed that the performance of composite oxide-supported Ni-based catalysts was superior to that of only γ-Al2O3-supported Ni-based catalysts.
Although Ru-based catalysts have been proven to present good activity in low-temperature CO2 methanation reactions, their performance at high temperatures has not been widely reported in the literature. Conversely, Ni has low activity at low temperatures but good activity at high temperatures. Based on these observations, bimetallic Ru-Ni catalysts could combine the advantages of each catalyst for use in a wide range of reaction temperatures for CO2 methanation. The Ru-Ni bimetallic catalyst has been widely adopted in many chemical reactions such as steam reforming of methane [26], dry reforming of methane [27], and CO cleanup in hydrogen-rich reformed gas for fuel cell applications [28,29]. Ru is the cheapest precious metal. Thus, Ru-Ni can keep the catalyst cost sufficiently low for industrial implementation. Polanski et al. [30] tested the catalytic performance of CO2 methanation using Ni-supported Ru nanoparticles in silica. Their result indicated that the Ru-Ni catalyst is very productive and efficient for CO2 methanation at low temperatures. In the study by Zhen et al. [31], bimetallic Ni-Ru catalysts supported on γ-Al2O3 were prepared using co-impregnation and sequential impregnation methods and used to investigate CO2 methanation performance for temperatures in the 250–500 °C range under atmospheric pressure conditions. They found that the CO2 methanation activities were highly dependent on the bimetallic catalyst preparation sequence. In a study by Navarro et al. [32], a structured Ru-Ni/MgAl2O4 catalyst was proposed for CO2 methanation. Their results indicated that the catalyst exhibited high stability in a long-term test.
Although Ru-Ni catalysts have been successfully applied in many applications, their efficacy in CO2 methanation involving higher CO2 concentration needs to be clarified. For further insights, Ru/Al2O3, Ni/Al2O3, and Ru–Ni/Al2O3 catalysts with various Ru and Ni loadings were prepared and tested for CO2 methanation in this study. The metal loading and reaction temperature effects on CO2 conversion, H2 efficiency, CH4 yield, and CO yield were investigated in detail. The thermal stability of the prepared catalyst is also reported in this study.

2. Results and Discussion

2.1. Experimental Parameters

A catalyst with a weight of 0.5 g was used in this study, resulting in a reaction packed bed length of 4.2 cm. Accordingly, the catalyst bed length to catalyst particle size ratio and the ratio of the inside reactor diameter to particle size were in the ranges of 35~84 and 3~8, respectively. This ensured that the back mixing and channeling effects were minimized in the packed-bed reactor [33].
The activity test was performed from 250 to 550 °C in 50 °C increments and operated at atmospheric pressure. The reactant volume flow rate was fixed at 55 sccm, which was composed of 5 sccm CO2, 25 sccm H2, and 25 sccm N2. This corresponded to a molar ratio of H2/CO2 = 5, while N2 served as the balance inert gas and to avoid the thermal runaway since CO2 methanation is an exothermic reaction [34,35,36]. Moreover, the N2 flow rate was used as the reference flow to evaluate the flow rates of the gas species involved in the product. Based on the total reactant flow rate and catalyst bed volume, the corresponding gas hourly space velocity (GHSV) was 5835 h−1. After removing water using a condenser, the dried product gas mixture was collected every 30 min for each temperature. To ensure stable reaction performance, the time on stream measurement for each temperature lasted at least 2 h.

2.2. Performance Indices

The CO2 methanation performance was characterized using CO2 conversion, H2 efficiency, CH4 yield, and CO yield, defined as follows,
C O 2  conversion:  X CO 2 = F ˙ CO 2 , in F ˙ CO 2 , out F ˙ CO 2 , in × 100 %
H 2  efficiency:  η H 2 = ( F ˙ CO 2 , in F ˙ CO 2 , out ) + ( F ˙ CO , out F ˙ CO , in ) F ˙ H 2 , in × 100 %
C H 4  yield:  Y CH 4 = F ˙ CH 4 , out F ˙ CO 2 , in × 100 %
CO yield:  Y CO = F ˙ CO , out F ˙ CO 2 , in × 100 %
In these equations, F ˙ CO 2 , in is the CO2 molar flow rate at the reactor inlet, F ˙ CO 2 , out , F ˙ CH 4 , out , and F ˙ CO , out are the molar flow rates of CO2, CH4, and CO at the reactor outlet, respectively. In Equation (4), CO2 conversion is defined based on CO2 consumption. H2 is expected to be the limiting reactant in large scale applications. Therefore, it is important to determine the amount of CO2 conversion compared to the amount of H2 available. This is defined as the H2 efficiency, as shown in Equation (5). Note that CO formation from converted CO2 is included in the H2 efficiency. In the CO2 methanation reaction, the CO2 is regarded as the feedstock, while CH4 and CO are the products from the reaction. Instead of selectivity, the inlet molar flow rate of CO2 is used as the reference to define the yields of CH4 and CO, as shown in Equations (6) and (7). The product yield for every mole of CO2 supply can be identified. The CH4 selectivity can be evaluated from CO2 conversion and CH4 yield, as defined in Equations (4) and (6), respectively.

