Next Article in Journal
Impact of Doping and Additive Applications on Photocatalyst Textural Properties in Removing Organic Pollutants: A Review
Next Article in Special Issue
Methanation of CO2 Using MIL-53-Based Catalysts: Ni/MIL-53–Al2O3 versus Ni/MIL-53
Previous Article in Journal
Immobilization of Fusarium solani Cutinase onto Magnetic Genipin-Crosslinked Chitosan Beads
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

CO2 Hydrogenation to Synthetic Natural Gas over Ni, Fe and Co–Based CeO2–Cr2O3

by
Chalempol Khajonvittayakul
1,
Vut Tongnan
1,
Suksun Amornraksa
1,
Navadol Laosiripojana
2,
Matthew Hartley
3 and
Unalome Wetwatana Hartley
1,2,*
1
Chemical and Process Engineering, The Sirindhorn International Thai-German Graduate School of Engineering (TGGS), King Mongkut’s University of Technology North Bangkok, Bangkok 10800, Thailand
2
Joint Graduate School of Energy and Environment (JGSEE), King Mongkut’s University of Technology Thonburi, Bangkok 10140, Thailand
3
Department of Chemical Engineering, Faculty of Engineering, King Mongkut’s University of Technology North Bangkok, Bangkok 10800, Thailand
*
Author to whom correspondence should be addressed.
Catalysts 2021, 11(10), 1159; https://doi.org/10.3390/catal11101159
Submission received: 29 August 2021 / Revised: 22 September 2021 / Accepted: 22 September 2021 / Published: 26 September 2021
(This article belongs to the Special Issue Frontiers in Catalysis for CO2 Methanation)

Abstract

:
CO2 methanation was studied over monometallic catalyst, i.e., Ni, Fe and Co; on CeO2-Cr2O3 support. The catalysts were prepared using one-pot hydrolysis of mixed metal nitrates and ammonium carbonate. Physicochemical properties of the pre- and post-exposure catalysts were characterized by X-Ray Powder Diffraction (XRD), Hydrogen Temperature Programmed Reduction (H2-TPR), and Field Emission Scanning Electron Microscope (FE-SEM). The screening of three dopants over CeO2-Cr2O3 for CO2 methanation was conducted in a milli-packed bed reactor. Ni-based catalyst was proven to be the most effective catalyst among all. Thus, a group of NiO/CeO2-Cr2O3 catalysts with Ni loading was investigated further. 40 % NiO/CeO2-Cr2O3 exhibited the highest CO2 conversion of 97.67% and CH4 selectivity of 100% at 290 °C. The catalytic stability of NiO/CeO2-Cr2O3 was tested towards the CO2 methanation reaction over 50 h of time-on-stream experiment, showing a good stability in term of catalytic activity.

1. Introduction

Global warming has caused several serious impacts on the environment in recent years. Increasing CO2 emission is anthropogenic in origin and is the main cause of global warming. Nowadays, many studies focused on two strategies to reduce atmospheric CO2 concentration; through carbon capture and CO2 conversion to biofuels [1,2]. The captured CO2 can be utilized and converted into fuels and chemicals via chemical processes such as dry reforming of methane for synthesis gas production, or CO2 hydrogenation to CH4, methanol or higher alcohols [3]. CO2 methanation is one of the promising processes which involves carbon recycle from abundant CO2. Methane, as a product of CO2 hydrogenation, is considered versatile and flexible as it can be injected directly into existing natural gas pipelines, or utilized as a raw material for chemical production [4]. This CO2 hydrogenation can be looked at as Power-to-Gas process (PtG) by its means to store (and transport) energy in the form of natural gas [5]. The process refers to a conversion of renewable electricity to a gaseous energy carrier via two pathways: (1) H2 production by water electrolysis, where wind or solar energy technologies could be integrated; and (2) H2 conversion to CH4, by methanation reaction with external CO2 capture [6]. CO2 methanation was firstly discovered and proposed as the Sabatier reaction: CO2 + 4H2 Catalysts 11 01159 i001 CH4 + 2H2O, ∆Hr298 = −164.8 kJ·mol−1 [7]. Although the reaction is highly exothermic and thermodynamically favored at high pressures/low temperatures (<400 °C) [4,8], there are significant kinetic limitations due to the high stability of CO2. Furthermore, heat accumulation from the reaction generally causes severe hotspots in the reactor, due to the heat transfer limitation within the process, leading to the catalyst deactivation and shortened catalyst lifespan [9]. Moreover, low operating temperature is favorable for CO disproportionation reaction (2CO Catalysts 11 01159 i001 CO2 + C, ∆Hr298 = −172.4 kJ·mol−1), resulting in unwanted coke deposition. In order to obtain the highest possible methane yield, it is necessary to invent a catalyst which enhances the reaction’s activity, withstands sintering and counters the coking phenomenon. Various active metals (such as Ni, Fe, Co, Ru, Rh, and Pd) have been used as an active site while metal oxides (such as CeO2, La2O3, MgO, γ-Al2O3, SiO2, TiO2, and ZrO2) have been useful as a support in a catalyst system for the CO2 methanation reaction [3,10,11,12,13]. Amongst these materials, CeO2 is so far found to be the most interesting support due to its high oxygen storage capacity (OSC) and its ability to disperse the active site [14]. In addition, CeO2 could promote the interaction between support and metal active component, such that the growth and dispersion of the metal active particles can be well distributed and controlled throughout the surface of the support, leading to the higher CO2 conversion [15]. The number oxygen vacancy can be tailor-made by substituting smaller transition metal ions (e.g., chromium ions) into CeO2. The higher number of lattice oxygen can combust coke deposits and reduce the chance of sintering [13,16,17,18]. According to previous research, Ni-, Fe-, Co doped on CeO2 have shown relatively high activities for CO2 methanation and possessed high stability when tested for 15 to 50 h reaction times [19,20,21,22,23].
In this work, Ni-, Fe- and Co- based CeO2/Cr2O3 were prepared using the one-pot hydrolysis method. The level of metal loading, operating temperature, reduction temperature and other relevant variables were observed as all of these parameters are well-known to influence the catalytic performance of the catalysts [24,25]. The physicochemical properties of the synthesized catalysts were examined, comparing pre- and post-exposure by X-Ray Diffractometer (XRD), Hydrogen Temperature-Programmed Reduction (H2-TPR), and Field Emission Scanning Electron Microscopes (FE-SEM). The catalyzation of CO2 methanation was conducted in a milli-packed bed reactor under atmospheric pressure where the operating temperature was varied from 200 to 350 °C. The reduction temperatures of 500 and 700 °C were chosen (via H2-TPR) for comparison purposes. WSHV was fixed at 27,624 mL·h1·gcat−1, and the stoichiometric reactants ratio was kept at 4 for all the experiments.

