Next Article in Journal
Carbon-Supported Pt-SnO2 Catalysts for Oxygen Reduction Reaction over a Wide Temperature Range: Rotating Disk Electrode Study
Next Article in Special Issue
Sulfur-Resistant CeO2-Supported Pt Catalyst for Waste-to-Hydrogen: Effect of Catalyst Synthesis Method
Previous Article in Journal
Au/CeO2 Photocatalyst for the Selective Oxidation of Aromatic Alcohols in Water under UV, Visible and Solar Irradiation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

CO2 Reforming of CH4 Using Coke Oven Gas over Ni/MgO-Al2O3 Catalysts: Effect of the MgO:Al2O3 Ratio

1
Department of Environmental Engineering, Yonsei University, 1 Yonseidae-gil, Wonju 26493, Gangwon-do, Korea
2
Department of Earth Resources and Environmental Engineering, Hanyang University, 222 Wangsimni-ro, Seongdong-gu, Seoul 04763, Korea
3
School of Chemical Engineering, Sungkyunkwan University (SKKU), 2066 Seobu-ro, Jangan-gu, Suwon 16419, Korea
*
Author to whom correspondence should be addressed.
Submission received: 5 November 2021 / Revised: 26 November 2021 / Accepted: 29 November 2021 / Published: 30 November 2021
(This article belongs to the Special Issue Nanocatalysts for Carbon Upcycling)

Abstract

:
Research is being actively conducted to improve the carbon deposition and sintering resistance of Ni-based catalysts. Among them, the Al2O3-supported Ni catalyst has been broadly studied for the dry reforming reaction due to its high CH4 activity at the beginning of the reaction. However, there is a problem of deactivation due to carbon deposition of Ni/Al2O3 catalyst and sintering of Ni, which is a catalytically active material. Supplementing MgO in Ni/Al2O3 catalyst can result in an improved MgAl2O4 spinel structure and basicity, which can be helpful for the activation of methane and carbon dioxide molecules. In order to confirm the optimal supports’ ratio in Ni/MgO-Al2O3 catalysts, the catalysts were prepared by supporting Ni after controlling the MgO:Al2O3 ratio stepwise, and the prepared catalysts were used for CO2 reforming of CH4 (CDR) using coke oven gas (COG). The catalytic reaction was conducted at 800 °C and at a high gas hourly space velocity (GHSV = 1,500,000 h−1) to screen the catalytic performance. The Ni/MgO-Al2O3 (MgO:Al2O3 = 3:7) catalyst showed the best catalytic performance between prepared catalysts. From this study, the ratio of MgO:Al2O3 was confirmed to affect not only the basicity of the catalyst but also the dispersion of the catalyst and the reducing property of the catalyst surface.

Graphical Abstract

1. Introduction

Since the adoption of the Paris Agreement on Climate Change in 2015, it has been necessary to put in place measures to cut down greenhouse gas emissions to limit the increase in the global average temperature to 1.5 °C relative to preindustrial levels [1,2]. To achieve the CO2 levels required by the Paris Agreement, we must reach zero emissions by 2050 [1]. However, zeroing greenhouse gases is still difficult due to the following issues. First, as the population increases, energy consumption continues to increase [3,4]. Among them, fossil fuels account for 81% of the total energy consumption [5]. Second, some carbon-based resources are essential for industries, such as steel making, where high-temperature heat is required [6,7,8]. Copious amounts of by-product gas are generated in each unit process in the steel industry, which is the core industry of our society [9]. Greenhouse gas emissions in the form of by-products accounted for approximately 9% of global emissions from 1900 to 2015 [10]. The by-product gas of the steel industry is largely divided into coke oven gas (COG), which is generated in the process of producing coke by oxidizing coal in a coke oven, and CO2 generated in a blast furnace when iron oxide is reduced to iron [11]. Coke oven gas, a representative steel by-product gas, contains approximately 27% methane (CH4), which causes greenhouse gas problems when emitted (coke oven gas composition: CH4: 28.0%, CO2: 2.6%, CO: 8.0%, H2: 55.6%, N2: 5.8%) [12]. A steam-reforming reaction (SRM: CH4 + H2O → CO + 3H2) that can produce syngas from methane is being studied to utilize COG, but there are economic problems with the additional supply of steam [13,14].
The CO2 reforming of CH4 (CDR: CH4 + CO2 → 2CO + 2H2) has the advantage of using CH4 and CO2 as reactants to produce syngas, which is an elemental raw material with high added value [15]. In addition, the syngas produced by the dry reforming reaction (H2:CO = 1:1) shows a favorable H2:CO ratio for oxo compound synthesis compared to the syngas produced through the steam reforming reaction (H2:CO = 3:1) [16]. However, as the dry reforming reaction proceeds at high temperatures above 700 °C, favorable conditions are formed for sintering of active metals and supports to occur, and catalyst deactivation may occur due to carbon deposition by methane decomposition [17]. In addition, the dry reforming reaction is a strong endothermic reaction, partially lowering the catalyst bed temperature and promoting the occurrence of the Boudouard reaction (2CO → C + CO2), causing carbon deposition on the catalyst surface, thereby deactivating the catalyst [18,19]. Therefore, to overcome these problems, it is necessary to develop a catalyst for dry reforming with high sintering and carbon deposition resistances.
Various studies have been conducted using noble metal (Ir, Rh, Ru, Pt, and Pd) catalysts, and strong sintering resistance and carbon deposition resistance have been demonstrated in the dry reforming reaction. However, there is a problem in entering the commercialization stage due to the low economic feasibility of the noble metal catalyst [20,21]. Therefore, although research on economical nickel (Ni)-based catalysts exhibiting excellent activity in dry reforming reactions is being actively conducted, the problem of catalyst deactivation due to the low sintering resistance and low carbon deposition resistance of Ni has become a major issue [22,23,24,25,26].
Research is being actively conducted to improve the catalyst deactivation problem of Ni-based catalysts. Out of them, the Al2O3-supported Ni catalyst has been broadly studied for the dry reforming reaction due to its high CH4 activity at the beginning of the reaction [27,28,29].
However, there is a problem of deactivation due to carbon deposition of the Ni/Al2O3 catalyst and sintering of Ni, which is a catalytically active material [30].
Therefore, many studies have been conducted to improve the performance of Ni/Al2O3 catalysts. Alipour et al. reported that adding alkaline earth metal oxides (CaO or MgO), dopants can enhance the interaction between Ni and the Al2O3 support [31]. Sengupta et al. demonstrated that the surface basicity of the alkaline earth metal oxide can improve the activity of methane molecules, thereby alleviating the problem of carbon deposition in the catalyst during the dry reforming reaction [32]. Guo et al. found that MgAl2O4 can be used as a support for the dry reforming catalytic reaction of Ni because of its high melting point, good chemical stability, and low acidity [33].
Based on the studies outlined above, the supplementing of MgO in Ni/Al2O3 catalyst can result in an improved MgAl2O4 spinel structure and basicity, which can be helpful for the activation of CH4 and CO2 molecules.
In this study, Ni/MgO-Al2O3 catalysts prepared with different MgO:Al2O3 ratios were used with coke oven gas (COG) to compare their dry reforming performance (gas hourly space velocity (GHSV) = 1,500,000 h−1) and to investigate how that performance relates to the characteristics of the catalysts. The correlations between the catalytic performance and the physicochemical properties were investigated.