2.3. Catalyst Characterization

FE-SEM-EDX (JSM-7401F, JEOL, (Tokyo, Japan) analysis was used to analyze the general morphology and the chemical composition of the prepared catalysts. A typical FE-SEM-EDX analysis for the 5 wt% Ru/Al2O3 catalyst is shown in Figure 1. It can be seen that Ru was well dispersed on alumina from FE-SEM micrograph because the particle brightness appeared to be quite homogeneous in the micrograph. The EDX analysis showed the presence of Ru, Al, and O. The Ru amount measured by EDX was close to the designed Ru loading of 5 wt% of Al2O3.
The catalyst crystal structure was characterized by X-ray diffractometer (XRD, Rigaku, Japan) using a Cu Kα radiation (λ = 1.5406 Å) at 40 kV and 40 mA. The 2θ angle was scanned at a rate of 1.5°/min within the 20° < 2θ < 90° range. XRD patterns of the calcined and reduced catalysts are shown in Figure 2. In Figure 2a for the calcined 10 wt% Ni/Al2O3, the peaks at 37.6°, 43.2°, and 63.0° were attributed to the (111), (200), and (220) planes of NiO species (JCPDS65-6920) [37], respectively. These peaks disappeared after reduction and the Ni crystalline corresponding to the diffraction peaks appeared at 44.5°, 51.7°, and 76.4°. As pointed out by Penkova et al. [38], the Ni loading is critical to the formation of nickel aluminate spinels (NiAl2O4). For the slow Ni loading cases in this study, the reduction conditions were adequate to reduce all the Ni in its metallic form [21,39]. Figure 2b shows the XRD patterns for the calcined and reduced 5 wt% Ru/Al2O3 catalyst. In the calcined Ru/Al2O3, RuO2 peaks appeared at 28.1°, 35.1°, and 54.4°. After reduction, these peaks disappeared and the Ru peaks appeared at 42.2°, 44°, 58.3°, 69.4°, and 78.2°. Figure 2c shows the XRD patterns of the calcined and reduced 1 wt% Ru-15 wt% Ni/Al2O3 catalysts. It can be seen that peaks corresponding to Ni, Ru, NiO, and RuO2 were observed in the reduced and calcined Ru-Ni/Al2O3 catalyst. The mean Ni or Ru crystallite size could be evaluated from XRD patterns shown in Figure 2a,b for 10 wt% Ni/Al2O3, and 5 wt% Ru/Al2O3, respectively. From the Scherrer equation, the estimated mean crystallite sizes of Ni and Ru were 22 nm and 10 nm, respectively. Note that bigger crystallite sizes resulting from high metal loading have been substantiated in the literature [40,41].

2.4. Ni Loading Effect

The time on stream measurement of CO2 methanation performance from 250 to 550 °C under various Ni loadings characterized by CO2 conversion, H2 efficiency, CH4 yield, and CO2 yield is shown in Figure 3. Because the product gas was collected and analyzed every 30 min for each temperature within 2 h, there were four data for each temperature, as shown in Figure 3. From the result shown in Figure 3, a stable reaction could be obtained within 2 h for each temperature. Figure 3a shows the result for CO2 conversion based on Equation (4). CO2 conversion can be enhanced by increasing the Ni loading for low-temperature regimes (250 and 300 °C). However, higher CO2 conversion was obtained for 15 wt% Ni/Al2O3 instead of 20 wt% /Al2O3 for temperatures in the 350 to 450 °C range. For temperatures higher than 500 °C, no noticeable difference was obtained in CO2 conversion for both 15 and 20 wt% catalysts. However, the Ni loading effect on CO2 conversion could still be found for temperatures higher than 500 °C. Based on Figure 3a, the Ni loading should be greater than 10 wt% to have effective CO2 conversion.
Figure 3b shows the H2 efficiency. In the stoichiometric CO2 methanation reaction with a molar ratio of H2/CO2 = 4, only 1/2 of the H2 is converted into CH4 and the other 1/2 is converted into water. Additionally, 2 moles of H2 are required to produce one mole of CH4. Therefore, the maximum theoretical H2 efficiency that could be achieved is only 25%. With a molar ratio of H2/CO = 5 used in this study, the maximum H2 efficiency would be 20%. Due to low catalytic activity, almost zero or low H2 efficiency results from low temperature (250 and 300 °C) and low Ni loadings. Similar to CO2 conversion, H2 efficiency increases with the temperature, indicating that more H2 is consumed in the reaction. For temperatures in the 350 to 450 °C range, higher H2 efficiency was obtained from 15 wt% Ni/Al2O3 instead of 20 wt% Ni/Al2O3. The maximum H2 efficiency was observed at 400 °C. For temperatures higher than 450 °C, H2 efficiency decreased with increased temperature. In this high-temperature regime, higher H2 efficiency was obtained from 15 wt% Ni/Al2O3.
In Figure 3c, the CH4 yield as a function of the reaction temperature and time on stream is shown. Similar to CO2 conversion, the CH4 yield increased with increased temperature and Ni loading. Because of low catalytic activity, low CH4 yield occurred when the temperature was low. The maximum CH4 yield was obtained at 400 °C. With temperatures higher than 400 °C, the CH4 yield decreased with increased temperature.
As shown in Figure 3d, CO formed in the reaction at high temperatures. As pointed out above, the CO2 methanation reaction is composed of two successive reactions: RWGS and CO methanation. Therefore, CO is one of the intermediates in CO2 methanation produced from RWGS. Because the RWGS reaction is endothermic and favored thermodynamically and kinetically at high temperatures, CO formation was found in the high-temperature regime, as shown in Figure 3d. Another reaction that contributes to CO formation is CO2 dissociation:
C O 2  dissociation:  CO 2 ads CO ads + O ads
As mentioned above [9,10], CO2 dissociation into CO is considered a rate-determining step of CO2 methanation.
To have a better understanding of the Ni loading and temperature effects on CO2 methanation, the averaged values of CO2 conversion, H2 efficiency, CH4 yield, and CO yield based on the results shown in Figure 3 are plotted in Figure 4. The equilibrium results calculated by taking CO2 methanation and RWGS reactions into account under the same experimental operating conditions are also shown in Figure 4. As shown in Figure 4a, it is indicated that CO2 conversion increased with increasing Ni loading. For the 20 wt% Ni/Al2O3 case, CO2 conversion was lower than that obtained from 15 wt% Ni/Al2O3 for temperatures higher than 350 °C. The reason for decreased CO2 conversion with high Ni loading may be due to decreased nickel dispersion resulting in bigger crystallite size [40,41]. Moreover, Figure 4a shows that the CO2 conversion increases with temperature reached a maximum value at 400 °C, and then decreased with temperature. Since the CO2 methanation is an exothermic reaction, the CO2 conversion decrease in high temperature was due to the reverse CO2 methanation reaction in which CO2 was reproduced. As shown in Figure 4a, the maximum CO2 conversion, i.e., 85%, was obtained using the 15 wt% Ni/Al2O3 catalyst at 400 °C. As the temperature exceeded 400 °C, the catalytic CO2 conversion obtained from 15 wt% Ni/Al2O3 and 20 wt% Ni/Al2O3 was about the same as the thermodynamic equilibrium result.
As also shown in Figure 4a, the catalytic CO2 conversion over Ni catalysts was lower than the equilibrium CO2 conversion, which gradually decreased with the increase in temperature due to the thermodynamic nature of CO2 methanation. From the stoichiometric reaction, the change in Gibbs free energy for CO2 methanation is:
Δ G 298 K = ( 173 + 0.1983 T )   kJ / mole
This is favorable for CH4 production from the chemical equilibrium viewpoint for Gibbs free energy ΔG being less than zero. In such cases, the equilibrium constant is positive. However, with the increase in temperature, ΔG increases, and the reaction shifts towards the reactants. This accounts for the decrease in equilibrium CO2 conversion at high temperature and is also called the reverse CO2 methanation reaction. The reason for the poor catalytic activity at low temperature is that CO2 methanation is an eight-electron process for fully reducing the CO2 (+4) into CH4 (−4), which usually has a great kinetic barrier to activate the stable CO2 molecule, and requires an efficient catalyst to decrease the activation energy [42].
The H2 efficiency corresponding to Figure 4a is shown in Figure 4b. Similar to CO2, Figure 4b shows that the H2 efficiency increases with the increase in Ni loading. H2 efficiency increased with temperature, reached a maximum value at 400 °C, and then decreased with temperature. The maximum H2 efficiency, i.e., 17%, was achieved with a 15 wt% Ni/Al2O3 catalyst.
Figure 4c shows the averaged CH4 yield based on the results shown in Figure 3c. Based on the CO2 conversion and H2 efficiency results, the CH4 yield can be enhanced by increasing the temperature and Ni loading. The maximum CH4 yield was obtained at 400 °C with a value of 80% from 15 wt% Ni/Al2O3 catalyst. At temperatures higher than 400 °C, CH4 yield decreased due to the reverse CO2 methanation mentioned above. Additionally, CH4 yield may also decrease due to methane-steam reforming (MSR) and CH4 decomposition reactions:
MSR:  CH 4 + H 2 O CO + 3 H 2 Δ H 298 K = 206   kJ / mole
C H 4  decomposition:  CH 4 2 H 2 + C Δ H 298 K = 75.6   kJ / mole
Both reactions are endothermic and favor high temperatures. Also, CO was formed in the RWGS reaction causing less carbon to be formed into CH4. Because of these reactions, it can be seen that catalytic CH4 yield departs from the thermodynamic equilibrium result to a higher extent compared with the CO2 conversion shown in Figure 4a. Note also that H2 can be produced from MSR and CH4 decomposition, while it is consumed in the RWGS reaction. The amount of net H2 production from these reactions may affect the H2 efficiency described in Figure 4b.
Figure 4d shows the CO yield as a function of the temperature and Ni loading. From Equation (10), CO may also be produced from MSR in addition to RWGS at high temperatures. Another possible reaction for CO formation is due to the Boudouard reaction:
CO 2 + C 2 CO Δ H 298 K = 172.4   kJ / mole
In this reaction, CO2 is reacted with C to form CO. Since this reaction is endothermic, it is favorable at high temperatures. Based on Figure 4d, less CO is produced when the Ni loading is high. When the Ni loading is low (5 wt% Ni/Al2O3 case), CO yield is high due to poor activity for converting CO into CH4. Moreover, the temperature at which CO forms can be as low as 300 °C for 5 wt% Ni/Al2O3 catalyst. Note that 15 wt% Ni/Al2O3 produces the lowest amount of CO among the catalysts studied.
In Equation (11), carbon is formed from CH4 decomposition and consumed in the Boudouard reaction, as shown in Equation (12). The net carbon production may be deposited onto the catalyst surface, leading to catalyst activity loss. Carbon balance can be performed based on the carbon supply in the reactant and carbon production in the product [43]. For CO2 methanation, the carbon supply is the CO2, while the carbon productions are in the forms of CO and CH4. Using the data shown in Figure 4 as an example, for T = 400 °C and 15% wt% Ni/Al2O3 catalyst, CO2 conversion was 88%, CH4 yield was 80%, and CO yield was 1%. This corresponds to about 7% of carbon deposition onto the catalyst and reactor surfaces. Note that CO yield was primarily due to RWGS which was not active for low temperature and produced low CO yield. Also, there was no carbon formation found from thermodynamic equilibrium analysis for the reactant composition used in this study, which agreed with the study by Janke et al. [44]. The 7% difference between the carbon supply and carbon production may have been due to experimental uncertainty.