2. Methodology

2.1. Catalyst Powder-Formed Preparation

Forty percent (by weight) x/CeO2-Cr2O3 (where x = Ni, Fe, and Co) catalysts were synthesized by a single step preparation using (NH4)2CO3 (PANREAC, 30% NH3) as a hydrolysis agent, the details of which are outlined in [3]. The relevant nitrate precursors Ni(NO3)2∙6H2O (CARLO ERBA, Cornaredo, Italy, ≥99.0%), Fe(NO3)3·9H2O (UNIVAR, Donners Grove, IL, UAS, ≥99.0%), Co(NO3)2∙6H2O (CARLO ERBA, ≥99.0%), Ce(NO3)3·6H2O (ALDRICH, St. Louis, MO, USA, ≥99.0%), and Cr(NO3)2·6H2O (ACROS, Merelbeke, Belgium, ≥99.0%) were dissolved in 50 mL distilled water where the ratio of active metal (Ni, Fe, and Co) to support (1 to 1 of CeO2/Cr2O3) was fixed at 40 to 60 by weight. Two molar (NH4)2CO3 solution was gradually dropped into the nitrate solutions until the pH reached 8.8–9.0. The mixture was continuously stirred while heated to 80 °C for 3 h. The solution’s temperature was then raised again to 120 °C to evaporate water and a dark blue gel was slowly obtained. The resulting material was then calcined in moving air at 500 °C with 10 °C/min of heating for 24 h before the black powder of the catalyst was achieved. The catalyst powder was then pressed, crushed, and sieved to gain its uniform particle size ranging from 75 to 180 µm in order to avoid pressure drop that could occur across the catalyst bed.

2.2. Characterizations

XRD analysis (Malvern PANalytical diffractometer)was performed using CuKα radiation (with λ = 1.5418 Å, 40 kV, 15 mA). The diffractogram patterns were recorded over 2-theta ranging from 10 to 80° with a scanning speed of 0.02° per second. The catalyst’s phase structures were identified using JCPDS cards (Joint Committee on Powder Diffraction Standards).
The optimal reduction temperature of the catalyst was screened using an in situ H2-TPR technique which was carried out in our lab-scale conventional packed-bed reactor, connected to a Quadrupole Mass Spectrometer (PFEIFFER, MS, Omnistar GSD 320, HAKUTO) operated in a SEM-MID mode. A total of 0.5 g of the catalyst sample was packed in a quartz tube reactor (i.d. = 10 mm) and pre-treated in 10% O2/Ar at 500 °C for 1 h, followed by Ar purging to clean the catalyst’s surface from any possible absorbed impurities. After the system reached ambient temperature, 5 % H2/Ar was introduced through the catalyst’s bed with a total flowrate of 100 mL·min−1 while the temperature was elevated to 950 °C at 5 °C/min.
Surface morphology and micro-structure of the catalysts, both pre- and post-exposure, were investigated using a Field Emission Scanning Electron Microscopes (FE-SEM, SU-8230 Hitachi, Japan) with an accelerating voltage of 15 kV.

2.3. CO2 Methanation Activity in a Packed-Bed Reactor

CO2 methanation was performed in a tubular packed-bed reactor under atmospheric pressure. A total of 0.2 g of catalyst was placed between two layers of quartz wool in the middle of the reactor (i.d. = 4 mm). The catalyst was reduced in 100 mL·min−1 of pure H2 for 2 h at the achieved reduction temperature (from H2-TPR where NiO reduced to metallic Ni at 500 °C while Co2O3 and Fe3O4 reduced to metallic Co and Fe at 850 °C) from the prior reaction. Next, the process was cooled down to the desired operating temperature, varying at 200, 210, 230, 250, 270, 290, 310, and 350 °C. Ar was purged in between to remove any excess H2. The mixture of gaseous reactant, CO2:H2:Ar at a ratio of 1:4:5 by volume, was injected through the catalyst’s bed. Total flow rate was set at 90 mL·min−1, giving WSHV at 27,624 mL·h−1·gcat−1. Moisture was condensed as a by-product using a cooler oil bath at the bottom of the reactor. After the process approached equilibrium, the dried gas products were automatically analyzed using gas chromatography coupled with a TCD detector (Shimadzu GC-2014ATF) every 7 min for 1 h. CO2 conversion (XCO2), CH4 selectivity (SCH4), and CH4 yield (YCH4) were calculated using the following formulas:
C O 2   c o n v e r s i o n ,   X C O 2   [ % ] = ( F C O 2 i n F C O 2 o u t F C O 2 i n ) × 100 %
C H 4   s e l e c t i v i t y ,   S C H 4   [ % ] = ( F C H 4 F C H 4 + F C O ) × 100 %
C H 4   y i e l d ,   Y C H 4   [ % ] = X C O 2 × S C H 4   100
F C O 2 i n and F C O 2 o u t represent volumetric flow rate of CO2 in the feed stream and outlet stream, respectively, whereas F C H 4 and F C O denote the volumetric flow rate of the product gas stream, CH4 and CO, respectively.

