Next Article in Journal
Partial Methane Oxidation in Fuel Cell-Type Reactors for Co-Generation of Energy and Chemicals: A Short Review
Next Article in Special Issue
Efficient Catalysts of Ethanol Steam Reforming Based on Perovskite-Fluorite Nanocomposites with Supported Ni: Effect of the Synthesis Methods on the Activity and Stability
Previous Article in Journal
Experimental and Theoretical Investigations of Low-Dimensional BiFeO3 System for Photocatalytic Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of Supports on the Catalytic Activity and Coke Resistance of Ni Catalyst in Dry Reforming of Methane

1
Hydrogen Research Department, Korea Institute of Energy Research, 152 Gajeong-ro, Yuseong-gu, Daejeon 34129, Korea
2
Department of Chemical and Biological Engineering, Korea University, 145 Anam-ro, Seongbuk-gu, Seoul 02841, Korea
3
Advanced Energy and System Engineering, University of Science and Technology (UST), 217 Gajeong-ro, Yuseong-gu, Daejeon 34113, Korea
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(2), 216; https://doi.org/10.3390/catal12020216
Submission received: 9 January 2022 / Revised: 1 February 2022 / Accepted: 10 February 2022 / Published: 14 February 2022
(This article belongs to the Special Issue Advances in Reforming Catalysts for Hydrogen Production)

Abstract

:
The dependence of the catalytic activity and coke resistance of Ni-based catalysts on the support type was investigated in the dry reforming of methane (DRM). Catalysts were prepared using incipient wetness impregnation and analyzed using ICP-OES, BET-BJH, XRD, H2-chemisorption, H2-TPR, and CO2-TPD. DRM was performed at 600–750 °C at 144,000 mL/gcat∙h of GHSV (CH4/CO2/N2 = 1/1/1). Ni/Al2O3 and Ni/MgO catalysts formed NiAl2O4 and NiO-MgO solid solutions, respectively, owing to strong binding between the metal and support. In contrast, MgO-Al2O3 and MgAl2O4 supports suppressed NiAl2O4 and NiO-MgO solid solution formation, due to Mg addition, with high metal dispersions of 4.6 and 6.6%, respectively. In the DRM reaction, the Ni/MgO-Al2O3 and Ni/MgAl2O4 catalysts showed high CH4 conversions of 78.1 and 76.8%, respectively, compared with Ni/Al2O3 and Ni/MgO at 750 °C. A stability test was performed at 600 °C for 20 h. A coke study of the spent catalysts was performed using SEM and TGA. Alkaline-earth metal-containing catalysts Ni/MgO-Al2O3 and Ni/MgAl2O4 with strong CO2 adsorption properties showed 20 wt% reduction in carbon deposition compared to commercial catalysts. Therefore, the support and basic properties of the catalyst significantly influenced the catalyst performance and coke resistance in the DRM.