2. Results and Discussion

MgO-Al2O3 was prepared by the one-step coprecipitation method with different MgO:Al2O3 ratios as a support. A Ni/MG catalyst was prepared by supporting a Ni precursor aqueous solution on the prepared MgO-Al2O3 support by the impregnation method, as shown in Scheme 1. Before chemisorption and CDR reaction, in situ reduction was performed for 3 h at 800 °C. Prior to XRD and TEM analyses, ex situ reduction was carried out for 3 h at 800 °C, and passivation was conducted for 3 h at 25 °C (Scheme 1).

2.1. Catalyst Characterization

The BET surface areas (S.A.) of catalysts are listed in Table 1. The BET S.A. of the prepared MgO-Al2O3 support decreased as the MgO content increased and further decreased as Ni was loaded onto the MgO-Al2O3 support. As a result, the BET S.A. of the prepared Ni/MG catalyst was shown to have the following order: Ni/MG10 (139.2 m2/g) > Ni/MG30 (120.8 m2/g) > Ni/MG50 (104.2 m2/g) > Ni/MG70 (104.0 m2/g) > Ni/MG90 (76.6 m2/g). When NiO is impregnated on a support, the BET S.A. of the catalyst has been shown to be reduced compared to that of the support because NiO fills the pores of the support [34,35,36]. Figure S1 shows the nitrogen isotherm adsorption and desorption results of the catalysts. The type IV hysteresis loop is identified in all catalysts, which indicates that the catalysts have mesoporous structures [16]. The mesoporous structure of Ni-based catalysts is known to provide accessible Ni active centers and to stabilize Ni particles by confinement effect [18]. The dispersion of Ni from the prepared catalysts was calculated using the results of the H2-chemisorption analysis and is shown in Table 1. A relatively high level of Ni dispersion was confirmed in the Ni/MG30, Ni/MG50, Ni/MG70, and Ni/MG90 catalysts compared to the Ni/MG10 catalyst. The Ni/MG30 catalyst showed the highest Ni dispersion (3.41%) among the prepared catalysts.
Figure 1 demonstrates the XRD analysis results of the Ni/MG catalysts by diverse MgO:Al2O3 ratios. Peaks at 31.4, 37.0, 45.0, 59.7, and 65.5° were attributed to spinel NiAl2O4 (JCPDS #10-0339), and only the highest Al2O3-containing Ni/MG10 catalyst was identified. Peaks at 31.3, 36.9, 44.8, 59.4, and 65.2° were assigned to the spinel MgAl2O4 (JCPDS #21-1152) crystals, which can be identified with the increasing MgO content. The Ni/MG30 catalyst exhibited the strongest spinel MgAl2O4 peak intensity. The peaks at approximately 37, 43, 62, 75, and 79° were due to the cubic structure of unreduced NiO (JCPDS #04-0835) and MgO (JCPDS #65-0476). It was difficult to differentiate the positions of the NiO and MgO XRD diffraction peaks. As the MgO content increased, the intensity of the peak corresponding to the NiO-MgO solid solution produced by the interaction of NiO and MgO remaining after MgAl2O4 formation increased. All the reduced catalysts exhibited Ni0 peaks at 44.5, 51.8, and 76.4°, which is an active phase during dry reforming of COG [14]. The crystallite size of Ni0 was calculated from the peak (51.8°) assigned to the Ni0 (2 0 0) plane using the Scherrer equation due to the peak overlap at the Ni0 (1 1 1) plane peak, and the results are given in Table 1. In the prepared catalysts, the Ni/MG30 catalyst exhibited the smallest Ni crystal size (9.4 nm).
The oxidation/reduction properties of the catalysts were the primary factors influencing the catalyst activity in reforming reactions. TPR analysis was organized to understand the oxidation/reduction characteristics of the Ni/MG catalysts prepared with various MgO:Al2O3 ratios, and the results are shown in Figure 2. All prepared catalysts demonstrated large reduction peaks at 700–850 °C [27]. This was a peak owing to the reduction of the complex NiO species that interacted strongly with the support.
The composite NiO reduction peak of the Ni/MG catalyst at low temperature appeared with the following order: Ni/MG30 < Ni/MG10 < Ni/MG70 < Ni/MG50 < Ni/MG90. In the case of catalyst Ni/MG10, a high Al2O3 content was reported to be advantageous for forming a NiAl2O4 spinel structure; and when the NiAl2O4 spinel structure was formed, a reduced peak appeared at high temperature because of the strong interaction between NiO and Al2O3 [37,38,39]. On the other hand, Ni/MG30 was reduced at a lower temperature than the Ni/MG10 catalyst by the reduction of the NiO interacting with the MgAl2O4 spinel structure. However, with an increase in MgO, the reduction peak shifted to a higher temperature because of the strong interaction following the formation of the NiO-MgO solid solution [40,41].
Table 2 shows the results of the reduction degree analysis for Ni/MG catalysts under the reaction conditions. The reduction degree value is closely related to the number of active species in the reaction conditions. Therefore, the Ni/MG30 catalyst showing the highest value was expected to show the highest activity when applied to the dry reforming reaction.
CO2-TPD analysis was used to examine the interaction of CO2 with the synthesized Ni/MG catalyst. As shown in Figure 3, the Ni/MG catalyst exhibited a major desorption peak at 100–400 °C, which means that there was a strong basic site on the Ni/MG surface. According to the literature, the chemical properties of the support, especially the basicity, promoted the activation of CO2 in the dry reforming reaction [42,43]. In addition, the conversion of both CH4 and CO2 was improved by increasing the removal rate of surface carbonaceous species, such as CHx (x = 0–3), generated by the dehydrogenation of CH4. The high CO2 adsorption capacity also improved the coke resistance of the catalyst [42,43]. Table 3 shows the amount of CO2 desorbed by the prepared MG supports and Ni/MG catalysts. In the MG supports, as the ratio of MgO increased, the basicity of the support increased stepwise and then decreased again at the ratio of MgO:Al2O3 = 9:1. The Ni/MG30 catalyst exhibited the highest CO2 adsorption amount (9.7 cm3/gcat).
TEM and EDS mapping analysis was performed to confirm the supported state and composition of the prepared catalyst. Figure 4 shows the TEM and EDS mapping images of Ni/MG30 catalyst. The composition of the prepared Ni/MG catalyst followed (Ni: 13.28%, Mg: 28.01%, Al: 58.70%).
The elemental composition analysis of the Ni/MG catalyst prepared by inductively coupled plasma (ICP) was analyzed and is listed in Table S1. In ICP analysis results, for all the prepared catalysts, almost same catalyst composition was achieved, which is what we intended.