3. Results and Discussion

3.1. Characterizations

3.1.1. XRD

XRD patterns of all the fresh catalysts (calcined in moving air at 500 °C) were achieved as shown in Figure 1. CeO2-Cr2O3 (■), as major crystals, were found in all samples and appeared to possess fluorite cubic structure [26], having two-theta position peaks at 28.57, 33.12, 47.49, 56.38, 58.94, 69.42, 76.74, 79.33, 88.44, and 95.34°. All the dopants, NiO (●), Fe2O3 (♦), and Co3O4 (▲) appeared as minor crystalline phases as shown in Figure 1a–c, respectively. NiO peaks were found at 37.31, 43.35, 63.00, and 75.49° (JCPDS No. 01-073-1519); whereas Fe2O3 peaks appeared at 24.19, 33.28, 35.68, 40.99, 49.57, 54.32, 62.64, and 64.15 (JCPDS No. 01-076-4579); and Co3O4 peaks were detected at 33.06, 36.67, 44.59, 59.08, and 64.92 (JCPDS No. 01-078-5631). Pure phase Cr2O3 was found in Co3O4-CeO2-Cr2O3 at 36.34, 44.60, 58.357, and 63.204° (JCPDS No. 00-001-1294), indicating that Cr2O3 cannot fully incorporate into the CeO2 lattice. This inhomogeneous solid solution depends on the size of ionic radii of the solutes. The ionic radius of Ce3+ appeared the largest (1.101 Å), followed by Cr3+ (0.80 Å) and Co4+ (0.61 Å) [27,28]. Thus, the Co4+ ion was able to compete with Cr3+ in becoming embedded into the CeO2 lattice, creating CeCoO3 perovskite [27,29], as it can be seen in Figure 1c. This phenomenon could cause a decay in the catalyst’s catalytic performance due to the loss of active sites, in this case, Co3O4. In addition, the average crystallite size of the NiO, Fe2O3, and Co3O4 on CeO2-Cr2O3 were calculated using the Scherrer’s equation at 14.56, 25.03, and 25.60 nm, respectively. The smaller active site could perhaps accommodate reactants better, rendering the chance of higher catalytic performance.

3.1.2. H2-TPR

Figure 2 shows the reduction profiles of pure CeO2 (a), CeO2-Cr2O3 (b), NiO/CeO2-Cr2O3 (c), Fe2O3/CeO2-Cr2O3 (d), and Co3O4/CeO2-Cr2O3 (e) catalysts. Pure CeO2 (a) had 2 small reduction peaks at 570 and >950 °C, corresponding to the surface reduction and bulk reduction of CeO2, respectively [30]. The first reduction peak of CeO2-Cr2O3 (b) appeared at 480 °C, where Cr6+ ions were reduced to Cr3+ ions. The reduction peak at 565 °C and >950 °C corresponded to the reduction of CeO2-Cr2O3 at surface and bulk oxygen, respectively [30,31,32,33]. Substitution of Cr2O3 into CeO2 was reported to enhance oxygen vacancy of the catalyst system [16,18,34], in which its H2 consumption was proven to be significantly higher than the pure CeO2.
Three distinct peaks at 325, 675, and 940 °C were found for NiO/CeO2-Cr2O3 catalyst (c). The first two peaks were identical to the reduction of Ni3+ to Ni2+ and to the reduction of Ni2+ to metallic nickel, respectively [34,35,36]. Some reduction of Cr6+ ions to Cr3+ ions could be combined in the first peak, whereas the second and the third peaks represented the reduction of the Cr2O3 incorporated within the CeO2 structure at the surface and bulk level, respectively. For Fe2O3/CeO2-Cr2O3 (d), the first peak appeared at 395 °C, representing the reduction of Cr6+ ions to Cr3+ ions, whereas its second and third peaks at 505 and 940 °C were attributed to the reduction of CeO2-Cr2O3 at the surface and bulk levels, respectively. The two reduction peaks observed at 505 °C and between 700 to 950 °C also represented the reduction of Fe2O3 to Fe3O4 and reduction of Fe3O4→FeO→metallic Fe, respectively [37,38]. Co2O3/CeO2-Cr2O3 (e) was detected at 445, 700, and >950 °C, and associated with (1) the reduction of Cr6+ ions to Cr3+ ions, (2) the reduction of Co3+ ions to Co2+ ions and the reduction of CeO2-Cr2O3 (and/or CeCoO3 perovskite) with surface oxygen, and (3) the reduction of Co2+ ions to metallic Co, and the reduction of CeO2-Cr2O3 (and/or CeCoO3 perovskite) with bulk oxygen [39]. The catalyst’s oxygen deficiency and number of active sites were interpreted from the hydrogen consumption, which was compared amongst all the catalysts and ordered as: NiO/CeO2-Cr2O3 > Fe2O3/CeO2-Cr2O3 > Co2O3/CeO2-Cr2O3.