1. Introduction

Recently, there has been an increased focus on problems related to climate change and the need for reducing carbon dioxide (CO2) emissions. The development of carbon dioxide utilization (CDU) technologies is necessary to reduce CO2 emissions over the long term. Among CDU technologies, the dry reforming of methane (DRM) reaction is an important technique that could be used to reduce the total annual anthropogenic greenhouse gas emissions (multi-GtCO2/year) [1,2]. An advantage of the DRM reaction is its low operational cost, as it does not require complicated gas separation [3]. The DRM reaction (Equation (1)) produces the beneficial syngas (H2 + CO) from CH4 and CO2 [4]. Syngas is a raw material for high value-added chemicals, such as oxo-alcohol, formaldehyde, and methanol. Notably, the product H2/CO ratio of the DRM reaction is 1, which renders the produced syngas suitable for the synthesis of methanol and dimethyl ether or Fischer–Tropsch synthesis [5]. The DRM reaction is strongly endothermic [6] and proceeds at a relatively high temperature (650–850 °C) [7]. The reverse water gas shift (RWGS) reaction (Equation (2)) occurs as a side reaction up to 820 °C under the DRM reaction conditions, due to the endothermic nature of the reaction. Since the RWGS reaction produces CO and H2O using CO2 and H2, the H2/CO ratio is lower than 1.
CH4 + CO2  2H2 + 2CO (ΔH298K = +247 kJ/mol)
CO2 + H2  H2O + CO (ΔH298K = +41 kJ/mol)
CH4  2H2 + C (△H298K = +74.9 kJ/mol)
2CO CO2 + C (△H298K = −172.4 kJ/mol)
Ni-based catalysts are the most widely used in the DRM reaction because of their high activity and low cost [8]. However, they can cause catalyst deactivation due to processes such as carbon deposition and sintering. Moreover, they have been reported to have three phases: Ni metallic state, Ni oxide, and Ni carbide phase, which influence carbon deposition. Among these, Ni metal atoms of the Ni carbide phase bind strongly with carbon; thus, easily causing carbon deposition [9]. In addition, at high reaction temperatures the active metal forms aggregates, owing to the sintering effect. The agglomeration of particles occurs randomly, causing low metal dispersion and low specific surface area at high temperatures. Moreover, the DRM reaction is more vulnerable to carbon deposition because of the higher carbon ratio in the reactants compared with other methane reforming reactions. Carbon deposition mainly occurs through CH4 decomposition (Equation (3)) and CO disproportionation (Equation (4)) reactions. As CH4 decomposition is endothermic, it mainly occurs above 557 °C, whereas the CO disproportionation reaction occurs below 700 °C, due to the exothermic nature of the reaction. Therefore, carbon deposition occurs in the temperature range of 557–700 °C [10]. The morphology of carbon exhibits variations in properties, which cause differences in the activity of the catalyst. In particular, filamentous carbon influences catalyst deactivation during the DRM reaction. Filamentous carbon is formed through diffusion, nucleation, and growth through Ni particles at temperatures > 450 °C. While filamentous carbon does not block the active sites of the Ni metal, it causes catalyst destruction and a pressure drop, due to pore plugging [11]. For this reason, the DRM reaction has limited commercial viability. In this context, the development of catalysts with high catalytic activity and coke resistance in the DRM reaction is essential.
In the preparation of DRM reaction catalysts, the type of support used has a significant influence on the catalytic activity. Among the various supports used in the DRM reaction, Zhang et al. reported that a ZrO2 support had a low metal dispersion, owing to its low specific surface area of 30 m2/g, and also had a weak interaction between NiO and the support [12]. Whereas, amorphous ZrO2 showed a high specific surface area of 200–300 m2/g and had weak thermal stability. However, a tetragonal phase was formed by calcination at temperatures above 450 °C, which remained stable at high temperatures, owing to its strong thermal stability [13]. Depending on the calcination temperature used, Al2O3 supports are divided into α-Al2O3, θ-Al2O3, and γ-Al2O3. In comparison to other Al2O3 supports, the γ-Al2O3 support has high catalytic activity and anti-coking properties, owing to its smaller particle size [14]. In the Ni/γ-Al2O3 catalyst, relatively fine Ni particles are formed, owing to the high specific surface area of γ-Al2O3. However, NiAl2O4 forms a spinel structure during the heat treatment process. The high thermal stability of NiAl2O4 spinel results in an increase in the reduction temperature and a reduction in the number of active sites, owing to the low metal dispersion [12]. The MgO support has a low specific surface area and forms a NiO-MgO solid solution in the presence of the Ni catalyst, which limits reduction and causes low metal dispersion [15]. In contrast, the MgO support enhances the adsorption of acidic CO2 gas, owing to the basic properties of MgO. As the adsorbed CO2 releases oxygen species upon dissociation, the MgO support enhances the coke resistance by promoting the oxidation of carbon formed on the catalyst surface [16,17]. In addition, owing to the basic properties of the Mg-Al mixed oxide support, the Mg-Al mixed oxide support improves the amount of CO2 adsorption, compared with the Al2O3 support [18]. A Mg-Al mixed oxide has a layered double hydroxide (LDH) structure, in which ion exchange between the cation and anion layers is possible. The Mg2+ and Al3+ ions are uniformly dispersed between the hydroxide layers, therefore the LDH structure can effectively disperse the nanoparticles, without particle agglomeration. In addition, the MgO-Al2O3 (LDH) support is transformed into a spinel structure during the heat treatment process and has high thermal stability. In particular, MgO-Al2O3 as a support for a Ni catalyst suppresses the formation of the NiAl2O4 and NiO-MgO solid solution, which is advantageous for high metal dispersion, owing to its large specific surface area [12,17]. Moreover, highly dispersed active metals on the MgO-Al2O3 support have strong metal-support interactions (SMSIs). Consequently, the sintering effect is inhibited [19,20].
Ni-based catalysts have a low coke resistance, but can improve the catalytic activity and durability by dispersing the active metals or enhancing their basic properties [21]. The γ-Al2O3 support has a large specific surface and forms fine particles, while the MgO support has excellent basic properties. The bimetallic MgO-Al2O3 and MgAl2O4 supports can, thus, overcome the disadvantages of the γ-Al2O3 and MgO supports. Therefore, in this study, the influence of various supports: γ-Al2O3, MgO, MgO-Al2O3, and MgAl2O4, was evaluated to improve the catalytic activity and coke resistance of Ni-based catalysts for the DRM reaction. In addition, the prepared catalysts were compared with commercial catalysts (Ni/Al2O3 and Ni/CaAl2O4) in the dry reforming of methane (DRM) reaction.