2.2. Reaction Results

Figure 5 shows the interrelationship between the physicochemical properties and the catalytic performance of the Ni/MG catalysts with various MgO:Al2O3 ratios. The BET S.A. of the prepared MgO-Al2O3 support decreased as the MgO content increased. In the case of the Ni/MG10 catalyst with the highest Al2O3 content in the support, the complex NiO reduction temperature increased, according to the formation of spinel NiAl2O4, and the catalyst showed low dispersion and a large crystallite size. As the MgO content increased, the complex NiO reduction temperature of complex NiO decreased according to the MgAl2O4 production. In the case of an excessive MgO content, a NiO-MgO solid solution, in which the excess MgO strongly interacted with NiO, was formed, the NiO reduction temperature increased to a high temperature, the dispersion decreased, and the crystallite size increased. The Ni/MG30 catalyst showed the highest CO2 adsorption amount. As a result, the Ni/MG30 catalyst showed an appropriate correlation between complex NiO and the MgO-Al2O3, having the largest number of Ni0 active sites.
Figure 6 shows the CH4 conversion of Ni/MG catalysts. To screen the prepared catalyst performance, the reaction was performed under high GHSV conditions (GHSV = 1,500,000 h−1) at 800 °C. As a result of the reaction, the Ni/MG30 catalyst had the highest CH4 conversion (84.2%), and the Ni/MG90 catalyst had the lowest CH4 conversion (66.6%). Even at high GHSV, all the prepared catalysts showed stable activity for 12 h.
Figure 7 shows the CO2 conversion of the Ni/MG catalysts. The CO2 conversion was approximately 3% higher than the CH4 conversion, and the H2 yield was lower than the theoretical H2 yield (H2 yield (%): Ni/MG10 (77.6), Ni/MG30 (81.9), Ni/MG50 (79.8), Ni/MG70 (79.4), Ni/MG90 (53.9)). This was judged to be the result of the reverse water gas shift reaction (RWGS: H2 + CO2 → H2O + CO) during the dry reforming reaction [44]. The detailed reaction results, including the feed gas and outlet gas compositions, are shown in Table 4.

