Next Article in Journal
High-Efficiency p-n Homojunction Perovskite and CIGS Tandem Solar Cell
Previous Article in Journal
Surface and Void Space Analysis of the Crystal Structures of Two Lithium Bis(pentafluoroethanesulfonyl)imide Salts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ni-CeO2 Catalyst with High Ni Loading Prepared by Salt-Assisted Solution Combustion for CO2 Methanation

1
Zhejiang Key Laboratory of Petrochemical Pollution Control, Zhejiang Ocean University, Zhoushan 316022, China
2
SINOPEC Dalian Research Institute of Petroleum and Petrochemicals, Dalian 116045, China
3
School of Food and Pharmaceutical, Zhejiang Ocean University, Zhoushan 316022, China
4
Department of Chemical and Biological Engineering, School of Engineering, Rensselaer Polytechnic Institute, Troy, NY 12180, USA
5
Department of Chemical Engineering, School of Petrochemical Engineering & Environment, Zhejiang Ocean University, Zhoushan 316022, China
6
National-Local Joint Engineering Laboratory of Harbor Oil & Gas Storage and Transportation Technology, Zhejiang Ocean University, Zhoushan 316022, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Submission received: 17 April 2022 / Revised: 10 May 2022 / Accepted: 13 May 2022 / Published: 15 May 2022

Abstract

:
An Ni-CeO2 catalyst with high Ni loading (50 wt.%) prepared by a salt-assisted solution combustion method was characterized by different methods and used for CO2 methanation. The specific surface area of the Ni-CeO2 catalyst prepared by salt-assisted solution combustion is 7 times that of the catalyst prepared by conventional solution combustion. The Ni-CeO2 catalyst prepared by salt-assisted solution combustion has smaller particle sizes of Ni and exhibits excellent activity at low temperatures. The high Ni loading and small Ni particle size can provide more metal Ni site and Ni-CeO2 interface, which help to improve the CO2 methanation performance.

1. Introduction

Excessive emission of carbon dioxide (CO2) will aggravate the greenhouse effect, and CO2 conversion has attracted more and more attention. CO2 methanation is one of the most popular technologies for CO2 conversion; it is also called the Sabatier reaction [1,2,3]. CO2 methanation provides practical ideas for the large-scale conversion of renewable hydrogen to CH4. CH4, the product of CO2 methanation, can be stored in natural gas network facilities.
CO2 methanation is a reaction in which CO2 is reduced to CH4 and subject to kinetic limitations [4,5,6]. Ru [7], Pd [8], Ni [9], etc., supported on different supports (SiO2 [8,10], ZrO2 [11,12], CeO2 [13,14], etc.), have been investigated for CO2 methanation. Supported Ni-based catalysts with relatively high activities as well as lower costs have become the most widely studied materials in recent years. In addition, CO2 methanation is a highly exothermic reaction, so the equilibrium conversion of carbon dioxide will decrease with the increase in reaction temperature [12]. Therefore, it is necessary to develop Ni-based CO2 methanation catalysts with high activity at low temperature.
Solution combustion methods have made outstanding contributions in the preparation of nanometer material [15,16]. Solution combustion methods have a series of superiorities in the preparation process, such as a relatively simple device, low cost involved, fast reaction process, and low preheating temperature [17,18,19]. However, the solution combustion method often results in the shortcoming of a low specific surface area as well as large particle size, which will decrease the catalyst activity.
Salt-assisted solution combustion is a new method to prepare nanomaterials with a high specific surface area, which can overcome the disadvantage of the low specific surface area of nanomaterials prepared by a conventional solution combustion method [20,21,22,23]. During combustion, inert salt such as NaCl and KCl can be deposited between nanoparticles, preventing the agglomeration and sintering of the nanoparticles. Then, by dissolving the salt with water or other solvents, nanomaterials with high specific surface areas can be obtained.
CeO2 has an excellent oxidation-reduction property and is widely used as a support in catalysts, which can inhibit the sintering of metal and minimize carbon deposition [24,25]. The oxygen vacancies in CeO2 caused by the Ce3+/Ce4+ ion pair make CO2 effectively activated, and a lot of studies have proved that using CeO2 can decrease Ni particle size in the catalyst and produce oxygen vacancies, and then the activity of the catalyst is boosted. Ni-CeO2 catalyst has been reported for CO2 methanation [26,27,28,29,30,31,32,33]. Ye et al. [29] found that the maximization of the Ni-CeO2 interface is the key to improve the catalyst activity in the low-temperature CO2 transformation process, and the Ni embedded into CeO2 can improve the stability of Ni nanoparticles. At a lower reaction temperature, the enriched Ni-CeO2 interface and Ni particles with higher dispersity can work synergistically, thus effectively completing the CO2 methanation process [30,31].
Theoretically, increasing Ni content and decreasing Ni particle size can increase the Ni-Ce interface. However, the existing literature on Ni-CeO2 catalysts is mainly regarding low Ni-content catalysts. This is because a higher Ni load is often accompanied by a larger Ni particle size, which reduces the catalytic activity.
In this paper, to the best of our knowledge, Ni-CeO2 (SSC) catalysts with 50% Ni loading were prepared by salt-assisted solution combustion and used for CO2 methanation for the first time. The salt-assisted solution combustion method can solve the problem of a low specific surface area caused by the conventional solution combustion method and obtain small Ni particles when the catalyst has a higher Ni loading. This enables the catalyst with higher Ni loadings to achieve high activity at low temperatures in the CO2 methanation reaction.