2. Results and Discussion

2.1. Characteristics

The characteristics of the catalysts are listed in Table 1. The Ni metal content of the catalysts was confirmed using ICP-OES analysis. The loading amounts of the prepared and commercial catalysts were approximately 10 and 15 wt%, respectively. Ni-based catalysts have a reduced specific surface area compared to that of the fresh support, due to the pore-blocking during the process of preparing the supported Ni metals. In particular, the Ni/MgO catalyst exhibited a significant decrease in the specific surface area of 17.7 m2/g after the Ni metal was supported. Zanganeh et al. reported that the crystallite size increased and the specific surface area decreased with increasing calcination temperature when NiO-MgO was calcined at 600–800 °C [22]. The Ni/MgAl2O4 catalyst maintained a relatively high specific surface area of 113.3 m2/g. Since the Ni/MgO-Al2O3 catalyst exhibited changes in the textural properties of the MgO-Al2O3 support during the heat treatment, its specific surface area was lower by 108.7 m2/g, compared to that of the Ni/MgAl2O4 catalyst. This was attributed to the lack of calcination of the MgO-Al2O3 (LDH) support used for the Ni catalyst preparation, which contained MgO and Al2O3 in a 30:70 (wt%) ratio. Among the prepared Ni-based catalysts, the Ni/Al2O3 catalyst had the lowest metal dispersion of 2.1%, and the metal dispersion of the Ni/MgO catalyst could not be measured because of the high reduction temperature. The Ni/MgO-Al2O3 catalyst, which exhibited changes in the textural properties of the support during heat treatment, showed a lower metal dispersion (4.6%) than the Ni/MgAl2O4 catalyst (6.6%). In addition, the commercial catalysts had very low metal dispersions, of less than 0.5%, compared those of the prepared catalysts.
The composition and crystalline phases of the catalysts were confirmed by XRD analysis. XRD patterns are shown in Figure 1. The Ni/Al2O3 catalyst formed a NiAl2O4 spinel crystalline phase. Nurunnabi et al. reported that the NiO peak shifted from 2θ = 43.26° to 2θ = 42.90–43.00° with the formation of a NiO-MgO solid solution [23]. The NiO peak of the prepared Ni/MgO catalyst also shifted to 2θ = 42.76°, owing to the formation of the NiO-MgO solid solution. In particular, Ni/MgO-Al2O3, Ni/MgAl2O4 catalysts displayed peaks corresponding to MgAl2O4 and NiO, and those corresponding to the NiAl2O4 and NiO-MgO solid solutions were not detected. The catalysts without alkaline earth metals presented peaks of NiAl2O4, but the Ca-containing commercial catalyst B showed a reduced formation of NiAl2O4 compared with catalysts using Al2O3 as a support. Therefore, the addition of alkaline earth metals effectively suppressed the formation of NiAl2O4 and NiO-MgO solid solutions.
H2-TPR analysis was conducted to analyze the reducibility of the catalysts (Figure 2). Pure NiO (unsupported) can be reduced in the 300–550 °C temperature range, while pure NiAl2O4 (unsupported) was reduced in the 730–900 °C temperature range [24,25]. In addition, a reduction peak associated with surface NiO weak interactions was formed in the temperature range of 500–550 °C, while the reduction peak formed in the temperature range of 700–800 °C was attributed to strong interactions between NiO and the support [26,27]. As bulk NiO has a weak interaction with the support, bulk NiO was mainly reduced at 250–500 °C [28]. A peak corresponding to the reduction of the commercial catalyst A was observed at 300–500 °C. The reduction peak corresponding to the commercial catalyst B was observed at 400–800 °C. Therefore, commercial catalysts were confirmed to have NiO with relatively weak interactions and bulk NiO. As confirmed by the XRD results, the Ni/Al2O3 catalyst, which was difficult to reduce due to the formation of NiAl2O4, underwent reduction at temperatures above 900 °C. In particular, because the Ni/MgO catalyst formed a stable NiO-MgO solid solution, the Ni/MgO catalyst was resistant to reduction up to 1000 °C. The Ni/MgO-Al2O3 and Ni/MgAl2O4 catalysts were mainly reduced at 800 °C. As the Ni/MgO-Al2O3 and Ni/MgAl2O4 catalysts had SMSIs, the Ni active metals were highly dispersed, and both Ni/MgO-Al2O3 and Ni/MgAl2O4 were reduced at higher temperatures compared to the commercial catalysts. In addition, because the Ni/MgO-Al2O3 and Ni/MgAl2O4 catalysts had a stable spinel structure (MgAl2O4) after calcination, the formation of NiAl2O4 or NiO-MgO solid solutions was suppressed. Therefore, the Ni/MgO-Al2O3 and Ni/MgAl2O4 catalysts displayed improved reducibility compared to the Ni/Al2O3 and Ni/MgO catalysts.
CO2-TPD analysis was conducted to determine the basic properties of the catalysts (Figure 3). In general, the catalytic CO2 adsorption significantly influenced the catalytic activity, as well as the oxygen species formed by the dissociation of the CO2 molecules, and suppressed carbon deposition [16]. The Ni/Al2O3 catalyst employing the acidic Al2O3 support showed low CO2 desorption characteristics of 26 μmol/g (Table 1). The Ni/MgO catalyst presented a CO2 desorption peak in the 100–700 °C temperature range, and exhibited the highest CO2 desorption amount (170 μmol/g), owing to the basic properties of the MgO support. In addition, because Ni/MgO-Al2O3 and Ni/MgAl2O4 catalysts contained the alkaline earth metal Mg, they exhibited improved CO2 adsorption properties compared to the Ni/Al2O3 catalyst. In particular, Ni/MgAl2O4, which had a high metal dispersion and large specific surface area, exhibited a higher CO2 desorption amount (125 μmol/g) than the Ni/MgO-Al2O3 catalyst (110 μmol/g). Commercial catalyst A, which employed the acidic Al2O3 support showed poor CO2 adsorption properties, due to the formation of bulk NiO. The commercial catalyst B presented CO2 desorption peaks at 200 and 500 °C, and an increased CO2 desorption amount of 81 μmol/g compared to the catalysts that employed Al2O3 as a support, due to the presence of the alkaline earth metal Ca.