2.5. Ru Loading Effect

The CO2 methanation performance using the Ru catalyst is shown in Figure 5. Similar to the Ni case, the CO2 conversion, H2 efficiency, CH4 yield, and CO yield values shown in Figure 5 are based on the averaged values obtained from the time on stream measurement which is provided in Figure S1 in the supplementary information. From Figure 5a, the Ru activity increased with temperature and Ru loading. Compared with the Ni catalyst results shown in Figure 4a, Ru is more active at low-temperature regimes. As shown in Figure 5a, the CO2 conversion was 38% with the 3 wt% Ru/Al2O3 catalyst, but only 10% CO2 conversion was obtained with the 20 wt% Ni/Al2O3 catalyst shown in Figure 4a. Similar to the Ni case, the CO2 conversion increased with temperature, reaching a maximum value at 400 °C, and then decreased as the temperature increased further. A maximum CO2 conversion of 90% could be obtained from 3 wt% Ru/Al2O3 catalyst, which was also 5% higher than that obtained from 15 wt% Ni/Al2O3, as shown in Figure 4a. For temperatures higher than 400 °C, CO2 conversion decreased dramatically with increased temperature. Compared with the result for Ni shown in Figure 4a, CO2 conversion did not follow the variation trend as it did in the thermodynamic equilibrium case. This was because Ru is more active for RWGS reactions in the higher temperature regime, as will be discussed later.
Figure 5b shows the H2 efficiency corresponding to the results shown in Figure 5a. Maximum H2 efficiency of 18% was obtained from the 5 wt% Ru/Al2O3 catalyst at 400 °C. It is interesting to note that low H2 efficiency was found for temperatures greater than 400 °C. At a temperature of 550 °C, very low H2 efficiency was found for the 1 wt% Ru and 3 wt% Ru cases. That is, H2 did not efficiently convert into CH4 in this high-temperature regime.
Figure 5c shows that the CH4 yield increased with increased Ru loading and temperature. At a temperature of 400 °C, a maximum CH4 yield of 80% could be obtained in the 5 wt% Ru/Al2O3 case. With temperatures higher than 400 °C, the CH4 yield decreased as the temperature increased. Note that low CH4 yields resulted in temperatures of 500 and 550 °C. As mentioned above, the amount of H2 converted into CH4 was low. With a temperature of 550 °C, Sharma et al. [45] reported that low CH4 yield resulted from using 5 wt% Ru/CeO2 as the catalyst in their CO2 methanation experiment. Similar to the Ni case, the CH4 yield may also be reduced by reactions such as MSR and CH4 decomposition.
Figure 5d shows the CO yield as functions of the Ru loading and temperature. For temperatures less than 400 °C, almost zero CO yield was obtained from all the Ru loadings studied. As the temperature increased beyond 400 °C the CO yield increased with increased temperature. In the high-temperature regime, CO2 conversion was still high, although the CH4 yield was low. From the measured CO results shown in Figure 5d, high CO2 conversion occurred partially due to CO formation. In the 1 wt% Ru/Al2O3 case, the CO yield was as high as 40%. This indicated that Ru is highly active for endothermic RWGS for CO production but not for the subsequent CO methanation for methane production. It was also expected that the H2 efficiency would be low since less H2 is consumed in RWGS for producing water.
Based on the results shown in Figure 5, it can be concluded that Ru has higher activity at low temperatures compared with Ni for CO2 methanation. However, Ru is not a suitable catalyst for CO2 methanation at temperatures higher than 400 °C due to low CH4 yield. In addition to the reaction temperature, Ru dispersion range on support is one of the factors that control the product from CO2 methanation, as pointed out in the study by Kwak et al. [34]. With a low Ru loading, atomically dispersed Ru on Al2O3 produced CO exclusively in CO2 methanation. With a higher Ru loading, 3D Ru clusters were formed that were highly active for CO2 methanation. It was also suspected that bigger crystallite size resulting from increased temperature was one of the factors that reduced the CH4 yield from the Ru catalyst.