3.2. Catalytic Performance Test

3.2.1. Choice of the Monometallic

Catalytic performance, in terms of CO2 conversion (Figure 3 (left)) and CH4 selectivity (Figure 3 (right)), of all the prepared catalysts was determined at various operating temperatures, ranging from 200 to 350 °C. CO2 conversion tended to increase with increasing temperature for all catalysts. Amongst all the selected metals, Ni was proven as the best monometallic active site for CeO2-Cr2O3, considering CO2 conversion, which was much higher than other metals (Fe and Co) starting at 260 °C. The highest CO2 conversion over Ni/CeO2-Cr2O3 was achieved at 290 °C, giving CO2 conversion of 90.19%. However, CO2 conversion decreased when the temperature was higher than 330 °C, due to its thermodynamic limitation [3,40,41]. In terms of CH4 selectivity, Ni also showed the best performance by giving complete selectivity at 100% during all temperatures (from 200 to 360 °C), followed by Fe which offered 94% of CH4 selectivity at its equilibrium at 290 °C. On the other hand, catalytic performance of Co as the monometallic dopant was incomparable to that of the other two, as it gave no reaction at low temperature (below 260 °C) and reached its maximum at 24% of CH4 selectivity at 270 °C. The CH4 selectivity was decreased at temperatures higher than 270 °C. This was due to the formation of CO as an unwanted product from the reverse water-gas shift reaction [19,42].

3.2.2. Effect of Metal Content on Catalytic Performance

The influence of metal content, doped on CeO2-Cr2O3, towards CO2 methanation was investigated over Ni-/CeO2-Cr2O3, where the Ni level was varied at 10, 20, 30, 40 and 50% by weight. Figure 4 showed that the percentage of all the selected Ni contents exhibited the same trend, where CO2 conversion was increased with increasing temperature and increasing amount of Ni content. The nickel content represented the amount of the active site for CO2 methanation reaction, thus, the higher level of Ni was unsurprisingly improved the efficiency of the reaction [41,43,44,45]. However, excess Ni loading could cause other problems, i.e., pore blockage, coagulation and obstruction of nano-channels [46,47,48]. For this reason, there was only a small difference in CO2 conversion, between using 40% and 50% Ni loading.

3.2.3. Effect of Reduction Temperature on Catalytic Performance

Two different reduction temperatures, at 500 and 700 °C, were selected for this study. Figure 5 presents relationship between CO2 conversion and reduction temperature of the catalyst at different operating temperatures. The results showed that the catalyst which reduced at 500 °C gave the highest CO2 conversion for all of the temperature ranges, compared to the one reduced at 700 °C. In addition, the decrease in CO2 conversion at the higher reduction temperature (700 °C) could also be the effect of the catalyst’s sintering, resulting in a lower number of active sites [49].

3.3. Catalytic Stability

The catalytic stability of the NiO/CeO2-Cr2O3 was measured in term of CO2 conversion and CH4 selectivity, illustrated in Figure 6. Approximately 97% of CO2 conversion and 100% of CH4 selectivity were achieved and maintained during 50 h of reaction time. XRD and SEM techniques were utilized for pre- and post-exposure characterization. XRD patterns of fresh NiO/CeO2-Cr2O3 catalyst was compared with the post-exposure one after the stability test, shown in Figure 7. Although both look quite similar, a decrease in full-width half-maximum (FWHM) was clearly noticed, indicating catalyst sintering. Compared to pre-exposure, post-exposure crystallite size was found to have increased from 11 to 13 nm, whereas particle size was doubled from 313 to 612 nm. However, the sign of sintering or deactivation was not clearly observed in TOS experiment. This could be due to the fact that the rate of reaction is rapid, to the point that the catalyst surface area becomes relatively less significant. No NiO peak was found on the diffraction pattern in either the pre- or post-exposure, indicating that the catalyst was fully reduced as peaks appeared at 44.508, 51.847, and 76.372 (JCPDS No. 00-004-0850). However, surface morphology of the pre- and post-exposure catalyst were found to be different, as shown in Figure 8. It can be seen that the particle size of the catalyst became larger after reaction due to its agglomeration, in an attempt to reduce its surface free energy.
The catalyst performance was compared between this work and other works that researched other catalysts (i.e., 10Ni/CeO2 [42], 10Ni/CeO2-ZrO2 (CZ) [19], 15Ni/CZ, 15Ni-3Fe/CZ, 15Ni-3Co/CZ [42], 40Ni/CZ [3], 15Ni-2Ce/Al2O3 [50], 5Ni/CZ [51], 20Ni/Al2O3 [52], and 40Ni-5Ce/Al2O3 [47]); as shown in Figure 9. Ni/CeO2-Cr2O3 catalyst can be deemed as a superior catalyst due to its high catalytic activity (YCH4 > 95% zone) at low operating temperatures.

4. Conclusions

The screening of monometallic catalysts (i.e., Ni, Fe and Co) doped on CeO2-Cr2O3 support was studied in a milli-packed bed reactor. All the catalysts were prepared using one-pot hydrolysis. Ni was proven to be the most effective dopant. The amount of Ni loading was found optimal at 40% by weight, giving CO2 conversion of 98.7% and CH4 selectivity of 100% at a relatively low temperature of 290 °C. At temperatures of 200 to 350 °C, the reaction was kinetically driven by the higher operating temperature. However, thermodynamic limitation took place at temperatures higher than 350 °C where a drop in catalytic performance was observed. The catalyst was also stable during 50 h time on stream experiment. Ni-CeO2/Cr2O3 was proven to be one of the highest potential catalysts for the CO2 hydrogenation process of CH4 production.