2.2. Catalytic Test in DRM

The catalytic activity of the DRM reaction was evaluated in the temperature range of 600–750 °C, and the CH4 and CO2 conversions are presented in Figure 4. Since the DRM is an endothermic reaction, it exhibited high CH4 and CO2 conversions with increasing reaction temperature. The Ni/MgO catalyst, which formed a NiO-MgO solid solution, was reduced at temperatures above 1000 °C; therefore, the Ni/MgO catalyst did not show catalytic activity in the DRM reaction. Since the Ni/Al2O3 catalyst also formed a NiAl2O4 spinel, this resulted in low metal dispersion and low reducibility of the catalyst. For this reason, the Ni/Al2O3 catalyst showed a low CH4 conversion (66%) at 750 °C, whereas, Ni/MgO-Al2O3 and Ni/MgAl2O4 catalysts exhibited high CH4 conversions of 78 and 77%, respectively, compared with Ni/Al2O3. This is because of the high dispersion of the Ni metals due to the suppression of the formed NiAl2O4 and NiO-MgO solid solutions. On the other hand, commercial catalysts A and B showed low CH4 conversions, due to the formation of bulk NiO, despite the presence of 15 wt% Ni metal. However, because the alkaline earth metal-containing commercial catalyst B formed a lower amount of NiAl2O4 compared with the commercial catalyst A, the catalyst B presented relatively high CH4 and CO2 conversions. This result indicates that the catalytic characteristics (such as metal dispersion and crystalline phase), which depend on the type of support employed, played an important role in determining the catalytic activity.
H2 and CO yields and H2/CO ratio are presented in Figure 4 and Table 2, respectively. The CO yield (56–86%) was higher than the H2 yield (34–74%). In addition, the H2/CO ratio of the produced syngas after the DRM reaction was 0.6–0.9. These results are attributed to the RWGS reaction (Equation (2)) as a side reaction, as mentioned above. Theoretically, the ratio of the H2/CO produced in the DRM reaction (Equation (1)) is 1. However, the H2/CO ratio was lower than 1, because the RWGS reaction occurs as a side reaction under the reaction conditions of the DRM process [29]. Among the tested catalysts, Ni/MgO-Al2O3 and Ni/MgAl2O4 presented ratios close to the equilibrium H2/CO ratio of 0.9.

2.3. Coke Formation Study

To investigate the stability of the reaction and coke resistance, a stability test was conducted in the absence of the diluent (Figure 5). Coke formation reactions of the CH4 decomposition reaction (Equation (3)) and CO disproportionation reaction (Equation (4)) occur simultaneously in the temperature range of 557–700 °C [10]. Thus, stability tests were performed at 600 °C and a TOS of 20 h. As the Ni/MgO catalyst did not exhibit catalytic activity in the catalytic test, in accordance with the reaction temperature, a stability test was conducted, except for the Ni/MgO catalyst. The Ni/Al2O3 catalyst tended toward constantly reduced CH4 and CO2 conversions over 20 h. The Ni/Al2O3 catalyst had 21% of the initial CH4 conversion, the Ni/Al2O3 catalyst decreased to 16% of CH4 conversion after 20 h. Ni/MgO-Al2O3 and Ni/MgAl2O4 catalysts, which added the Mg alkaline earth metal, had high CH4 and CO2 conversions compared to the Ni/Al2O3 catalyst, owing to the improved reducibility of the catalyst and high metal dispersion. In particular, the Ni/MgO-Al2O3 and Ni/MgAl2O4 catalysts showed higher CH4 and CO2 conversions than the commercial catalyst A. On the other hand, the commercial catalyst B exhibited a CH4 conversion of 20–22%, and had a higher CH4 conversion than the catalyst that employed Al2O3 as a support.
To investigate the carbon formed on the surface of the catalyst, SEM and TGA analyses were performed on the recovered spent catalysts. The morphologies of the formed carbon were analyzed using SEM images, as shown in Figure 6. Sehested reported that three types of carbon were formed during reforming reactions: pyrolytic carbon, encapsulating carbon, and filamentous carbon [30]. Pyrolytic carbon is formed at temperatures higher than 600 °C, owing to thermal decomposition of hydrocarbon compounds. Since the encapsulating carbon covered particles of the catalyst, it caused immediate catalyst deactivation. Filamentous carbon is formed and grown between the Ni particles and the support, which causes catalyst pore plugging. For this reason, filamentous carbon results in catalyst deactivation, causing catalyst destruction and a pressure drop. The formation of filamentous carbon was also confirmed on the surface of the spent catalysts after the stability test, and SEM images of all the analyzed catalysts indicated the formation of a large amount of filamentous carbon on the catalyst surface.
Quantitative analysis of the amount of deposited carbon was performed through weight loss analysis using TGA up to 800 °C. The TGA results are shown in Figure 7 and Table 3. The weight loss of the spent catalyst was classified according to the temperature range; desorption of H2O and CO2 occurred in the temperature range of 250–300 °C, and the oxidation of carbon due to the release of CO and CO2 occurred in the temperature range of 480–640 °C [31]. In addition, graphitic filamentous carbon is mainly oxidized at a high temperature of 530 °C [32]. Quantitative analysis of the formed carbon was performed using the recovered spent catalysts containing alkaline earth metals, which showed a relatively high catalytic activity during the stability test. Results from the TG profiles indicated that most of the weight loss of the spent catalysts occurred at temperatures above 500 °C. The amount of deposited carbon was calculated from the weight loss, in the temperature range of 50–800 °C (Table 3). The amount of deposited carbon on the Ni/MgO-Al2O3 and Ni/MgAl2O4 catalysts was 21.4 and 29.3 wt%, respectively. In comparison, the commercial catalyst B contained twice the amount of deposited carbon (47.1 wt%). These results are related to the CO2 adsorption characteristics of the catalysts [33]. Since CO2 is an acidic gas, CO2 adsorption was enhanced by the addition of basic materials, such as Mg and Ca. In addition, oxygen species produced by the dissociation of CO2 molecules are used in the oxidation of deposited carbon on the catalyst surface. Therefore, increasing the basic properties of the DRM catalyst could improve catalyst durability. Therefore, commercial catalyst B, which showed relatively low CO2 adsorption properties (81 μmol/g), contained a large amount of carbon compared to the Ni/MgO-Al2O3 and Ni/MgAl2O4 catalysts, which had high CO2 adsorption properties, of 110 and 125 μmol/g, respectively.