3. Materials and Methods

3.1. Preparation of Catalysts

MgO–Al2O3 was prepared using a one-step coprecipitation method with different MgO:Al2O3 ratios as a support. Mg(NO3)2·6H2O (99%, Aldrich) and Al(NO3)2·6H2O (98%, Aldrich) were used as the precursors. KOH (95%, Samchun) was utilized as the precipitation agent. To support preparation, the precursors were stoichiometrically quantified and then dissolved in distilled water and a constant temperature at 80 °C. In a precipitation step, 15 wt% KOH solution was added at 0.8 mL/min to reach pH 10.5. The prepared solution was aged for 3 days and was then washed several times with distilled water to remove impurities, including the remaining K+ and NO3- ions. The precipitate was sufficiently dried at 100 °C to remove moisture. After that, it was calcined at 800 °C for 6 h. The Ni/MgO-Al2O3 catalyst was prepared through an impregnation method by supporting the Ni(NO3)2·6H2O (97%, Junsei) on the prepared MgO-Al2O3 support. The Ni loading was 15 wt. %, and the catalyst was dried at 100 °C and then calcined at 800 °C for 6 h. The finished Ni/MgO-Al2O3 catalyst was named according to the MgO:Al2O3 ratio, as indicated in Table 5.

3.2. Catalyst Characterization

Transmission electron microscopy (TEM) images of the samples were obtained using a JEM-F200 (JEOL, Japan) microscope. Samples were blended with ethanol by using ultra-sonication for 30 min. The prepared suspension was deposited on a grid. The elemental composition of prepared catalysts was analyzed by an inductively coupled, plasma-optical emission spectrometer (Varian, USA). The Brunauer–Emmett–Teller specific surface areas of the samples were investigated by the N2 adsorption/desorption isotherms at −196 °C using an ASAP 2010 (Micromeritics, USA). X-ray diffraction (XRD) patterns were analyzed using a Rigaku Ultima IV diffractometer (RIGAKU, Japan). Detailed procedures of XRD and Ni0 crystallite size calculation method were reported in our previous work [14]. H2-chemisorption data were obtained using an Autochem 2920 (Micromeritics, USA). The redox properties of the catalysts were evaluated via H2 temperature-programmed reduction (H2-TPR) using Autochem 2920. The reduction degree was estimated using the following equation, derived from the integration of the peak area below the TPR at 800 °C and 1000 °C:
Reduction   degree   % = H 2   consumption   amount   800   ° C H 2   consumption   amount   1000   ° C × 100
H2-chemisorption data were obtained using an Autochem 2920 (Micromeritics, USA). Detailed procedures of H2-chemisorption were reported in our previous work [14]. CO2 temperature-programmed desorption (CO2-TPD) analyses were performed on an Autochem 2920 (Micromeritics, USA) instrument. In the pretreatment step, samples were fixed in a quartz U-tube and exposed with He at 300 °C for 10 min. The CO2 adsorption step was performed in 10% CO2/He for 30 min at 100 °C. Subsequently, physically adsorbed CO2 was removed by purging with He for 1 h, and the treated samples were heated in a He flow from 100 to 1000 °C at a rate of 10 °C/min. The desorbed CO2 was measured by a thermal conductivity detector (TCD).

3.3. Catalyst activity

The catalytic activity tests were performed under atmospheric pressure at 800 °C using a fixed-bed tubular quartz reactor with an inner diameter of 4 mm. The reactor temperature was monitored in real time and controlled by a temperature controller box. During the test procedure, 10 mg of the catalyst, 150 mg of the diluent, and a simulated coke oven gas (as the reactant gas: CH4: 28.04%, CO2: 2.56%, CO: 7.99%, H2: 55.57%, N2: 5.84%) were used, and CO2 was supplied to increase the CO2:CH4 ratio to 1.2:1. The catalyst was loaded into the tubular quartz reactor, and a catalytic test was conducted at 800 °C. A high GHSV of 1.5 million h−1 was applied for 12 h to screen the activity. Before each measurement, the sample was reduced in situ at 800 °C for 3 h under 5% H2/N2 conditions. Each gas was controlled using an independent mass flow controller (MFC). Effluent gases from the reactor were chilled and allowed to pass through a moisture trap to remove the remaining H2O. Then, the product gases were analyzed online using a micro-gas chromatograph (Agilent 3000, USA). The CH4, CO2 conversions, H2, CO yields, and H2/CO ratio were defined as follows:
C H 4   conversion   % = C H 4 in C H 4 out [ C H 4 ] in × 100
C O 2   conversion   % = C O 2 in C O 2 out [ C O 2 ] in × 100
H 2   yield   % = H 2 out H 2 in 2 × C H 4 in × 100
CO   yield   % = C O out C O in C H 4 in + C O 2 in × 100
H 2 / CO   ratio = H 2 out [ CO ] out
where [X]in and [X]out are the inlet and outlet concentrations of X, respectively. In the same way, Xin and Xout are the inlet and outlet amounts of X, respectively.

4. Conclusions

To confirm the effect of the MgO:Al2O3 ratio in the Ni/MG catalyst for the dry reforming of COG, Ni/MG catalysts were prepared with varying MgO:Al2O3 ratios (1:9, 3:7, 5:5, 7:3, 9:1). The Ni/MG catalyst prepared with a MgO:Al2O3 ratio of 3:7 (Ni/MG30) showed the highest CH4 and CO2 conversion (84.2%, 86.9%) even at high GHSV (1.5 million h−1) and was selected as the optimal production ratio.
In the case of the Ni/MG10 catalyst with the highest Al2O3 content in the support, the complex NiO reduction temperature was increased. In addition, it showed the formation of spinel NiAl2O4, with low dispersion and a large crystallite size. As the MgO content increased, the reduction temperature of complex NiO in Ni/MG30 catalyst decreased according to MgAl2O4 production. In the case of an excessive MgO content, a NiO-MgO solid solution, in which the excess MgO strongly interacted with NiO was formed, the NiO reduction temperature increased to a high temperature, the dispersion degree decreased, and the crystallite size increased. This resulted in the Ni/MG30 catalyst having the highest number of Ni0 active sites. As a result, the Ni/MG30 catalyst showed an appropriate correlation between complex NiO and the MgO-Al2O3 and showed the highest basicity. Also, due to the high reducibility and dispersibility of the catalyst surface, it has the largest number of Ni0 active sites. Consequently, the Ni/MG30 catalyst showed the highest catalytic activity between the prepared catalysts.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/catal11121468/s1. Figure S1: Adsorption/desorption isotherms of Ni/MG catalysts with different MgO:Al2O3 ratios. Table S1: The elemental composition of Ni/MG catalysts determined by ICP-OES.