Author Contributions

Conceptualization, U.W.H. and V.T.; methodology, V.T.; software, C.K.; validation, S.A., M.H.; formal analysis, S.A.; investigation, C.K.; resources, N.L. and U.W.H.; data curation, N.L.; writing—original draft preparation, V.T. and C.K.; writing—review and editing, M.H.; visualization, N.L.; supervision, N.L. and U.W.H.; project administration, V.T.; funding acquisition, U.W.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Research Council of Thailand (NRCT) contract number (1) N41A640149, (2) 668/2563 and (3) 130/2563; and King Mongkut’s University of Technology North Bangkok contract number (1) KMUTNB-BasicR-64-34, (2) KMUTNB-BasicR-64-31, and (3) KMUTNB-PHD-62-10.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Blunden, J.; Arndt, D.S. State of the Climate in 2018. Bull. Am. Meteorol. Soc. 2019, 100, S1–S429. [Google Scholar] [CrossRef] [Green Version]
  2. Cheah, W.Y.; Ling, T.C.; Juan, J.C.; Lee, D.-J.; Chang, J.-S.; Show, P.L. Biorefineries of carbon dioxide: From carbon capture and storage (CCS) to bioenergies production. Bioresour. Technol. 2016, 215, 346–356. [Google Scholar] [CrossRef]
  3. Nie, W.; Zou, X.; Chen, C.; Wang, X.; Ding, W.; Lu, X. Methanation of Carbon Dioxide over Ni–Ce–Zr Oxides Prepared by One-Pot Hydrolysis of Metal Nitrates with Ammonium Carbonate. Catalysts 2017, 7, 104. [Google Scholar] [CrossRef]
  4. Gao, J.; Liu, Q.; Gu, F.; Liu, B.; Zhong, Z.; Su, F. Recent advances in methanation catalysts for the production of synthetic natural gas. RSC Adv. 2015, 5, 22759–22776. [Google Scholar] [CrossRef]
  5. Chisholm, G.; Cronin, L. Hydrogen from Water Electrolysis; Elsevier: Amsterdam, The Netherlands, 2016; pp. 315–343. [Google Scholar]
  6. Huynh, H.L.; Yu, Z. CO2 Methanation on Hydrotalcite-Derived Catalysts and Structured Reactors: A Review. Energy Technol. 2020, 8. [Google Scholar] [CrossRef] [Green Version]
  7. Kierzkowska-Pawlak, H.; Tyczkowski, J.; Balcerzak, J.; Tracz, P. Advances in plasma produced CoOx-based nanocatalysts for CO2 methanation. Catal. Today 2019, 337, 162–170. [Google Scholar] [CrossRef]
  8. Zhao, B.; Chen, Z.; Chen, Y.; Ma, X. Syngas methanation over Ni/SiO2 catalyst prepared by ammonia-assisted impregnation. Int. J. Hydrog. Energy 2017, 42, 27073–27083. [Google Scholar] [CrossRef]
  9. Rostrup-Nielsen, J.; Pedersen, K.; Sehested, J. High temperature methanation: Sintering and structure sensitivity. Appl. Catal. A Gen. 2007, 330, 134–138. [Google Scholar] [CrossRef]
  10. Moghaddam, S.V.; Rezaei, M.; Meshkani, F.; Daroughegi, R. Carbon dioxide methanation over Ni-M/Al2O3 (M: Fe, CO, Zr, La and Cu) catalysts synthesized using the one-pot sol-gel synthesis method. Int. J. Hydrog. Energy 2018, 43, 16522–16533. [Google Scholar] [CrossRef]
  11. Shen, W.-J.; Okumura, M.; Matsumura, Y.; Haruta, M. The influence of the support on the activity and selectivity of Pd in CO hydrogenation. Appl. Catal. A Gen. 2001, 213, 225–232. [Google Scholar] [CrossRef]
  12. Takoudis, C.G. The catalytic synthesis of hydrocarbons from H2/CO mixtures over the group VIII metals: Comments on methanation kinetics. J. Catal. 1982, 78, 265. [Google Scholar] [CrossRef]
  13. Tsiotsias, A.I.; Charisiou, N.D.; Yentekakis, I.V.; Goula, M.A. The Role of Alkali and Alkaline Earth Metals in the CO2 Methanation Reaction and the Combined Capture and Methanation of CO2. Catalysts 2020, 10, 812. [Google Scholar] [CrossRef]
  14. Wang, F.; He, S.; Chen, H.; Wang, B.; Zheng, L.; Wei, M.; Evans, D.G.; Duan, X. Active Site Dependent Reaction Mechanism over Ru/CeO2 Catalyst toward CO2 Methanation. J. Am. Chem. Soc. 2016, 138, 6298–6305. [Google Scholar] [CrossRef] [PubMed]
  15. Le, T.A.; Kim, M.S.; Lee, S.H.; Kim, T.W.; Park, E.D. CO and CO2 methanation over supported Ni catalysts. Catal. Today 2017, 293-294, 89–96. [Google Scholar] [CrossRef]
  16. Liu, Q.; Zhong, Z.; Gu, F.; Wang, X.; Lu, X.; Li, H.; Xu, G.; Su, F. CO methanation on ordered mesoporous Ni–Cr–Al catalysts: Effects of the catalyst structure and Cr promoter on the catalytic properties. J. Catal. 2016, 337, 221–232. [Google Scholar] [CrossRef]