2.6. Ru-Ni Loading Effect

CO2 methanation performance using Ru-Ni catalysts is shown in Figure 6 and Figure 7. Similar to the Ni and Ru cases discussed above, CO2 conversion, H2 efficiency, CH4 yield, and CO yield are based on the averaged time-on-stream measured data, which are provided in Figure S2 in the supplementary information. The CO2 methanation performance using Ni, Ru, and Ru-Ni catalysts are plotted together. In Figure 6, CO2 methanation performance using 1 wt% Ru-10 wt% Ni/Al2O3 catalyst is compared with those using 10 wt% Ni/Al2O3 and 1 wt% Ru/Al2O3 catalysts. From Figure 6a for a temperature of 250 °C, the CO2 conversion increased from 7% with 10 wt% Ni/Al2O3 to 15% when 1 wt% Ru-10 wt% Ni/Al2O3 catalyst was used. As pointed out in studies by Zhen et al. [31] and Skriver and Rosengaard [46], due to the difference in surface free energy, the Ru segregation tendency on the Ni catalyst surface can be identified. This tendency leads to higher bimetallic Ru-Ni catalyst activity compared with monometallic Ni or Ru catalyst, as clearly shown in Figure 6a. The maximum CO2 conversion was obtained at 400 °C, with a value of 88%, as shown in Figure 6a. In the high-temperature regime, CO2 conversions obtained from Ru-Ni and Ni catalysts followed the same variation trend as that of the equilibrium result. It was also seen that 1 wt% Ru-10 wt% Ni/Al2O3 catalysts could improve the CO2 methanation activity as compared with the monoatomic 1 wt% Ru catalyst in the high-temperature regime.
Figure 6b shows the H2 efficiency corresponding to the results shown in Figure 6a. With the increased CO2 conversion using the Ru-Ni catalyst, the H2 efficiency was also enhanced using the Ru-Ni catalyst, as shown in Figure 6b. Similarly, the CH4 yield was also improved, as shown in Figure 6c. Significant improvement could be obtained compared with the 1 wt% Ru catalyst case for the high-temperature regime. Using the Ru-Ni catalyst, the CO yield could be kept at the same level as that of Ni catalyst, as shown in Figure 6d.
In Figure 7, CO2 methanation performance using 1 wt% Ru-15 wt% Ni/Al2O3 catalyst is compared with those using 15 wt% Ni/Al2O3 and 1 wt% Ru/Al2O3 catalysts. As discussed in Figure 4 CO2 methanation performance could be enhanced by increasing the Ni loading; also, the 15 wt% Ni/Al2O3 catalyst produced maximum CO2 conversion at 400 °C. With the addition of 1 wt% Ru, Figure 7 clearly shows that further enhancement of CO2 methanation performance was obtained. As shown in Figure 7a, the CO2 conversion improved from 14% using 15 wt% Ni/Al2O3 to 40% using 1 wt% Ru-15 wt% Ni/Al2O3 at a temperature of 250 °C. CO2 conversion with a value of 80% could be obtained for temperatures in the 300 to 400 °C range. The corresponding H2 efficiency is shown in Figure 7b indicating that higher H2 efficiency could be obtained in the low-temperature regime. In Figure 7c, the CH4 yield could be improved from 5% with 15 wt% Ni/Al2O3 catalyst to 38% at 250 °C when a 1 wt% Ru-15 wt% Ni/Al2O3 catalyst was used. With higher CH4 yield, CO formation could be kept at the same level as the thermodynamic equilibrium result, as shown in Figure 7d.

2.7. Stability Test

The thermal stability of 1 wt% Ru-10 wt% Ni/Al2O3 catalyst was tested using three continuous ascending-descending temperature cycles. That is, the temperature was raised from 250 °C to 550 °C at 50 °C increments and then reduced from 550 °C to 250 °C also at 50 °C increments. The test procedure was the same as that described in Figure 3. For each temperature, the time-on-stream measurement lasted two hours. The detailed time-on-stream measurement is shown in Figure S3 in the supplementary information. To complete three cycles, the catalyst was under testing for a total of 74 h. The test result is shown in Figure 8. Note that the 1 wt% Ru–10 wt% Ni/Al2O3 catalyst used for the result shown in Figure 6 was reused for the stability test. As shown in Figure 8, it showed that good catalyst thermal stability could be obtained. After three cycles, the maximum CO2 conversion changed from 82% to 78%, as shown in Figure 8a; the maximum H2 efficiency changed from 15% to 14%, as shown in Figure 8b; the maximum CH4 yield changed from 78% to 73%, as shown in Figure 8c; and the maximum CO yield changed from 18% to 16%, as shown in Figure 8d. The reason for the decrease in CO2 methanation performance after a long period of operation may be due to the gradual re-oxidization of Ni to NiO and oxidization Ru to RuO2 by CO2, or to the water produced. Because Ru has better resistance to carbon deposition than Ni [47], the Ru contained in the Ru-Ni catalyst played an important role in enhancing the activity and stability of the catalysts and achieved high CO2 conversion and CH4 yield [48,49].