Author Contributions

Conceptualization, formal analysis, H.-R.P., B.-J.K. and Y.-L.L.; writing—original draft preparation, visualization, validation, H.-R.P., methodology, S.-Y.A. and K.-J.K.; investigation, H.-R.P.; resources H.-R.P., B.-J.K. and Y.-L.L.; data curation, H.-R.P., B.-J.K., Y.-L.L., S.-Y.A., K.-J.K., G.-R.H. and S.-J.Y.; writing—review and editing, H.-R.P., B.-J.K., Y.-L.L., S.-Y.A., K.-J.K., G.-R.H., S.-J.Y., B.-H.J. and J.W.B.; supervision, H.-S.R.; project administration, H.-S.R.; funding acquisition, H.-S.R. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the “Next Generation Carbon Upcycling Project” (Project No. 2017M1A2A2044372) through the National Research Foundation (NRF) funded by the Ministry of Science and ICT, Korea.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Adoption of the Paris Agreement, L.9; UNFCCC: France, Paris, 2015; pp. 2–32.
  2. Rogelj, J.; den Elzen, M.; Höhne, N.; Fransen, T.; Fekete, H.; Winkler, H.; Schaeffer, R.; Sha, F.; Riahi, K.; Meinshausen, M. Paris agreement climate proposals need a boost to keep warming well below 2 degrees Celsius. Nature 2016, 534, 631–639. [Google Scholar] [CrossRef] [Green Version]
  3. Energy Technology Perspectives 2016: Towards Sustainable Urban Energy Systems; IEA: France, Paris, 2016; pp. 25–76.
  4. Umar, M.; Ji, X.; Kirikkaleli, D.; Alola, A.A. The imperativeness of environmental quality in the United States transportation sector amidst biomass-fossil energy consumption and growth. J. Clean. Prod. 2021, 285, 124863. [Google Scholar] [CrossRef]
  5. Al-mulali, U. Exploring the bi-directional long run relationship between energy consumption and life quality. Renew. Sustain. Energy Rev. 2016, 54, 824–837. [Google Scholar] [CrossRef]
  6. Friedmann, S.J.; Fan, Z.; Tang, K. Low-Carbon Heat Solutions for Heavy Industry: Sources, Options, and Costs Today; CGEP, Columbia University: New York, NY, USA, 2019. [Google Scholar]
  7. He, K.; Wang, L. A review of energy use and energy-efficient technologies for the iron and steel industry. Renew. Sustain. Energy Rev. 2017, 70, 1022–1039. [Google Scholar] [CrossRef]
  8. Tian, S.; Jiang, J.; Zhang, Z.; Manovic, V. Inherent potential of steelmaking to contribute to decarbonization targets via industrial carbon capture and storage. Nat. Commun. 2018, 9, 1–8. [Google Scholar] [CrossRef] [PubMed]
  9. Wang, P.; Ryberg, M.; Yang, Y.; Feng, K.; Kara, S.; Hauschild, M.; Chen, W.-Q. Efficiency stagnation in global steel production urges joint supply- and demand-side mitigation efforts. Nat. Commun. 2021, 12, 2066–2076. [Google Scholar] [CrossRef] [PubMed]
  10. Naito, M.; Takeda, K.; Matsui, Y. Ironmaking technology for the last 100 years: Deployment to advanced technologies from introduction of technological know-how, and evolution to next-generation process. ISIJ Int. 2015, 55, 7–35. [Google Scholar] [CrossRef] [Green Version]
  11. Uribe-Soto, W.; Portha, J.-F.; Commenge, J.-M.; Falk, L. A review of thermochemical processes and technologies to use steelworks off-gases. Renew. Sustain. Energy Rev. 2017, 74, 809–823. [Google Scholar] [CrossRef]
  12. Park, J.E.; Koo, K.Y.; Jung, U.H.; Lee, J.H.; Roh, H.-S.; Yoon, W.L. Syngas production by combined steam and CO2 reforming of coke oven gas over highly sinter-stable La-promoted Ni/MgAl2O4 catalyst. Int. J. Hydrogen Energy 2015, 40, 13909–13917. [Google Scholar] [CrossRef]
  13. Aghaali, M.H.; Firoozi, S. Enhancing the catalytic performance of Co substituted NiAl2O4 spinel by ultrasonic spray pyrolysis method for steam and dry reforming of methane. Int. J. Hydrogen Energy 2021, 46, 357–373. [Google Scholar] [CrossRef]
  14. Lee, Y.-L.; Kim, B.-J.; Park, H.-R.; Ahn, S.-Y.; Kim, K.-J.; Roh, H.-S. Customized Ni–MgO–Al2O3 catalyst for carbon dioxide reforming of coke oven gas: Optimization of preparation method and co-precipitation pH. J. CO2 Util. 2020, 42, 101354–101362. [Google Scholar] [CrossRef]