  17. Dufour, J.; Martos, C.; Ruiz, A.; Ayuela, F. Effect of the precursor on the activity of high temperature water gas shift catalysts. Int. J. Hydrog. Energy 2013, 38, 7647–7653. [Google Scholar] [CrossRef]
  18. Lohitharn, N.; Goodwinjr, J. Impact of Cr, Mn and Zr addition on Fe Fischer–Tropsch synthesis catalysis: Investigation at the active site level using SSITKA. J. Catal. 2008, 257, 142–151. [Google Scholar] [CrossRef]
  19. Ashok, J.; Ang, M.; Kawi, S. Enhanced activity of CO2 methanation over Ni/CeO2-ZrO2 catalysts: Influence of preparation methods. Catal. Today 2017, 281, 304–311. [Google Scholar] [CrossRef]
  20. Wang, X.; Zhu, L.; Liu, Y.; Wang, S. CO2 methanation on the catalyst of Ni/MCM-41 promoted with CeO2. Sci. Total Environ. 2018, 625, 686–695. [Google Scholar] [CrossRef] [PubMed]
  21. Branco, J.; Silva, R.; Ferreira, A. Methanation of CO2 over Cobalt-Lanthanide Aerogels: Effect of Calcination Temperature. Catalysts 2020, 10, 704. [Google Scholar] [CrossRef]
  22. Wang, H.; Tsilomelekis, G. Catalytic performance and stability of Fe-doped CeO2 in propane oxidative dehydrogenation using carbon dioxide as an oxidant. Catal. Sci. Technol. 2020, 10, 4362–4372. [Google Scholar] [CrossRef]
  23. Konishcheva, M.; Potemkin, D.; Snytnikov, P.; Zyryanova, M.; Pakharukova, V.; Simonov, P.; Sobyanin, V. Selective CO methanation in H2-rich stream over Ni–, Co– and Fe/CeO2: Effect of metal and precursor nature. Int. J. Hydrog. Energy 2015, 40, 14058–14063. [Google Scholar] [CrossRef]
  24. 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]
  25. Aziz, M.; Jalil, A.; Triwahyono, S.; Saad, M. CO2 methanation over Ni-promoted mesostructured silica nanoparticles: Influence of Ni loading and water vapor on activity and response surface methodology studies. Chem. Eng. J. 2015, 260, 757–764. [Google Scholar] [CrossRef]
  26. Singh, P.; Hegde, M.S. Ce0.67Cr0.33O2.11: A New Low-Temperature O2 Evolution Material and H2 Generation Catalyst by Thermochemical Splitting of Water. Chem. Mater. 2010, 22, 762–768. [Google Scholar] [CrossRef]
  27. Hu, Q.; Yue, B.; Shao, H.; Yang, F.; Wang, J.; Wang, Y.; Liu, J. Facile syntheses of cerium-based CeMO3 (M = Co, Ni, Cu) perovskite nanomaterials for high-performance supercapacitor electrodes. J. Mater. Sci. 2020, 55, 8421–8434. [Google Scholar] [CrossRef]
  28. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  29. Ghiasi, M.; Malekzadeh, A. Structural Features of (Ce, La or Sr)(Mn or Co)O3 Nano-Perovskites as a Catalyst for Carbon Monoxide Oxidation. Acta Met. Sin. 2014, 27, 635–641. [Google Scholar] [CrossRef]
  30. Qiu, Y.; Ye, N.; Situ, D.; Zuo, S.; Wang, X. Study of Catalytic Combustion of Chlorobenzene and Temperature Programmed Reactions over CrCeOx/AlFe Pillared Clay Catalysts. Materials 2019, 12, 728. [Google Scholar] [CrossRef] [Green Version]
  31. Li, W.; Liu, H.; Chen, Y. Promotion of transition metal oxides on the NH3-SCR performance of ZrO2-CeO2 catalyst. Front. Environ. Sci. Eng. 2017, 11, 6. [Google Scholar] [CrossRef]
  32. Yang, P.; Yang, S.; Shi, Z.; Meng, Z.; Zhou, R. Deep oxidation of chlorinated VOCs over CeO2 -based transition metal mixed oxide catalysts. Appl. Catal. B Environ. 2015, 162, 227–235. [Google Scholar] [CrossRef]
  33. Liu, H.; Wei, L.; Yue, R.; Chen, Y. CrOx–CeO2 binary oxide as a superior catalyst for NO reduction with NH3 at low temperature in presence of CO. Catal. Commun. 2010, 11, 829–833. [Google Scholar] [CrossRef]
  34. Xu, X.; Li, L.; Yu, F.; Peng, H.; Fang, X.; Wang, X. Mesoporous high surface area NiO synthesized with soft templates: Remarkable for catalytic CH4 deep oxidation. Mol. Catal. 2017, 441, 81–91. [Google Scholar] [CrossRef]
  35. Khajonvittayakul, C.; Tongnan, V.; Kangsadan, T.; Laosiripojana, N.; Jindasuwan, S.; Hartley, U.W. Thermodynamic and mechanism study of syngas production via integration of nitrous oxide decomposition and methane partial oxidation in the presence of 10%NiO–La0.3Sr0.7Co0.7Fe0.3O3−δ. React. Kinet. Mech. Catal. 2019, 127, 839–855. [Google Scholar] [CrossRef]