3. Experimental

3.1. Catalyst Preparation

Ni/Al2O3, Ru/Al2O3, and Ru-Ni/Al2O3 with various metal loadings were prepared in this study using the wet incipient impregnation method. Spherical Al2O3 particles with an average diameter of 0.5~1.2 mm and a specific surface area of 220 m2/g were used as the catalyst support (Aldrich, USA). Aqueous Ni(NO3)2·6H2O and Ru(NO)(NO3)3 (Merck, USA) were used as the Ni and Ru precursors, respectively. The bimetallic catalyst Ru-Ni/Al2O3 was prepared by adding Ru to Ni/Al2O3 using sequential impregnation [50]. All of the prepared catalysts in this study are listed in Table 1. The chemical composition was verified using XRF (X-MET5100, Oxford Instruments, UK) and the accuracy was shown to be within 5%. Figure 9 shows the prepared Al2O3-supported Ni, Ru, and Ru-Ni catalysts. After impregnation, the catalysts were dried in an oven at a temperature of 110 °C for 24 h. The catalysts were then calcined in the air flow at 700 °C for 2 h.

3.2. Experimental Setup

A schematic diagram of the experimental setup is shown in Figure 10a. The CO2 and N2 were supplied from gas bottles while H2 was supplied from a hydrogen generator (H2PEM-165, USA). The flow rate of each gas species was controlled using a mass flow controller MFC 5850 E (Brooks, USA). The gases were mixed in a mixer before entering the reactor. The CO2 methanation was carried out using a conventional catalytic packed-bed tubular reactor operated isothermally in a temperature-controllable oven. A quartz tube with an inner diameter of 4 mm was used as the reactor in which the catalyst was loaded in the central section with two ends fixed using quartz wool, as shown in Figure 10b. The catalyst was activated using in situ reduction at 700 °C with 5% H2 in N2 before each experimental run [21,39]. After the reduction, the temperature was lowered to 250 °C, which is the initial activity test temperature. The activity test was performed from 250 to 550 °C. A thermocouple inserted into the catalyst bed was used to monitor the reaction temperature. After removing the product water with a condenser, the product gas was collected and analyzed using gas chromatography (Agilent 6890, USA).

4. Conclusions

CO2 methanation was studied experimentally using Ni-, Ru-, and Ru-Ni-supported by Al2O3 as the catalysts in a fixed-bed reactor at atmospheric pressure. The H2/CO2 molar ratio and space velocity were fixed at 5 and 5835 h−1, respectively. Based on the measured results, the following conclusions can be made:
(1)
For the three types of catalyst studied, the optimum reaction temperature was found to be 400 °C. At this temperature, maximum CO2 conversion, maximum H2 efficiency, maximum CH4 yield, and minimum CO yield were obtained for all metal loadings studied.
(2)
CO2 methanation performance at low temperatures could be enhanced by increasing the catalyst loading for all the catalysts studied. Compared with Ni, Ru was more active in the low-temperature regime. At higher temperature regimes, CO2 methanation performance followed a variation trend resulting from the thermodynamic equilibrium of the Ni catalyst. The measured results indicated that Ru was active for the reverse water-gas shift reaction in the high-temperature regime. This led to a low CH4 yield and high CO yield.
(3)
The measured data indicated that the bimetallic Ru-Ni catalyst was more active compared with the monometallic Ni or Ru catalysts for CO2 methanation in a low-temperature regime.
(4)
The Ru-Ni catalyst showed good thermal stability, i.e., about 3% and 5% decreases in CO2 conversion and CH4 yield resulted after 74 h testing, respectively.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/10/10/1112/s1, Figure S1: Time on stream measured CO2 methanation performance as functions of the Ru loading and reaction temperature. (a) CO2 conversion, (b) H2 efficiency, (c) CH4 selectivity, (d) CO selectivity, (e) CH4 yield, and (f) CO yield. Figure S2: Time on stream measured CO2 methanation performance as functions of the Ru-Ni catalysts and reaction temperature. (a) CO2 conversion, (b) H2 efficiency, (c) CH4 selectivity, (d) CO selectivity, (e) CH4 yield, and (f) CO yield. Figure S3: Time on stream measured thermal stability test of 1wt% Ru-10wt% Ni/Al2O3 catalyst under three continuous ascending-descending temperature change cycles between 250 and 550 °C. (a) CO2 conversion, (b) H2 efficiency, (c) CH4 selectivity, (d) CO selectivity, (e) CH4 yield, and (f) CO yield.