2. Materials and Methods

2.1. Catalyst Preparation

An Ni-CeO2(SSC) catalyst with 50 wt.% Ni loading was prepared by a salt-assisted solution combustion method. For salt-assisted solution combustion, Ni(NO3)2·6H2O and Ce(NO3)3·6H2O were added as oxidants, glycine was added as fuel and NaCl as salt. An amount of 2.48 g of Ni(NO3)2·6H2O, 1.26 g of Ce(NO3)3·6H2O, 0.89 g of glycine, and 0.35 g of NaCl were put into beaker and dissolved in 20 mL of deionized water. The beaker was heated at 210 °C until the water evaporated and a combustion reaction took place to produce a solid powder. The solid powder was calcinated at 450 °C for 4 h. After cooling, the sample was washed several times with deionized water to remove NaCl. Silver nitrate solution was used to detect whether there was Cl residue in the washing solution. The Ni-CeO2 (SSC) catalyst was obtained after drying at 120 °C.
For comparison, the Ni-CeO2(SC) catalyst with 50 wt.% Ni loading was prepared by a solution combustion method without adding NaCl. In this case, the sample was calcined at 450 °C for 4 h.
The Ni and Na content were determined by atomic absorption spectrophotometer (AAS). The Ni content of the Ni-CeO2(SSC) and Ni-CeO2(SC) catalysts was 46.3 wt.% and 48.5 wt.%, respectively, which is close to the theoretical value. The Na residue in the Ni-CeO2(SSC) catalyst was 0.06 wt.%.

2.2. Catalyst Characterization

Nitrogen adsorption and desorption were measured using an Autosorb-iQ analyzer (Quanta chrome Instruments, Boynton Beach, FL, USA) at −196 °C. The specific surface area was calculated by BET method. The average pore size distribution was obtained from the desorption isotherm using the BJH method.
The X-ray diffraction (XRD) measurements were carried out on a DX-2700 X-ray diffractometer (Haoyuan Instrument, Dandong, China) using Cu Kα radiation. The scanning range of 2θ was from 10° to 80° with a step size of 0.02°/1 s. The crystallite sizes in the catalysts were obtained with the Scherrer equation.
H2 temperature programmed reduction (H2-TPR) was operated on a multifunctional adsorption instrument TP-5080 (xianquan Industrial and trading Co., Ltd., Tianjin, China), and the instrument was used with a thermal conductivity detection device (TCD). A quartz tube was charged with 50 mg of catalyst and then pretreated with Ar (20 mL/min) at 400 °C for 25 min. After cooling, the reactor was purged using 5% H2/Ar (20 mL/min) for 0.5 h. Finally, H2-TPR tests were performed in 5% H2/Ar (20 mL/min) from 25 °C to 850 °C (10 °C /min).
The tested instrument of H2 temperature programmed desorption (H2-TPD) is consistent with H2-TPR. First, 50 mg of catalyst was reduced at 450 °C with 5% H2/Ar (20 mL/min) for 40 min, then it was reduced to 25 °C and adsorbed at 50 °C for 15min. After being purged with Ar for 1h, desorption was performed in Ar (20 mL/min) from 30 to 830 °C (10 °C/min). The specific surface area of Ni was calculated from the amount of H2 desorption using the assumption that the stoichiometric ratio of H/Nisurface is 1.
TEM measurements were performed on a Tecnai G2 F20 (FEI company, USA) microscope. Before the test, the catalyst powder was evenly dispersed in ethanol after ultrasonic treatment for 30 min, and then the solution was dripped onto the carbon-supported copper mesh by a capillary tube, dried and analyzed.