  15. Li, Z.; Jiang, B.; Wang, Z.; Kawi, S. High carbon resistant Ni@Ni phyllosilicate@SiO2 core shell hollow sphere catalysts for low temperature CH4 dry reforming. J. CO2 Util. 2018, 27, 238–246. [Google Scholar] [CrossRef]
  16. Karam, L.; Miglio, A.; Specchia, S.; Hassan, N.E.; Massiani, P.; Reboul, J. PET waste as organic linker source for the sustainable preparation of MOF-derived methane dry reforming catalysts. Mater. Adv. 2021, 2, 2750–2758. [Google Scholar] [CrossRef]
  17. Liu, W.; Li, L.; Zhang, X.; Wang, Z.; Wang, X.; Peng, H. Design of Ni-ZrO2@SiO2 catalyst with ultra-high sintering and coking resistance for dry reforming of methane to prepare syngas. J. CO2 Util. 2018, 27, 297–307. [Google Scholar] [CrossRef]
  18. Jang, W.J.; Shim, J.O.; Kim, H.M.; Yoo, S.Y.; Roh, H.S. A review on dry reforming of methane in aspect of catalytic properties. Catal Today 2019, 324, 15–26. [Google Scholar] [CrossRef]
  19. Li, X.; Li, D.; Tian, H.; Zeng, L.; Zhao, Z.J.; Gong, J. Dry reforming of methane over Ni/La2O3 nanorod catalysts with stabilized Ni nanoparticles. Appl. Catal. B Environ. 2017, 202, 683–694. [Google Scholar] [CrossRef]
  20. Li, G.; Cheng, H.; Zhao, H.; Lu, X.; Xu, Q.; Wu, C. Hydrogen production by CO2 reforming of CH4 in coke oven gas over Ni–Co/MgAl2O4 catalysts. Catal Today 2018, 318, 46–51. [Google Scholar] [CrossRef]
  21. Jang, W.J.; Kim, H.M.; Shim, J.O.; Yoo, S.Y.; Jeon, K.W.; Na, H.S.; Lee, Y.L.; Jeong, D.W.; Bae, J.W.; Nah, I.W.; et al. Key properties of Ni–MgO–CeO2, Ni–MgO–ZrO2, and Ni–MgO–Ce(1−x)Zr(x)O2 catalysts for the reforming of methane with carbon dioxide. Green Chem. 2018, 20, 1621–1633. [Google Scholar] [CrossRef]
  22. Ren, H.-P.; Ding, S.-Y.; Ma, Q.; Song, W.-Q.; Zhao, Y.-Z.; Liu, J.; He, Y.-M.; Tian, S.-P. The Effect of Preparation Method of Ni-Supported SiO2 Catalysts for Carbon Dioxide Reforming of Methane. Catalysts 2021, 11, 1221. [Google Scholar] [CrossRef]
  23. Talkhoncheh, S.K.; Haghighi, M. Syngas production via dry reforming of methane over Ni-based nanocatalyst over various supports of clinoptilolite, ceria and alumina. J. Nat. Gas Sci. Eng. 2015, 23, 16–25. [Google Scholar] [CrossRef]
  24. Mozammel, T.; Dumbre, D.; Hubesch, R.; Yadav, G.D.; Selvakannan, P.R.; Bhargava, S.K. Carbon Dioxide Reforming of Methane over Mesoporous Alumina Supported Ni(Co), Ni(Rh) Bimetallic, and Ni(CoRh) Trimetallic Catalysts: Role of Nanoalloying in Imoriving the Stability and Nature of Coking. Energy Fuels 2020, 34, 16433–16444. [Google Scholar] [CrossRef]
  25. Jang, W.J.; Jeong, D.W.; Shim, J.O.; Kim, H.M.; Han, W.B.; Bae, J.W.; Roh, H.S. Metal oxide (MgO, CaO, and La2O3) promoted Ni-Ce0.8Zr0.2O2 catalysts for H2 and CO production from two major greenhouse gases. Renew. Energy 2015, 79, 91–95. [Google Scholar] [CrossRef]
  26. Han, J.W.; Park, J.S.; Choi, M.S.; Lee, H.J. Uncoupling the size and support effects of Ni catalysts for dry reforming of methane. Appl. Catal. B Environ. 2017, 203, 625–632. [Google Scholar] [CrossRef]
  27. Margossian, T.; Larmier, K.; Kim, S.M.; Krumeich, F.; Fedorov, A.; Chen, P.; Müller, C.R.; Copéret, C. Molucularly Tailored Nickel Precursor and Support Yield a Stable Methane Dry Reforming Catalyst with Superior Metal Utilization. J. Am. Chem. Soc. 2017, 139, 6919–6927. [Google Scholar] [CrossRef] [PubMed]
  28. Shen, D.; Huo, M.; Li, L.; Lyu, S.; Wang, J.; Wang, X.; Zhang, Y.; Li, J. Effect of alumina morphology on dry reforming of methane over Ni/Al2O3 catalysts. Catal. Sci. Technol. 2020, 10, 510–516. [Google Scholar] [CrossRef]
  29. Shamskar, F.R.; Rezaei, M.; Meshkani, F. The influence of Ni loading on the activity and coke formation of ultrasound-assisted co-precipitated Ni–Al2O3 nanocatalyst in dry reforming of methane. Int. J. Hydrogen Energy 2017, 42, 4155–4164. [Google Scholar] [CrossRef] [Green Version]
  30. Hou, Z.; Yokota, O.; Tanaka, T.; Yashima, T. Characterization of Ca-promoted Ni/-Al2O3 catalyst for CH4 reforming with CO2. Appl. Catal. A Gen. 2003, 253, 381–387. [Google Scholar] [CrossRef]