Author Contributions

Conceptualization, R.-Y.C.; methodology, R.-Y.C.; formal analysis, C.-C.W.; investigation, R.-Y.C. and C.-C.W.; data curation, C.-C.W.; writing—original draft preparation, R.-Y.C.; writing—review and editing, R.-Y.C.; visualization, R.-Y.C.; supervision, R.-Y.C.; project administration, R.-Y.C.; funding acquisition, R.-Y.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Science and Technology, Taiwan, under the grant number MOST106-2221-E-005-073-MY3, and the APC was funded by National Chung Hsing University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Centi, G.; Perathoner, S. CO2-based energy vectors for the storage of solar energy. Greenh. Gases Sci. Technol. 2011, 1, 21–35. [Google Scholar] [CrossRef]
  2. Luo, X.; Wang, J.; Dooner, M.; Clarke, J. Overview of current development in electrical energy storage technologies and the application potential in power system operation. Appl. Energy 2015, 137, 511–536. [Google Scholar] [CrossRef] [Green Version]
  3. Suberua, M.Y.; Mustafa, M.Y.; Bashir, N. Energy storage systems for renewable energy power sector integration and mitigation of intermittency. Renew. Sustain. Energy Rev. 2014, 35, 499–514. [Google Scholar] [CrossRef]
  4. Ghaib, K.; Ben-Fares, F. Power-to-Methane: A state-of-the-art review. Renew. Sustain. Energy Rev. 2018, 81, 433–446. [Google Scholar] [CrossRef]
  5. Schiebahn, S.; Grube, T.; Robinius, M.; Tietze, V.; Kumar, B.; Stolten, D. Power to gas: Technological overview, systems analysis and economic assessment for a case study in Germany. Int. J. Hydrog. Energy 2015, 40, 4285–4294. [Google Scholar] [CrossRef]
  6. Younas, M.; Kong, L.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]
  7. Eckle, S.; Anfang, H.G.; Behm, R.J. Reaction intermediates and side products in the methanation of CO and CO2 over supported Ru catalysts in H2-rich reformate gases. J. Phys. Chem. C 2011, 115, 1361–1367. [Google Scholar] [CrossRef]
  8. Fisher, I.A.; Bell, A.T. A comparative study of CO and CO2 hydrogenation over Rh/SiO2. J. Catal. 1996, 162, 54–65. [Google Scholar] [CrossRef]
  9. Williams, K.J.; Boffa, A.B.; Salmeron, M.; Bell, A.T.; Somorjai, G.A. The kinetics of CO2 hydrogenation on a Rh foil promoted by titania overlayers. Catal. Lett. 1991, 9, 415–426. [Google Scholar] [CrossRef]
  10. Swalus, C.; Jacquemin, M.; Poleunis, C.; Bertrand, P.; Ruiz, P. CO2 methanation on Rh/γ-Al2O3 catalyst at low temperature: “in situ” supply of hydrogen by Ni/activated carbon catalyst. Appl Catal B Environ. 2012, 125, 41–50. [Google Scholar] [CrossRef]
  11. Shanmugam, V; Neuberg, S; Zapf, R; Pennemann, H; Kolb, G. Effect of support and chelating ligand on the synthesis of Ni catalysts with high activity and stability for CO2 Methanation. Catalysts 2020, 10, 493. [Google Scholar] [CrossRef]
  12. Frontera, P.; Macario, A.M.; Ferraro, M.; Antonucci, P. Supported catalysts for CO2 methanation: A review. Catalysts 2017, 7, 59. [Google Scholar] [CrossRef]
  13. 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]
  14. Pandey, D.; Ray, K.; Bhardwaj, R.; Bojja, S.; Chary, K.V.R.; Deo, G. Promotion of unsupported nickel catalyst using iron for CO2 methanation. Int. J. Hydrog. Energy 2018, 43, 4987–5000. [Google Scholar] [CrossRef]
  15. Wang, X.; Hong, Y.; Shi, H.; Szanyi, J. Kinetic modeling and transient DRIFTS–MS studies of CO2 methanation over Ru/Al2O3 catalysts. J. Catal. 2016, 343, 185–195. [Google Scholar] [CrossRef] [Green Version]
  16. Hemmingsson, F.; Schaefer, A.; Skoglundh, M.; Carlsson, P. CO2 methanation over Rh/CeO2 studied with infrared modulation excitation spectroscopy and phase sensitive detection. Catalysts 2020, 10, 601. [Google Scholar] [CrossRef]
  17. Ronsch, S.; Schneider, J.; Matthischke, S.; Schluter, M.; Gotz, M.; Lefebvre, J.; Prabhakaran, P.; Bajohr, S. Review on methanation-From fundamentals to current projects. Fuel 2016, 166, 276–296. [Google Scholar] [CrossRef]
  18. Martins, J.A.; Faria, A.C.; Soria, M.A.; Miguel, C.V.; Rodrigues, A.E.; Madeira, L.M. CO2 methanation over hydrotalcite-derived nickel/ruthenium and supported ruthenium catalysts. Catalysts 2019, 9, 1008. [Google Scholar] [CrossRef] [Green Version]
  19. Garbarino, G.; Bellotti, D.; Finocchio, E.; Magistri, L.; Busca, G. Methanation of carbon dioxide on Ru/Al2O3: Catalytic activity and infrared study. Catal. Today 2016, 277, 21–28. [Google Scholar] [CrossRef]
  20. Duyar, M.S.; Ramachandran, A.; Wang, C.; Farrauto, R.J. Kinetics of CO2 methanation over Ru/γ-Al2O3 and implications for renewable energy storage applications. J. CO2 Util. 2015, 12, 27–33. [Google Scholar] [CrossRef]
  21. Rahmani, S.; Rezaei, M.; Meshkani, F. Preparation of highly active nickel catalysts supported on mesoporous nano crystalline γ-Al2O3 for CO2 methanation. J. Ind. Eng. Chem. 2014, 20, 1346–1352. [Google Scholar] [CrossRef]
  22. Lee, C.J.; Lee, D.H.; Kim, T. Enhancement of methanation of carbon dioxide using dielectric barrier discharge on a ruthenium catalyst at atmospheric conditions. Catal. Today 2017, 293–294, 97–104. [Google Scholar] [CrossRef]
  23. Xu, L.; Wang, F.; Chen, M.; Nie, D.; Lian, X.; Lu, Z.; Chen, H.; Zhang, K.; Ge, P. CO2 methanation over rare earth doped Ni based mesoporous catalysts with intensified low-temperature activity. Int. J. Hydrog. Energy 2017, 42, 15523–15539. [Google Scholar] [CrossRef]