3. Materials and Methods

3.1. Materials

Ni(NO3)2∙6H2O (97%, Junsei) was used as the active metal precursor; γ-Al2O3 (PURALOX TH 100/150, SASOL), MgO (≥99%, Sigma-Aldrich), MgO-Al2O3 hydrotalcite (MgO:Al2O3 = 30:70, SASOL), and MgAl2O4 (MG30 pre-calcined at 800 °C, SASOL) were used as supports for the Ni-based catalyst. Ni-based commercial catalysts Ni/Al2O3 (FCR-4; Clariant) and Ni/CaAl2O4 (R-70; Clariant)) were used as reference catalysts.

3.2. Catalyst Preparation

Ni-based catalysts were prepared using the incipient wetness impregnation method. Nickel nitrate was dissolved in distilled water and impregnated with γ-Al2O3. The prepared Ni/Al2O3 catalyst was dried overnight at room temperature, followed by calcination at 800 °C for 6 h in air. Ni/MgO, Ni/MgO-Al2O3, and Ni/MgAl2O4 catalysts were also prepared by the same method, and 10 wt% Ni active metal was supported on the catalysts.

3.3. Characteristics

The active metal content supported on the catalyst was analyzed using inductively coupled plasma-optical emission spectroscopy (ICP-OES, AVIO500, Perkin Elmer). The crystalline phases and composition of the catalysts were analyzed using an X-ray diffractometer (XRD, SmartLab High Temp, Rigaku, Tokyo, Japan). Diffraction peaks were obtained in the 2θ range of 10–90° with a step size of 0.02°. The specific surface area and pore size distribution of the catalyst were obtained using the Brunauer–Emmett–Teller and BarrettJoyner–Halenda methods (BET-BJH, BELSORP-MAX, MicrotracBEL Corp., Osaka, Japan). A sample (0.2 g) was pretreated at 300 °C for 3 h. N2 adsorption/desorption isotherms and pore size distributions were analyzed at −196 °C using liquid nitrogen. H2-chemisorption (BEL-METAL-3, MicrotracBEL Corp., Osaka, Japan) was performed to determine the metal dispersion of the catalyst. The sample (50 mg) was reduced at 700 °C for 2 h under H2 flow, and H2 gas was adsorbed at 50 °C using 20% H2/Ar gas. H2-temperature programmed reduction (H2-TPR, BELCAT-B, MicrotracBEL Corp., Osaka, Japan) analysis was performed to investigate the reducibility of the catalyst and the interaction between the active metal and support. The analysis was performed by increasing the temperature up to 1000 °C (10 °C/min) under 10% H2/Ar flow using a 50 mg sample. The CO2 adsorption properties were analyzed through CO2-temperature programmed desorption (CO2-TPD, BELCAT-B, MicrotracBEL Corp., Osaka, Japan) analysis. The sample (50 mg) was reduced at 700 °C for 2 h under 10% H2/Ar flow. The CO2 was adsorbed at 50 °C using 10% CO2/He gas and desorbed by increasing the temperature to 800 °C (10 °C/min). In addition, the morphology of the formed carbon in the spent catalysts was analyzed using scanning electron microscope (SEM, Regulus 8220, HITACHI, Tokyo, Japan) analysis. Quantitative analysis of carbon on the spent catalysts was performed using thermogravimetric analysis (TGA, TGA 92-18, Setaram, Caluire-et-Cuire, France). The weight loss was measured by loading 7 mg of the sample. The sample was oxidized by increasing the temperature to 800 °C at 10 °C/min in air.

3.4. Catalytic Tests

Catalytic activity was evaluated in a fixed bed reactor system. The 3/8′′ quartz reactor was filled with a catalyst (50 mg, 150–250 μm) and 1100 °C pre-calcined MgAl2O4 as a diluent (500 mg, 150–250 μm) in a ratio of 1:10. Before the catalytic test, the catalyst was reduced at 700 °C for 2 h with 10 vol.% H2/N2 gas. The evaluation of the catalytic activity was performed in a temperature range of 600–750 °C. The feeds of CH4, CO2, and N2 were injected to equal ratio. The total flow rate was 120 mL/min, and GHSV corresponded to 144,000 mL/gcat∙h. The reaction temperature was controlled using a K-type thermocouple (TC), the TC was located inside the catalyst layer. The produced gas was detected using a micro-gas chromatograph (3000A Micro GC, Agilent, CA, USA) equipped with a thermal conductivity detector (TCD) using MolSeive 5A and PlotU columns.

4. Conclusions

DRM reaction over Ni-based catalysts was studied as a function of the type of catalyst support used (Al2O3, MgO, MgO-Al2O3 (LDH), and MgAl2O4 (spinel)). The type of support significantly influenced the reducibility and metal dispersion of the Ni-based catalysts. The Ni/Al2O3 and Ni/MgO catalysts had low metal dispersion, due to the formation of NiAl2O4 and NiO-MgO solid solutions, respectively. In addition, these catalysts had a strong interaction between the Ni active metal and support, indicated by the high temperature (more than 900 °C) required for reduction. In contrast, Ni/MgO-Al2O3 and Ni/MgAl2O4 catalysts suppressed the formation of NiAl2O4 and NiO-MgO solid solutions, exhibited high metal dispersion, and were reduced at lower temperatures compared with the Ni/Al2O3 and Ni/MgO catalysts. Moreover, Ni/MgO-Al2O3 and Ni/MgAl2O4 catalysts containing Mg had enhanced CO2 adsorption properties. In the DRM reaction, Ni/MgO-Al2O3 and Ni/MgAl2O4 catalysts exhibited high CH4 conversion (>75%) at 750 °C compared to commercial catalysts. Furthermore, Ni/MgO-Al2O3 and Ni/MgAl2O4 catalysts showed enhanced coke resistance compared to commercial catalyst B. Therefore, the metal dispersion, basic properties, and metal-support interaction of the catalyst in accordance with the type of support were confirmed to significantly influence the catalyst performance in the DRM reaction.