  31. Alipoura, Z.; Rezaei, M.; Meshkani, F. Effect of Ni loadings on the activity and coke formation of MgO-modified Ni/Al2O3 nanocatalyst in dry reforming of methane. J. Energy Chem. 2014, 23, 633–638. [Google Scholar] [CrossRef]
  32. Sengupta, S.; Deo, G. Modifying alumina with CaO or MgO in supported Ni and Ni–Co catalysts and its effect on dry reforming of CH4. J. CO2 Util. 2015, 10, 67–77. [Google Scholar] [CrossRef]
  33. Guo, J.; Lou, H.; Zheng, X. The deposition of coke from methane on a Ni/MgAl2O4 catalyst. Carbon 2007, 45, 1314–1321. [Google Scholar] [CrossRef]
  34. Zhang, Q.L.; Qiu, C.T.; Xu, H.D.; Lin, T.; Lin, Z.E.; Gong, M.C. Low-temperature selective catalytic reduction of NO with NH3 over monolith catalyst of MnOx/CeO2-ZrO2-Al2O3. Catal. Today 2011, 175, 171–176. [Google Scholar] [CrossRef]
  35. Ko, J.H.; Park, S.H.; Jeon, J.K.; Kim, S.S.; Kim, S.C.; Kim, J.M.; Chang, D.; Park, Y.K. Low temperature selective catalytic reduction of NO with NH3 over Mn supported on Ce0.65Zr0.35O2 prepared by supercritical method: Effect of Mn precursors on NO reduction. Catal. Today 2012, 185, 290–295. [Google Scholar] [CrossRef]
  36. Shen, B.; Zhang, X.; Ma, H.; Yao, Y.; Liu, T. A comparative study of Mn/CeO2, Mn/ZrO2 and Mn/Ce-ZrO2 for low temperature selective catalytic reduction of NO with NH3 in the presence of SO2 and H2O. J. Environ. Sci. 2013, 25, 791–800. [Google Scholar] [CrossRef]
  37. Rynkowski, J.M.; Paryjczak, T.; Lenik, M. On the nature of oxidic nickel phases in NiO/y-Al2O3 catalysts. Appl. Catal. A 1993, 106, 73–82. [Google Scholar] [CrossRef]
  38. Salagre, P.; Fierro, J.L.G.; Medina, F.; Sueiras, J.E. Characterization of nickel species on several y-alumina supported nickel sample. J. Mol. Catal. A 1996, 106, 125–134. [Google Scholar] [CrossRef]
  39. Turlier, P.; Praliaud, H.; Moral, P.; Martin, G.A.; Dalmon, J.A. Influence of the nature of the support on the reducibility and catalytic properties of nickel: Evidence for a new type of metal support interaction. Appl. Catal. 1985, 19, 287–300. [Google Scholar] [CrossRef]
  40. Xu, D.; Li, W.; Ge, Q.; Xu, H. A novel process for converting coalmine-drained methane gas to syngas over nickel–magnesia solid solution catalyst. Fuel Process. Technol. 2005, 86, 995–1006. [Google Scholar] [CrossRef]
  41. Arena, F.; Frusteri, F.; Giordano, N. Temperature-programmed Reduction Study of NiO-MgO Interactions in Magnesia-supported Ni Catalysts and NiO-MgO Physical Mixture. J. Chem. Soc. Faraday Trans. 1990, 86, 2663–2669. [Google Scholar] [CrossRef]
  42. Wang, Y.; Ohtsuka, Y. Mn-based binary oxides as catalysts for the conversion of methane to C2 hydrocarbons with carbon dioxide as oxidant. Appl. Catal. A 2001, 219, 183–193. [Google Scholar] [CrossRef]
  43. Zangeneh, F.T.; Sahebdelfar, S.; Ravanchi, T.M. Conversion of carbon dioxide to valuable petrochemicals: An approach to clean development mechanism. J. Nat. Gas Chem. 2011, 20, 219–231. [Google Scholar] [CrossRef]
  44. Roh, H.-S.; Potdar, H.S.; Jun, K.-W.; Kim, J.-W.; Oh, Y.-S. Carbon dioxide reforming of methane over Ni incorporated into Ce–ZrO2 catalysts. Appl. Catal. A 2004, 276, 231–239. [Google Scholar] [CrossRef]
Scheme 1. Schematic of the preparation, characterization, and reaction conditions of the Ni/MG catalysts.
Scheme 1. Schematic of the preparation, characterization, and reaction conditions of the Ni/MG catalysts.
Catalysts 11 01468 sch001
Figure 1. XRD results of Ni/MG catalysts with different MgO/Al2O3 ratios.
Figure 1. XRD results of Ni/MG catalysts with different MgO/Al2O3 ratios.
Catalysts 11 01468 g001
Figure 2. TPR results of Ni/MG catalysts with different MgO/Al2O3 ratios.
Figure 2. TPR results of Ni/MG catalysts with different MgO/Al2O3 ratios.
Catalysts 11 01468 g002
Figure 3. CO2-TPD results of Ni/MG catalysts with different MgO:Al2O3 ratios.
Figure 3. CO2-TPD results of Ni/MG catalysts with different MgO:Al2O3 ratios.
Catalysts 11 01468 g003
Figure 4. (A) TEM and (B) EDS mapping images of Ni/MG30 catalyst.