  24. Muroyama, H.; Tsuda, Y.; Asakoshi, T.; Masitah, H.; Okanishi, T.; Matsui, T.; Eguchi, K. Carbon dioxide methanation over Ni catalysts supported on various metal oxides. J. Catal. 2016, 343, 178–184. [Google Scholar] [CrossRef] [Green Version]
  25. Abate, S.; Mebrahtu, C.; Giglio, E.; Deorsola, F.; Bensaid, S.; Perathoner, S.; Pirone, R.; Centi, G. Catalytic performance of γ-Al2O3-ZrO2-TiO2-CeO2 composite oxide supported Ni-based catalysts for CO2 Methanation. Ind. Eng. Chem. Res. 2016, 55, 4451–4460. [Google Scholar] [CrossRef]
  26. Nawfal, M.; Gennequin, C.; Labaki, M.; Nsouli, B.; Abi-Aada, E. Hydrogen production by methane steam reforming over Ru supported on Ni–Mg–Al mixed oxides prepared via hydrotalcite route. Int. J. Hydrog. Energy 2015, 40, 1269–1277. [Google Scholar] [CrossRef]
  27. Luisetto, I.; Sarno, C.; De Felicis, D.; Basoli, F.; Battocchio, C.; Tuti, S.; Licoccia, S.; Di Bartolomeo, E. Ni supported on γ-Al2O3 promoted by Ru for the dry reforming of methane in packed and monolithic reactors. Fuel Process. Technol. 2017, 158, 130–140. [Google Scholar] [CrossRef]
  28. Tada, S.; Minori, D.; Otsuka, F.; Kikuchi, R.; Osada, K.; Akiyama, K.; Satokawa, S. Effect of Ru and Ni ratio on selective CO methanation over Ru–Ni/TiO2. Fuel 2014, 129, 219–224. [Google Scholar] [CrossRef]
  29. Gao, Z.; Cui, L.; Ma, H. Selective methanation of CO over Ni/Al2O3 catalyst: Effects of preparation method and Ru addition. Int. J. Hydrog. Energy 2016, 41, 5484–5493. [Google Scholar] [CrossRef]
  30. Polanski, J.; Siudyga, T.; Bartczaka, P.; Kapkowskia, M.; Ambrozkiewicz, W.; Nobis, A.; Sitko, R.; Klimontko, J.; Szade, J.; Lelatko, J. Oxide passivated Ni-supported Ru nanoparticles in silica: A new catalyst for low-temperature carbon dioxide methanation. Appl. Catal. B Environ. 2017, 206, 16–23. [Google Scholar] [CrossRef]
  31. Zhen, W.; Li, B.; Lu, G.; Ma, J. Enhancing catalytic activity and stability for CO2 methanation on Ni–Ru/γ-Al2O3 via modulating impregnation sequence and controlling surface active species. RSC Adv. 2014, 4, 16472–16479. [Google Scholar] [CrossRef]
  32. Navarro, J.C.; Centeno, M.A.; Laguna, O.H.; Odriozola, J.A. Ru-Ni/MgAl2O4 structured catalyst for CO2 methanation. Renew. Energy 2020, 161, 120–132. [Google Scholar] [CrossRef]
  33. Froment, G.F.; Bischoff, K.B. Chemical Reactor Analysis and Design; Wiley: New York, NY, USA, 1990. [Google Scholar]
  34. Kwak, J.H.; Kovarik, L.; Szanyi, J. CO2 Reduction on supported Ru/Al2O3 catalysts: Cluster size dependence of product selectivity. ACS Catal. 2013, 3, 2449–2455. [Google Scholar] [CrossRef]
  35. Kim, M.J.; Youn, J.; Kim, H.J.; Seo, M.W.; Lee, D.; Go, K.S.; Lee, K.B.; Jeon, S.G. Effect of surface properties controlled by Ce addition on CO2 methanation over Ni/Ce/Al2O3 catalyst. Int. J. Hydrog. Energy 2020, 45, 24595–24603. [Google Scholar] [CrossRef]
  36. 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]
  37. Zhu, H.; Razzaq, R.; Li, C.; Muhmmad, Y.; Zhang, S. Catalytic methanation of carbon dioxide by active oxygen material CexZr1−xO2 supported Ni-Co bimetallic nanocatalysts. AICHE J. 2013, 59, 2567–2576. [Google Scholar] [CrossRef]
  38. Penkova, A.; Bobadilla, L.; Ivanova, S.; Domínguez, M.I.; Romero-Sarria, F.; Roger, A.C.; Centeno, M.A.; Odriozola, J.A. Hydrogen production by methanol steam reforming on NiSn/MgO–Al2O3 catalysts: The role of MgO addition. Appl. Catal. A Gen. 2011, 392, 184–191. [Google Scholar] [CrossRef]
  39. Garbarino, G.; Riani, P.; Magistri, L.; Busca, G. A study of the methanation of carbon dioxide on Ni/Al2O3 catalysts at atmospheric pressure. Int. J. Hydrog. Energy 2014, 39, 11557–11565. [Google Scholar] [CrossRef]
  40. 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, 60, 661–668. [Google Scholar] [CrossRef]
  41. Marconi, E.; Tuti, S.; Luisetto, I. Structure-sensitivity of CO2 Methanation over nanostructured Ni Supported on CeO2 Nanorods. Catalysts 2019, 9, 375. [Google Scholar] [CrossRef] [Green Version]
  42. 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]
  43. Zhang, J.; Ren, M.; Li, X.; Hao, Q.; Chen, H.; Ma, X. Ni-based catalysts prepared for CO2 reforming and decomposition of methane. Energy Convers. Manag. 2020, 205, 112419. [Google Scholar] [CrossRef]
  44. Janke, C.; Duyar, M.S.; Hoskins, M.; Farrauto, R. Catalytic and adsorption studies for the hydrogenation of CO2 to methane. Appl. Catal. B Environ. 2014, 152–153, 184–191. [Google Scholar] [CrossRef]
  45. Sharma, S.; Hu, Z.; Zhang, P.; McFarland, E.W.; Metiu, H. CO2 methanation on Ru-doped ceria. J. Catal. 2011, 278, 297–309. [Google Scholar] [CrossRef]
  46. Skriver, H.L.; Rosengaard, N.M. Surface energy and work function of elemental metals. Phys. Rev. B Condens. Matter Mater. Phys. 1992, 46, 7157. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Rostrup-Nielsen, J.R.; Hansen, J.H.B. CO2-reforming of methane over transition metals. J. Catal. 1993, 144, 38–49. [Google Scholar] [CrossRef]
  48. Jalama, K. Carbon dioxide hydrogenation over nickel-, ruthenium-, and copper-based catalysts: Review of kinetics and mechanism. Catal. Rev. 2017, 59, 95–164. [Google Scholar] [CrossRef]