Author Contributions

Conceptualization, U.H.J. and K.Y.K.; methodology, D.H.S.; formal analysis, Y.E.K. and H.B.I.; data curation, T.H.L.; writing-original draft preparation, D.H.S.; writing—review and editing, K.Y.K.; supervision, K.Y.K. and K.B.L.; project administration, U.H.J. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Ministry of Trade, Industry, and Energy (MOTIE) and the Korea Institute of Energy Technology Evaluation and Planning (KETEP) of the Republic of Korea (No. 20203030030080).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Subramanian, S.; Song, Y.; Kim, D.; Yavuz, C.T. Redox and nonredox CO2 utilization: Dry reforming of methane and catalytic cyclic carbonate formation. ACS Energy Lett. 2020, 5, 1689–1700. [Google Scholar] [CrossRef]
  2. Parsapur, R.K.; Chatterjee, S.; Huang, K.-W. The insignificant role of dry reforming of methane in CO2 emission relief. ACS Energy Lett. 2020, 5, 2881–2885. [Google Scholar] [CrossRef]
  3. San-José-Alonso, D.; Juan-Juan, J.; Illán-Gómez, M.; Román-Martínez, M. Ni, Co and bimetallic Ni–Co catalysts for the dry reforming of methane. Appl. Catal. A 2009, 371, 54–59. [Google Scholar] [CrossRef]
  4. Edwards, J.; Maitra, A. The chemistry of methane reforming with carbon dioxide and its current and potential applications. Fuel Process. Technol. 1995, 42, 269–289. [Google Scholar] [CrossRef]
  5. Zhang, M.; Shengfu, J.; Linhua, H.; Fengxiang, Y.; Chengyue, L.; Hui, L. Structural characterization of highly stable Ni/SBA-15 catalyst and its catalytic performance for methane reforming with CO2. Chin. J. Catal. 2006, 27, 777–781. [Google Scholar] [CrossRef]
  6. Arora, S.; Prasad, R. An overview on dry reforming of methane: Strategies to reduce carbonaceous deactivation of catalysts. RSC Adv. 2016, 6, 108668–108688. [Google Scholar] [CrossRef]
  7. Abdullah, B.; Abd Ghani, N.A.; Vo, D.-V.N. Recent advances in dry reforming of methane over Ni-based catalysts. J. Clean. Prod. 2017, 162, 170–185. [Google Scholar] [CrossRef] [Green Version]
  8. Budiman, A.W.; Song, S.-H.; Chang, T.-S.; Shin, C.-H.; Choi, M.-J. Dry reforming of methane over cobalt catalysts: A literature review of catalyst development. Catal. Surv. Asia 2012, 16, 183–197. [Google Scholar] [CrossRef]
  9. Praserthdam, S.; Somdee, S.; Rittiruam, M.; Balbuena, P.B. Computational study of the evolution of Ni-based catalysts during the dry reforming of methane. Energy Fuels 2020, 34, 4855–4864. [Google Scholar] [CrossRef]
  10. Wang, S.; Lu, G.; Millar, G.J. Carbon dioxide reforming of methane to produce synthesis gas over metal-supported catalysts: State of the art. Energy Fuels 1996, 10, 896–904. [Google Scholar] [CrossRef]
  11. Ginsburg, J.M.; Piña, J.; El Solh, T.; De Lasa, H.I. Coke formation over a nickel catalyst under methane dry reforming conditions: Thermodynamic and kinetic models. Ind. Eng. Chem. Res. 2005, 44, 4846–4854. [Google Scholar] [CrossRef]
  12. ZHANG, R.-j.; XIA, G.-f.; LI, M.-f.; Yu, W.; Hong, N.; LI, D.-d. Effect of support on the performance of Ni-based catalyst in methane dry reforming. J. Fuel Chem. Technol. 2015, 43, 1359–1365. [Google Scholar] [CrossRef]
  13. Rezaei, M.; Alavi, S.; Sahebdelfar, S.; Bai, P.; Liu, X.; Yan, Z.-F. CO2 reforming of CH4 over nanocrystalline zirconia-supported nickel catalysts. Appl. Catal. B 2008, 77, 346–354. [Google Scholar] [CrossRef]
  14. Bychkov, V.Y.; Tyulenin, Y.P.; Firsova, A.; Shafranovsky, E.; Gorenberg, A.Y.; Korchak, V. Carbonization of nickel catalysts and its effect on methane dry reforming. Appl. Catal. A 2013, 453, 71–79. [Google Scholar] [CrossRef]
  15. Li, Y.; Lu, G.; Ma, J. Highly active and stable nano NiO–MgO catalyst encapsulated by silica with a core–shell structure for CO2 methanation. RSC Adv. 2014, 4, 17420–17428. [Google Scholar] [CrossRef]
  16. Tomishige, K.; Yamazaki, O.; Chen, Y.; Yokoyama, K.; Li, X.; Fujimoto, K. Development of ultra-stable Ni catalysts for CO2 reforming of methane. Catal. Today 1998, 45, 35–39. [Google Scholar] [CrossRef]
  17. Özdemir, H.; Öksüzömer, M.F.; Gürkaynak, M.A. Preparation and characterization of Ni based catalysts for the catalytic partial oxidation of methane: Effect of support basicity on H2/CO ratio and carbon deposition. Int. J. Hydrog. Energy 2010, 35, 12147–12160. [Google Scholar] [CrossRef]