  36. Khajonvittayakul, C.; Tongnan, V.; Namo, N.; Phonbubpha, C.; Laosiripojana, N.; Hartley, M.; Hartley, U.W. Tar steam reforming for synthesis gas production over Ni-based on ceria/zirconia and La0.3Sr0.7Co0.7Fe0.3O3 in a packed-bed reactor. Chemosphere 2021, 277. [Google Scholar] [CrossRef]
  37. Saharuddin, T.S.T.; Samsuri, A.; Salleh, F.; Othaman, R.; Bin Kassim, M.; Hisham, M.W.M.; Yarmo, M.A. Studies on reduction of chromium doped iron oxide catalyst using hydrogen and various concentration of carbon monoxide. Int. J. Hydrog. Energy 2017, 42, 9077–9086. [Google Scholar] [CrossRef]
  38. Reddy, G.K.; Gunasekara, K.; Boolchand, P.; Smirniotis, P.G. Cr- and Ce-Doped Ferrite Catalysts for the High Temperature Water−Gas Shift Reaction: TPR and Mossbauer Spectroscopic Study. J. Phys. Chem. C 2010, 115, 920–930. [Google Scholar] [CrossRef]
  39. Díaz, E.; Ordóñez, S.; Vega, A.; Coca, J. Catalytic combustion of hexane over transition metal modified zeolites NaX and CaA. Appl. Catal. B Environ. 2005, 56, 313–322. [Google Scholar] [CrossRef]
  40. Stangeland, K.; Kalai, D.; Li, H.; Yu, Z. CO2 Methanation: The Effect of Catalysts and Reaction Conditions. Energy Procedia 2017, 105, 2022–2027. [Google Scholar] [CrossRef]
  41. Tada, S.; Shimizu, T.; Kameyama, H.; Haneda, T.; Kikuchi, R. Ni/CeO2 catalysts with high CO2 methanation activity and high CH4 selectivity at low temperatures. Int. J. Hydrog. Energy 2012, 37, 5527–5531. [Google Scholar] [CrossRef]
  42. Pastor-Pérez, L.; Saché, E.; Jones, C.; Gu, S.; Arellano-Garcia, H.; Reina, T. Synthetic natural gas production from CO2 over Ni-x/CeO2-ZrO2 (x = Fe, Co) catalysts: Influence of promoters and space velocity. Catal. Today 2018, 317, 108–113. [Google Scholar] [CrossRef]
  43. Aziz, M.; Jalil, A.; Triwahyono, S.; Mukti, R.; Taufiq-Yap, Y.; Sazegar, M. Highly active Ni-promoted mesostructured silica nanoparticles for CO2 methanation. Appl. Catal. B Environ. 2014, 147, 359–368. [Google Scholar] [CrossRef]
  44. Lu, B.; Kawamoto, K. Preparation of the highly loaded and well-dispersed NiO/SBA-15 for methanation of producer gas. Fuel 2013, 103, 699–704. [Google Scholar] [CrossRef]
  45. Yamasaki, M.; Habazaki, H.; Asami, K.; Izumiya, K.; Hashimoto, K. Effect of tetragonal ZrO2 on the catalytic activity of Ni/ZrO2 catalyst prepared from amorphous Ni–Zr alloys. Catal. Commun. 2006, 7, 24–28. [Google Scholar] [CrossRef]
  46. Perkas, N.; Amirian, G.; Zhong, Z.; Teo, J.; Gofer, Y.; Gedanken, A. Methanation of Carbon Dioxide on Ni Catalysts on Mesoporous ZrO2 Doped with Rare Earth Oxides. Catal. Lett. 2009, 130, 455–462. [Google Scholar] [CrossRef]
  47. Gac, W.; Zawadzki, W.; Rotko, M.; Słowik, G.; Greluk, M. CO2 Methanation in the Presence of Ce-Promoted Alumina Supported Nickel Catalysts: H2S Deactivation Studies. Top. Catal. 2019, 62, 524–534. [Google Scholar] [CrossRef] [Green Version]
  48. Wongsartsai, C.; Tongnan, V.; Sornchamni, T.; Siri-Nguan, N.; Laosiripojana, N.; Hartley, M.; Hartley, U.W. CO2 utilization via methanation using 40%Ni/CexCr1-xO2 as a novel catalyst: A comparative study of packed-bed and micro-channel reactors. React. Kinet. Mech. Catal. 2020, 131, 1–17. [Google Scholar] [CrossRef]
  49. Bacariza, C.; Graça, I.; Westermann, A.; Ribeiro, M.F.; Lopes, J.; Henriques, C. CO2 Hydrogenation Over Ni-Based Zeolites: Effect of Catalysts Preparation and Pre-reduction Conditions on Methanation Performance. Top. Catal. 2015, 59, 314–325. [Google Scholar] [CrossRef]
  50. Liu, H.; Zou, X.; Wang, X.; Lu, X.; Ding, W. Effect of CeO2 addition on Ni/Al2O3 catalysts for methanation of carbon dioxide with hydrogen. J. Nat. Gas Chem. 2012, 21, 703–707. [Google Scholar] [CrossRef]
  51. Ocampo, F.; Louis, B.; Kiennemann, A.; Roger, A.-C. CO2 methanation over Ni-Ceria-Zirconia catalysts: Effect of preparation and operating conditions. In IOP Conference Series: Materials Science and Engineering; IOP Publishing: Bristol, UK, 2011; Volume 19, p. 012007. [Google Scholar] [CrossRef]