  49. Nguyen, T.T.M.; Wissing, L.; Skjoth-Rasmussen, M.S. High temperature methanation: Catalyst considerations. Catal. Today 2013, 215, 233–238. [Google Scholar] [CrossRef]
  50. Yeung, C.; Tsang, S.C. Some optimization in preparing core-shell Pt–ceria catalysts for water gas shift reaction. J. Mol. Catal. A Chem. 2010, 322, 17–25. [Google Scholar] [CrossRef]
Figure 1. FE-SEMEDX micrographs of fresh 5 wt% Ru/Al2O3 showing catalyst morphology and EDX analysis recorded on an area of the picture.
Figure 1. FE-SEMEDX micrographs of fresh 5 wt% Ru/Al2O3 showing catalyst morphology and EDX analysis recorded on an area of the picture.
Catalysts 10 01112 g001
Figure 2. XRD patterns of (a) 10 wt% Ni/Al2O3, (b) 5 wt% Ru/Al2O3, and (c) 1 wt% Ru-10 wt% Ni/Al2O3 catalysts.
Figure 2. XRD patterns of (a) 10 wt% Ni/Al2O3, (b) 5 wt% Ru/Al2O3, and (c) 1 wt% Ru-10 wt% Ni/Al2O3 catalysts.
Catalysts 10 01112 g002
Figure 3. Time on stream measured CO2 methanation performance as functions of the Ni loading and reaction temperature. (a) CO2 conversion, (b) H2 efficiency, (c) CH4 yield, and (d) CO yield.
Figure 3. Time on stream measured CO2 methanation performance as functions of the Ni loading and reaction temperature. (a) CO2 conversion, (b) H2 efficiency, (c) CH4 yield, and (d) CO yield.
Catalysts 10 01112 g003
Figure 4. Averaged CO2 methanation performance as a function of the temperature using Ni catalyst based on Figure 3. (a) CO2 conversion, (b) H2 efficiency, (c) CH4 yield, and (d) CO yield.
Figure 4. Averaged CO2 methanation performance as a function of the temperature using Ni catalyst based on Figure 3. (a) CO2 conversion, (b) H2 efficiency, (c) CH4 yield, and (d) CO yield.
Catalysts 10 01112 g004
Figure 5. Averaged CO2 methanation performance as a function of the temperature using Ru catalyst based on time on stream measurement. (a) CO2 conversion, (b) H2 efficiency, (c) CH4 yield, and (d) CO yield.
Figure 5. Averaged CO2 methanation performance as a function of the temperature using Ru catalyst based on time on stream measurement. (a) CO2 conversion, (b) H2 efficiency, (c) CH4 yield, and (d) CO yield.
Catalysts 10 01112 g005
Figure 6. Averaged CO2 methanation performance as a function of the temperature using 1 wt% Ru-10 wt% Ni/Al2O3 catalyst based on time on stream measurement. (a) CO2 conversion, (b) H2 efficiency, (c) CH4 yield, and (d) CO yield.
Figure 6. Averaged CO2 methanation performance as a function of the temperature using 1 wt% Ru-10 wt% Ni/Al2O3 catalyst based on time on stream measurement. (a) CO2 conversion, (b) H2 efficiency, (c) CH4 yield, and (d) CO yield.
Catalysts 10 01112 g006
Figure 7. Averaged CO2 methanation performance as a function of temperature using 1 wt% Ru-15 wt% Ni/Al2O3 catalyst based on time on stream measurement. (a) CO2 conversion, (b) H2 efficiency, (c) CH4 yield, and (d) CO yield.
Figure 7. Averaged CO2 methanation performance as a function of temperature using 1 wt% Ru-15 wt% Ni/Al2O3 catalyst based on time on stream measurement. (a) CO2 conversion, (b) H2 efficiency, (c) CH4 yield, and (d) CO yield.
Catalysts 10 01112 g007
Figure 8. Thermal stability test of 1 wt% Ru–10 wt% Ni/Al2O3 catalyst under three continuous ascending-descending temperature change cycles between 250 and 550 °C. (a) CO2 conversion, (b) H2 efficiency, (c) CH4 yield, and (d) CO yield.
Figure 8. Thermal stability test of 1 wt% Ru–10 wt% Ni/Al2O3 catalyst under three continuous ascending-descending temperature change cycles between 250 and 550 °C. (a) CO2 conversion, (b) H2 efficiency, (c) CH4 yield, and (d) CO yield.
Catalysts 10 01112 g008
Figure 9. Prepared catalysts used in CO2 methanation.
Figure 9. Prepared catalysts used in CO2 methanation.
Catalysts 10 01112 g009
Figure 10. (a) Schematic of the experimental setup and (b) catalyst bed.
Figure 10. (a) Schematic of the experimental setup and (b) catalyst bed.
Catalysts 10 01112 g010
Table 1. Abbreviation and metal loading amount for Al2O3-supported Ni, Ru, and Ru-Ni catalysts.
Table 1. Abbreviation and metal loading amount for Al2O3-supported Ni, Ru, and Ru-Ni catalysts.
Catalyst TypeCatalyst
Ni catalyst5 wt% Ni/Al2O3
10 wt% Ni/Al2O3
15 wt% Ni/Al2O3
20 wt% Ni/Al2O3
Ru catalyst1 wt% Ru/Al2O3
3 wt% Ru/Al2O3
5 wt% Ru/Al 2O3
Bimetallic catalyst1 wt% Ru-10 wt% Ni/Al2O3
1 wt% Ru-15 wt% Ni/Al2O3
Note: 1. The weight percentage is based on the weight of the support. 2. The chemical composition was verified by XRF (X-MET5100, Oxford Instruments, UK) and the accuracy is within 5%.

Share and Cite

MDPI and ACS Style

Chein, R.-Y.; Wang, C.-C. Experimental Study on CO2 Methanation over Ni/Al2O3, Ru/Al2O3, and Ru-Ni/Al2O3 Catalysts. Catalysts 2020, 10, 1112. https://doi.org/10.3390/catal10101112

AMA Style

Chein R-Y, Wang C-C. Experimental Study on CO2 Methanation over Ni/Al2O3, Ru/Al2O3, and Ru-Ni/Al2O3 Catalysts. Catalysts. 2020; 10(10):1112. https://doi.org/10.3390/catal10101112

Chicago/Turabian Style

Chein, Rei-Yu, and Chih-Chang Wang. 2020. "Experimental Study on CO2 Methanation over Ni/Al2O3, Ru/Al2O3, and Ru-Ni/Al2O3 Catalysts" Catalysts 10, no. 10: 1112. https://doi.org/10.3390/catal10101112

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