  18. Bobadilla, L.; Garcilaso, V.; Centeno, M.; Odriozola, J. CO2 reforming of methane over Ni-Ru supported catalysts: On the nature of active sites by operando DRIFTS study. J. CO2 Util. 2018, 24, 509–515. [Google Scholar]
  19. Koo, K.Y.; Roh, H.-S.; Seo, Y.T.; Seo, D.J.; Yoon, W.L.; Park, S.B. Coke study on MgO-promoted Ni/Al2O3 catalyst in combined H2O and CO2 reforming of methane for gas to liquid (GTL) process. Appl. Catal. A 2008, 340, 183–190. [Google Scholar] [CrossRef]
  20. Djaidja, A.; Libs, S.; Kiennemann, A.; Barama, A. Characterization and activity in dry reforming of methane on NiMg/Al and Ni/MgO catalysts. Catal. Today 2006, 113, 194–200. [Google Scholar] [CrossRef]
  21. Gao, X.; Ge, Z.; Zhu, G.; Wang, Z.; Ashok, J.; Kawi, S. Anti-coking and anti-sintering Ni/Al2O3 catalysts in the dry reforming of methane: Recent progress and prospects. Catalysts 2021, 11, 1003. [Google Scholar] [CrossRef]
  22. Zanganeh, R.; Rezaei, M.; Zamaniyan, A. Preparation of nanocrystalline NiO–MgO solid solution powders as catalyst for methane reforming with carbon dioxide: Effect of preparation conditions. Adv. Powder Technol. 2014, 25, 1111–1117. [Google Scholar] [CrossRef]
  23. Nurunnabi, M.; Li, B.; Kunimori, K.; Suzuki, K.; Fujimoto, K.-I.; Tomishige, K. Performance of NiO–MgO solid solution-supported Pt catalysts in oxidative steam reforming of methane. Appl. Catal. A 2005, 292, 272–280. [Google Scholar] [CrossRef]
  24. Mori, H.; Wen, C.-j.; Otomo, J.; Eguchi, K.; Takahashi, H. Investigation of the interaction between NiO and yttria-stabilized zirconia (YSZ) in the NiO/YSZ composite by temperature-programmed reduction technique. Appl. Catal. A 2003, 245, 79–85. [Google Scholar] [CrossRef]
  25. Becerra, A.M.; Castro-Luna, A.E. An investigation on the presence of NiAl2O4 in a stable Ni on alumina catalyst for dry reforming. J. Chil. Chem. Soc. 2005, 50, 465–469. [Google Scholar] [CrossRef]
  26. Guo, J.; Lou, H.; Zhao, H.; Chai, D.; Zheng, X. Dry reforming of methane over nickel catalysts supported on magnesium aluminate spinels. Appl. Catal. A 2004, 273, 75–82. [Google Scholar] [CrossRef]
  27. Alipour, Z.; Rezaei, M.; Meshkani, F. Effect of alkaline earth promoters (MgO, CaO, and BaO) on the activity and coke formation of Ni catalysts supported on nanocrystalline Al2O3 in dry reforming of methane. J. Ind. Eng. Chem. 2014, 20, 2858–2863. [Google Scholar] [CrossRef]
  28. Gao, J.; Jia, C.; Zhang, M.; Gu, F.; Xu, G.; Su, F. Effect of nickel nanoparticle size in Ni/α-Al2O3 on CO methanation reaction for the production of synthetic natural gas. Catal. Sci. Technol. 2013, 3, 2009–2015. [Google Scholar] [CrossRef]
  29. Serrano-Lotina, A.; Daza, L. Influence of the operating parameters over dry reforming of methane to syngas. Int. J. Hydrog. Energy 2014, 39, 4089–4094. [Google Scholar] [CrossRef]
  30. Sehested, J. Four challenges for nickel steam-reforming catalysts. Catal. Today 2006, 111, 103–110. [Google Scholar] [CrossRef]
  31. Tsyganok, A.I.; Tsunoda, T.; Hamakawa, S.; Suzuki, K.; Takehira, K.; Hayakawa, T. Dry reforming of methane over catalysts derived from nickel-containing Mg–Al layered double hydroxides. J. Catal. 2003, 213, 191–203. [Google Scholar] [CrossRef]
  32. Wang, S.; Lu, G.Q. Effects of promoters on catalytic activity and carbon deposition of Ni/γ-Al2O3 catalysts in CO2 reforming of CH4. J. Chem. Technol. Biotechnol. 2000, 75, 589–595. [Google Scholar] [CrossRef]
  33. Singha, R.K.; Yadav, A.; Agrawal, A.; Shukla, A.; Adak, S.; Sasaki, T.; Bal, R. Synthesis of highly coke resistant Ni nanoparticles supported MgO/ZnO catalyst for reforming of methane with carbon dioxide. Appl. Catal. B 2016, 191, 165–178. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of Ni catalysts with various supports.
Figure 1. XRD patterns of Ni catalysts with various supports.
Catalysts 12 00216 g001
Figure 2. H2-TPR profiles of Ni catalysts with various supports.
Figure 2. H2-TPR profiles of Ni catalysts with various supports.
Catalysts 12 00216 g002
Figure 3. CO2-TPD profiles of Ni catalysts with various supports.
Figure 3. CO2-TPD profiles of Ni catalysts with various supports.