  52. Sepehri, S.; Rezaei, M. Preparation of Highly Active Nickel Catalysts Supported on Mesoporous Nanocrystallineγ-Al2O3for Methane Autothermal Reforming. Chem. Eng. Technol. 2015, 38, 1637–1645. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of (a) 40% NiO/CeO2-Cr2O3, (b) 40% Fe2O3/CeO2-Cr2O3, and (c) 40% Co3O4/CeO2-Cr2O3 after calcined at 500 °C, where NiO (●), Fe2O3 (♦), Co3O4 (▲), Cr2O3 (▼), and CeO2-Cr2O3 (■) phases.
Figure 1. XRD patterns of (a) 40% NiO/CeO2-Cr2O3, (b) 40% Fe2O3/CeO2-Cr2O3, and (c) 40% Co3O4/CeO2-Cr2O3 after calcined at 500 °C, where NiO (●), Fe2O3 (♦), Co3O4 (▲), Cr2O3 (▼), and CeO2-Cr2O3 (■) phases.
Catalysts 11 01159 g001
Figure 2. H2-TPR profiles of (a) pure CeO2, (b) CeO2-Cr2O3, (c) 40%wt. NiO/CeO2-Cr2O3, (d) 40%wt. Fe2O3/CeO2-Cr2O3, and (e) 40%wt. Co2O3/CeO2-Cr2O3 catalyst calcined at 500 °C.
Figure 2. H2-TPR profiles of (a) pure CeO2, (b) CeO2-Cr2O3, (c) 40%wt. NiO/CeO2-Cr2O3, (d) 40%wt. Fe2O3/CeO2-Cr2O3, and (e) 40%wt. Co2O3/CeO2-Cr2O3 catalyst calcined at 500 °C.
Catalysts 11 01159 g002
Figure 3. Catalytic performance of CO2 methanation in terms of CO2 conversion (Left) and CH4 selectivity (Right) over different metals: (▼) Ni; (♦) Fe; and (●) Co; on CeO2-Cr2O3 support.
Figure 3. Catalytic performance of CO2 methanation in terms of CO2 conversion (Left) and CH4 selectivity (Right) over different metals: (▼) Ni; (♦) Fe; and (●) Co; on CeO2-Cr2O3 support.
Catalysts 11 01159 g003
Figure 4. Catalytic performance of y %wt. NiO/CeO2-Cr2O3 catalyst for CO2 methanation.
Figure 4. Catalytic performance of y %wt. NiO/CeO2-Cr2O3 catalyst for CO2 methanation.
Catalysts 11 01159 g004
Figure 5. The catalytic performance with different reduction temperature using Ni-based on CeO2-Cr2O3 catalyst.
Figure 5. The catalytic performance with different reduction temperature using Ni-based on CeO2-Cr2O3 catalyst.
Catalysts 11 01159 g005
Figure 6. Time on steam of 40%wt. NiO/CeO2-Cr2O3 after H2/CO2 = 4:1 exposure at 290 °C and 27,624 mL.h−1·gcat−1 for 50 h.
Figure 6. Time on steam of 40%wt. NiO/CeO2-Cr2O3 after H2/CO2 = 4:1 exposure at 290 °C and 27,624 mL.h−1·gcat−1 for 50 h.
Catalysts 11 01159 g006
Figure 7. XRD patterns of 40%wt. NiO/CeO2-Cr2O3: (a) before reaction and (b) after stability, where Ni (*) and CeO2-Cr2O3 (■) phases.
Figure 7. XRD patterns of 40%wt. NiO/CeO2-Cr2O3: (a) before reaction and (b) after stability, where Ni (*) and CeO2-Cr2O3 (■) phases.
Catalysts 11 01159 g007
Figure 8. Surface morphology of 40%wt. NiO/CeO2-Cr2O3 at magnitude × 100,000: before (Left) reaction and after (Right) stability.
Figure 8. Surface morphology of 40%wt. NiO/CeO2-Cr2O3 at magnitude × 100,000: before (Left) reaction and after (Right) stability.
Catalysts 11 01159 g008
Figure 9. The summaries of the catalytic performances toward CO2 methanation over various catalysts at H2/CO2 = 4:1.
Figure 9. The summaries of the catalytic performances toward CO2 methanation over various catalysts at H2/CO2 = 4:1.
Catalysts 11 01159 g009
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Khajonvittayakul, C.; Tongnan, V.; Amornraksa, S.; Laosiripojana, N.; Hartley, M.; Hartley, U.W. CO2 Hydrogenation to Synthetic Natural Gas over Ni, Fe and Co–Based CeO2–Cr2O3. Catalysts 2021, 11, 1159. https://doi.org/10.3390/catal11101159

AMA Style

Khajonvittayakul C, Tongnan V, Amornraksa S, Laosiripojana N, Hartley M, Hartley UW. CO2 Hydrogenation to Synthetic Natural Gas over Ni, Fe and Co–Based CeO2–Cr2O3. Catalysts. 2021; 11(10):1159. https://doi.org/10.3390/catal11101159

Chicago/Turabian Style

Khajonvittayakul, Chalempol, Vut Tongnan, Suksun Amornraksa, Navadol Laosiripojana, Matthew Hartley, and Unalome Wetwatana Hartley. 2021. "CO2 Hydrogenation to Synthetic Natural Gas over Ni, Fe and Co–Based CeO2–Cr2O3" Catalysts 11, no. 10: 1159. https://doi.org/10.3390/catal11101159

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