Figure 4. (A) TEM and (B) EDS mapping images of Ni/MG30 catalyst.
Catalysts 11 01468 g004
Figure 5. The relationships among the physicochemical properties of the catalysts and their catalytic performance.
Figure 5. The relationships among the physicochemical properties of the catalysts and their catalytic performance.
Catalysts 11 01468 g005
Figure 6. CH4 conversion vs. time on stream for Ni/MG catalysts with diverse MgO:Al2O3 ratios (CO2/CH4 = 1.2; reaction temperature = 800 °C; GHSV =1,500,000 h−1).
Figure 6. CH4 conversion vs. time on stream for Ni/MG catalysts with diverse MgO:Al2O3 ratios (CO2/CH4 = 1.2; reaction temperature = 800 °C; GHSV =1,500,000 h−1).
Catalysts 11 01468 g006
Figure 7. CO2 conversion vs. time on stream over Ni/MG catalysts with diverse MgO:Al2O3 ratios (CO2/CH4 = 1.2; reaction temperature = 800 °C; GHSV = 1,500,000 h−1).
Figure 7. CO2 conversion vs. time on stream over Ni/MG catalysts with diverse MgO:Al2O3 ratios (CO2/CH4 = 1.2; reaction temperature = 800 °C; GHSV = 1,500,000 h−1).
Catalysts 11 01468 g007
Table 1. Characteristics of Ni/MG catalysts with different MgO:Al2O3 ratios.
Table 1. Characteristics of Ni/MG catalysts with different MgO:Al2O3 ratios.
CatalystSupport BET
S.A. (m2/g) 1
Catalyst BET
S.A. (m2/g) 1
Ni Dispersion
(%) 2
Crystallite Size of Ni0 (nm) 3
Ni/MG10202.6139.21.4613.6
Ni/MG30200.9120.83.419.4
Ni/MG50164.2104.23.329.7
Ni/MG70161.5104.03.189.9
Ni/MG90102.076.63.2411.9
1 Predicted from N2 adsorption–desorption at −196 °C. 2 Predicted from H2-chemisorption contemplating the Ni reduction degree. 3 Calculated from the Scherrer equation based on the XRD result.
Table 2. Reduction degree results of Ni/MG catalysts with different MgO/Al2O3 ratios.
Table 2. Reduction degree results of Ni/MG catalysts with different MgO/Al2O3 ratios.
CatalystNi/MG10Ni/MG30Ni/MG50Ni/MG70Ni/MG90
Reduction
degree (%) 1
8496886751
1 Calculated from (H2 consumption amount (800 °C)/H2 consumption amount (1000 °C)) × 100.
Table 3. CO2-TPD results of MG supports and Ni/MG catalysts with different MgO:Al2O3 ratios.
Table 3. CO2-TPD results of MG supports and Ni/MG catalysts with different MgO:Al2O3 ratios.
CatalystSupport Desorbed CO2
(cm3/gcat)
Total Desorbed CO2
(cm3/gcat)
Ni/MG1018.06.1
Ni/MG3058.89.7
Ni/MG5071.28.2
Ni/MG70154.38.0
Ni/MG9039.87.7
Table 4. CO2 reforming of CH4 using COG over Ni/MG30 catalyst (reaction results for 12 h).
Table 4. CO2 reforming of CH4 using COG over Ni/MG30 catalyst (reaction results for 12 h).
Gas Composition (%) 1
ComponentCH4CO2N2H2COTotal
Reactant gas (COG) + CO221.3925.674.4642.396.10100.00
Product gas2.652.683.5355.2735.87100.00
Yield (%)Conversion (%)H2/CO ratio
H2COCH4CO2Reactant gas (COG) + CO2Product gas
64.083.284.486.86.951.54
1 Evaluated by micro-GC.
Table 5. Ni/MgO-Al2O3 catalysts’ naming convention.
Table 5. Ni/MgO-Al2O3 catalysts’ naming convention.
MgO:Al2O3Name
1:9Ni/MG10
3:7Ni/MG30
5:5Ni/MG50
7:3Ni/MG70
9:1Ni/MG90
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Park, H.-R.; Kim, B.-J.; Lee, Y.-L.; Ahn, S.-Y.; Kim, K.-J.; Hong, G.-R.; Yun, S.-J.; Jeon, B.-H.; Bae, J.W.; Roh, H.-S. CO2 Reforming of CH4 Using Coke Oven Gas over Ni/MgO-Al2O3 Catalysts: Effect of the MgO:Al2O3 Ratio. Catalysts 2021, 11, 1468. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11121468

AMA Style

Park H-R, Kim B-J, Lee Y-L, Ahn S-Y, Kim K-J, Hong G-R, Yun S-J, Jeon B-H, Bae JW, Roh H-S. CO2 Reforming of CH4 Using Coke Oven Gas over Ni/MgO-Al2O3 Catalysts: Effect of the MgO:Al2O3 Ratio. Catalysts. 2021; 11(12):1468. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11121468

Chicago/Turabian Style

Park, Ho-Ryong, Beom-Jun Kim, Yeol-Lim Lee, Seon-Yong Ahn, Kyoung-Jin Kim, Ga-Ram Hong, Seong-Jin Yun, Byong-Hun Jeon, Jong Wook Bae, and Hyun-Seog Roh. 2021. "CO2 Reforming of CH4 Using Coke Oven Gas over Ni/MgO-Al2O3 Catalysts: Effect of the MgO:Al2O3 Ratio" Catalysts 11, no. 12: 1468. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11121468

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