2.3. Catalyst Performance Test

Catalytic performance tests were performed at atmospheric pressure, and the catalyst was mixed with quartz sand at a ratio of 1:10 and then loaded into a quartz tube reactor (inner diameter is 8 mm, outer diameter is 12 mm, length is 49 cm). The reactor temperature was controlled by a PID temperature controller, and a thermocouple was placed in the middle of catalyst bed. After reduction at 450 °C for 40 min and cooling to 250 °C, the reaction was carried out in a gas environment of 40 mL/min of H2, 10 mL/min of CO2, and 50 mL/min of N2. The aging test of the catalysts was carried out at WHSV of 300,000 mL g−1 h−1. The temperature range was 250–450 °C. The outlet gas of the reactor was first condensed and dehydrated, and then analyzed by a PGENERAL G5 on-line gas chromatography. The gas chromatography was equipped with a TCD detector and a TDX-01 column and used H2 as carrier gas.
CO2 conversion and CH2 selectivity are calculated as follows:
X CO 2 = F CO 2 , in F CO 2 , out F CO 2 , in × 100 %
S CH 4 = F CH 4 , out F CO 2 , in F CO 2 , out × 100 %
where F is the molar flow rate.

3. Results and Discussion

The nitrogen adsorption/desorption isotherms of Ni-CeO2(SSC) and Ni-CeO2(SC) catalysts are shown in Figure 1a. From the IUPAC classification, the isotherms showed typical type IV isotherms and hysteresis loops [34], which indicates that the catalyst contains a mesopore structure. Analyzing Figure 1b, the mesoporous structure in the Ni-CeO2(SSC) catalyst increased significantly after NaCl was used. As shown in Table 1, the specific surface area of Ni-CeO2(SSC) is about 7 times that of the Ni-CeO2(SC) catalyst, and the pore volume of Ni-CeO2(SSC) is also higher than that of Ni-CeO2(SC). Therefore, the addition of NaCl allowed the pore structure of the catalyst to be optimized, while the SBET of the catalyst was greatly increased.
After the catalyst was reduced at 450 °C for 40 min, XRD tests were performed, and the results are collated in Figure 2. The diffraction peaks at 44° and 51.8° are attributed to metal Ni [35], while others are attributed to CeO2. The addition of NaCl increases the half peak width of all diffraction peaks. It suggests that the addition of NaCl reduces the particle size of Ni and CeO2 in the catalyst, and the dispersity of Ni in the catalysts was enhanced. The data in Table 2 about the crystal sizes of catalysts were calculated according to the Scherrer equation [36]. From the data in Table 2, it can be concluded that the Ni and CeO2 particle sizes of the reduced Ni-CeO2(SSC) is much smaller than that of the Ni-CeO2(SC). After NaCl was used, the Ni crystal size decreased from 21.4 nm to 7.2 nm, and the specific surface area of Ni (SNi) increased from 5.87 to 11.45 m2/gcat. Kang [37] prepared 24 wt.% Ni/CeO2-HH-2 catalyst with the solution combustion method, and the grain size of Ni reached 23 nm, while the SBET was about 14 m2/g, which is similar to the Ni-CeO2(SC) in our work.
The catalyst reduced at 450 °C for 40 min was characterized by TEM, and the results are shown in Figure 3. As shown in Figure 3a, the particle size of the Ni-CeO2(SC) catalyst is about 20 nm; large Ni and CeO2 particles can be found in Figure 3b. By taking the statistics of Figure 3c, the particle size of Ni-CeO2(SSC) catalysts is about 10 nm. As shown in Figure 3d, the small particles of Ni and CeO2 are observed to have obvious contact in the Ni-CeO2(SSC) catalyst. It can be seen in Figure 3e,f that the particle size in the catalyst of Ni-CeO2(SSC) is far less than Ni-CeO2(SC). This result is consistent with that obtained by XRD, indicating that the salt-assisted solution combustion method allows the particle size of catalysts to decrease significantly and form a rich metal–support interface.
Figure 4 shows the H2-TPR spectra of the Ni-CeO2(SC) and Ni-CeO2(SSC) catalysts. The Ni-CeO2(SC) catalyst exhibits a distinct peak centered at 404 °C. Combining the results of XRD and TEM, this reduction peak should be due to the decrease in bulk NiO. A weak wide peak around 780 °C was caused by the decrease in bulk CeO2 [38]. The reduction peak for superficial CeO2 was usually between 400 and 600 °C [39], which should be covered by the peak of NiO.