Catalysts 12 00216 g003
Figure 4. CH4 conversion (a), CO2 conversion (b), H2 yield (c), and CO yield (d) with reaction temperature over Ni catalysts (reaction conditions: Temp. = 600–750 °C, CH4/CO2/N2 = 1/1/1, GHSV = 144,000 mL/gcat∙h).
Figure 4. CH4 conversion (a), CO2 conversion (b), H2 yield (c), and CO yield (d) with reaction temperature over Ni catalysts (reaction conditions: Temp. = 600–750 °C, CH4/CO2/N2 = 1/1/1, GHSV = 144,000 mL/gcat∙h).
Catalysts 12 00216 g004
Figure 5. CH4 conversion (a) and CO2 conversion (b) with long-term test over Ni catalysts (reaction conditions: Temp. = 600 °C, CH4/CO2/N2 = 1/1/1, GHSV = 144,000 mL/gcat∙h, TOS = 20 h).
Figure 5. CH4 conversion (a) and CO2 conversion (b) with long-term test over Ni catalysts (reaction conditions: Temp. = 600 °C, CH4/CO2/N2 = 1/1/1, GHSV = 144,000 mL/gcat∙h, TOS = 20 h).
Catalysts 12 00216 g005
Figure 6. SEM images of fresh (a) and spent (b) catalysts with various supports (reaction conditions: Temp. = 600 °C, CH4/CO2/N2 = 1/1/1, GHSV = 144,000 mL/gcat∙h, TOS = 20 h).
Figure 6. SEM images of fresh (a) and spent (b) catalysts with various supports (reaction conditions: Temp. = 600 °C, CH4/CO2/N2 = 1/1/1, GHSV = 144,000 mL/gcat∙h, TOS = 20 h).
Catalysts 12 00216 g006
Figure 7. TG profiles of spent catalysts under air (reaction conditions: Temp. = 600 °C, CH4/CO2/N2 = 1/1/1, GHSV = 144,000 mL/gcat∙h, TOS = 20 h).
Figure 7. TG profiles of spent catalysts under air (reaction conditions: Temp. = 600 °C, CH4/CO2/N2 = 1/1/1, GHSV = 144,000 mL/gcat∙h, TOS = 20 h).
Catalysts 12 00216 g007
Table 1. Characteristics of supports and Ni-based catalysts.
Table 1. Characteristics of supports and Ni-based catalysts.
Ni Content (wt%)Specific Surface Area 1 (m2/g)Metal Dispersion 2 (%)H2 Uptake 3 (μmol/g)CO2 Desorption Amount 4 (μmol/g)
Al2O3-132.2---
MgO124.0
MgO-Al2O3177.6
MgAl2O4135.1
Ni/Al2O39.2103.52.19426
Ni/MgO9.217.7N/A 539170
Ni/MgO-Al2O38.9108.74.698110
Ni/MgAl2O49.1113.36.696125
Commercial A15.63.70.41059
Commercial B15.924.60.518281
1 Estimated from N2 adsorption at −196 °C. 2 Estimated from H2-Chemisorption at 50 °C. 3 Estimated from H2-TPR in the temperature range 30–1000 °C. 4 Estimated from CO2-TPD analysis in the temperature range 50–800 °C. 5 N/A: Not available.
Table 2. H2/CO ratio of Ni catalysts with various supports (reaction conditions: Temp. = 750 °C, CH4/CO2/N2 = 1/1/1, GHSV = 144,000 mL/gcat∙h).
Table 2. H2/CO ratio of Ni catalysts with various supports (reaction conditions: Temp. = 750 °C, CH4/CO2/N2 = 1/1/1, GHSV = 144,000 mL/gcat∙h).
Ni/Al2O3Ni/MgONi/MgO-Al2O3Ni/MgAl2O4Commercial ACommercial B
H2/CO ratio0.8N/A 10.90.90.60.8
1 N/A: Not available.
Table 3. Amount of coke over Ni catalysts from TG profiles (reaction conditions: Temp. = 600 °C, CH4/CO2/N2 = 1/1/1, GHSV = 144,000 mL/gcat∙h, TOS = 20 h).
Table 3. Amount of coke over Ni catalysts from TG profiles (reaction conditions: Temp. = 600 °C, CH4/CO2/N2 = 1/1/1, GHSV = 144,000 mL/gcat∙h, TOS = 20 h).
CatalystsAmount of Coke 1 [wt%]
Ni/MgO-Al2O321.4
Ni/MgAl2O429.3
Commercial B47.1
1 Calculated from the following equation; ( Amount   of   coke ) = ( s a m p l e   w e i g h t   a t   50   ° C ) ( s a m p l e   w e i g h t   a t   800   ° C ) ( s a m p l e   w e i g h t   a t   50   ° C ) × 100 .
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Song, D.H.; Jung, U.H.; Kim, Y.E.; Im, H.B.; Lee, T.H.; Lee, K.B.; Koo, K.Y. Influence of Supports on the Catalytic Activity and Coke Resistance of Ni Catalyst in Dry Reforming of Methane. Catalysts 2022, 12, 216. https://doi.org/10.3390/catal12020216

AMA Style

Song DH, Jung UH, Kim YE, Im HB, Lee TH, Lee KB, Koo KY. Influence of Supports on the Catalytic Activity and Coke Resistance of Ni Catalyst in Dry Reforming of Methane. Catalysts. 2022; 12(2):216. https://doi.org/10.3390/catal12020216

Chicago/Turabian Style

Song, Da Hye, Un Ho Jung, Young Eun Kim, Hyo Been Im, Tae Ho Lee, Ki Bong Lee, and Kee Young Koo. 2022. "Influence of Supports on the Catalytic Activity and Coke Resistance of Ni Catalyst in Dry Reforming of Methane" Catalysts 12, no. 2: 216. https://doi.org/10.3390/catal12020216

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