With regard to the Ni-CeO2(SSC) catalyst, an indistinct reduction peak can be observed at around 250 °C, which is due to the oxygen vacancy generated in NixCe1-xOy solid solution adsorbing oxygen and reducing it [38]. A few Ni2+ ions can enter the CeO2 crystal lattice to form a NixCe1-xOy solid solution, and the solid solution causes the generation of oxygen vacancies [38]. Some oxygen is more easily adsorbed into the oxygen vacancy (Ni-Vox-Ce) of the solid solution, which can be reduced to below 300 °C. This indicates that the use of NaCl promotes the generation of the NixCe1-xOy solid solution. In addition, the oxygen vacancies (Ni-Vox-Ce) in the NixCe1-xOy solid solution play a huge part in promoting the activity of the CO2 methanation reaction. The wide reduction peak from 300 to 500 °C belongs to the reduction peak of NiO. Most of the reduction peaks are concentrated at about 350 °C, which is the result of reduction after the weak interaction between highly dispersed NiO and CeO2, while a slightly wider reduction peak found near 440 °C belongs to the reduction of surface CeO2, and the highly dispersed NiO strongly interacted with CeO2 [38]. The reduction of surface CeO2 is usually between 400 and 600 °C. The TPR results indicate that the salt-assisted solution combustion method makes it easy to form NixCe1-xOy solid solution in the catalyst; at the same time, the interaction between NiO and CeO2 was increased.
The performance of the two catalysts was tested for CO2 methanation. As shown in Figure 5a,b, when the temperature range is 250 to 350 °C, the two groups of catalysts showed similar regularities. The CO2 conversion of the Ni-CeO2(SSC) catalyst is much higher than that of the Ni-CeO2(SC) catalyst. At 400 °C, the CH4 selectivity of the two catalysts began to decrease (as shown in Figure 5b), which was due to the emergence of a small amount of CO from the reverse water-gas shift reaction at high temperature. Especially when the temperature was at 250 °C, the CO2 conversion (9.9%) of Ni-CeO2(SSC) was about 3 times the CO2 conversion (3.4%) of Ni-CeO2(SC).
Figure 6 shows the CO2 conversion during the stability test of Ni-CeO2 (SSC) at 350 °C for 50 h. The sample shows high stability, and the carbon dioxide conversion rate is only reduced by 6.2% within 50 h. This could be due to the interaction between the small Ni and CeO2 particles, which can prevent the nickel particles from sintering during the reaction.
According to the performance test and characterization results of the catalysts, it can be seen that the salt-assisted solution combustion method can significantly improve the specific surface area and reduce the metal and support particle sizes of the high metal loading Ni-CeO2 catalyst, while facilitating the formation of oxygen vacancies (Ni-Vox-Ce), thus improving the activity of the catalyst. The small Ni particle size provides more metallic Ni surface for adsorption and activation of hydrogen, which then can react with CO2 adsorbed on CeO2 to form CH4. The enriched Ni-CeO2 interface makes it easier for adsorbed hydrogen to react with adsorbed CO2. Rui et al. [30] found that the abundant Ni-CeO2 interfacial sites had more excellent activity than the inerratic lattice oxygen atoms in CeO2. Tada et al. [31] reported that the CO2 conversion rate had no direct connection with the number of oxygen vacancies on CeO2 (Ce-Vox-Ce) but was affected by the number of Ni-Vox-Ce sites.

4. Conclusions

The Ni-CeO2(SSC) catalyst with a high Ni loading of 50 wt.% prepared by a salt-assisted solution combustion method exhibits a high CO2 conversion rate in CO2 methanation. Compared with the Ni-CeO2(SC) catalyst, the Ni-CeO2(SSC) catalyst has a larger specific surface area and smaller Ni particle size, and thus has a rich metal surface and metal–support interface, so it has good performance in methanation reactions. A salt-assisted solution combustion method can overcome the problems of particle sintering and low specific surface area caused by the traditional solution combustion method and provides a simple strategy for the preparation of high metal loading and high dispersed metal catalyst.

Author Contributions

Conceptualization, L.W.; data curation, C.H.; formal analysis, C.H., Z.C., J.Y., X.L., Z.J. and Y.Z.; funding acquisition, Z.C. and L.W.; investigation, C.H. and L.W.; writing—original draft, C.H., H.L. and L.W.; writing—review and editing, S.Y., X.Z. and L.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Science and Technology Foundation of Zhoushan (No. 2022C41002) and SINOPEC (No. 112109).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nie, W.; Zou, X.; Shang, X.; Wang, X.; Ding, W.; Lu, X. CeO2-assisted Ni nanocatalysts supported on mesoporous γ-Al2O3 for the production of synthetic natural gas. Fuel 2017, 202, 135–143. [Google Scholar] [CrossRef]
  2. Wang, W.; Wang, S.; Ma, X.; Gong, J. Recent advances in catalytic hydrogenation of carbon dioxide. Chem. Soc. Rev. 2011, 40, 3703–3727. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Pandey, D.; Ray, K.; Bhardwaj, R.; Bojja, S.; Chary, K.; Deo, G. Promotion of unsupported nickel catalyst using iron for CO2 methanation. Int. J. Hydrog. Energy 2018, 43, 4987–5000. [Google Scholar] [CrossRef]
  4. Biset-Peiró, M.; Guilera, J.; Zhang, T.; Arbiol, J.; Andreu, T. On the role of ceria in Ni-Al2O3 catalyst for CO2 plasma methanation. Appl. Catal. A Gen. 2019, 575, 223–229. [Google Scholar] [CrossRef]
  5. Xu, J.; Su, X.; Duan, H.; Hou, B.; Lin, Q.; Liu, X.; Pan, X.; Pei, G.; Geng, H.; Huang, Y. Influence of pretreatment temperature on catalytic performance of rutile TiO2-supported ruthenium catalyst in CO2 methanation. J. Catal. 2016, 333, 227–237. [Google Scholar] [CrossRef]
  6. Su, X.; Xu, J.; Liang, B.; Duan, H.; Hou, B.; Huang, Y. Catalytic carbon dioxide hydrogenation to methane: A review of recent studies. J. Energy Chem. 2016, 25, 553–565. [Google Scholar] [CrossRef]
  7. 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]
  8. Kim, H.Y.; Lee, H.M.; Park, J.-N. Bifunctional mechanism of CO2 methanation on Pd-MgO/SiO2 catalyst: Independent roles of MgO and Pd on CO2 methanation. J. Phys. Chem. C 2010, 114, 7128–7131. [Google Scholar] [CrossRef]
  9. Wei, W.; Jinlong, G. Methanation of carbon dioxide: An overview. Front. Chem. Sci. Eng. 2011, 5, 2–10. [Google Scholar] [CrossRef]
  10. 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]
  11. Zhao, K.; Wang, W.; Li, Z. Highly efficient Ni/ZrO2 catalysts prepared via combustion method for CO2 methanation. J. CO2 Util. 2016, 16, 236–244. [Google Scholar] [CrossRef]
  12. Tan, J.; Wang, J.; Zhang, Z.; Ma, Z.; Wang, L.; Liu, Y. Highly dispersed and stable Ni nanoparticles confined by MgO on ZrO2 for CO2 methanation. Appl. Surf. Sci. 2019, 481, 1538–1548. [Google Scholar] [CrossRef]
  13. Atzori, L.; Cutrufello, M.G.; Meloni, D.; Cannas, C.; Gazzoli, D.; Monaci, R.; Sini, M.F.; Rombi, E. Highly active NiO-CeO2 catalysts for synthetic natural gas production by CO2 methanation. Catal. Today 2018, 299, 183–192. [Google Scholar] [CrossRef]
  14. Ge, Y.; He, T.; Han, D.; Li, G.; Zhao, R.; Wu, J. Plasma-assisted CO2 methanation: Effects on the low-temperature activity of an Ni–Ce catalyst and reaction performance. R. Soc. Open Sci. 2019, 6, 190750. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Varma, A.; Mukasyan, A.S.; Rogachev, A.S.; Manukyan, K.V. Solution Combustion Synthesis of Nanoscale Materials. Chem. Rev. 2016, 116, 14493–14586. [Google Scholar] [CrossRef] [PubMed]
  16. Carlos, E.; Martins, R.; Fortunato, E.; Branquinho, R. Solution combustion synthesis: Towards a sustainable approach for metal oxides. Chem.—A Eur. J. 2020, 26, 9099–9125. [Google Scholar] [CrossRef] [PubMed]
  17. Cross, A.; Roslyakov, S.; Manukyan, K.V.; Rouvimov, S.; Rogachev, A.S.; Kovalev, D.; Wolf, E.E.; Mukasyan, A.S. In situ preparation of highly stable Ni-based supported catalysts by solution combustion synthesis. J. Phys. Chem. C 2014, 118, 26191–26198. [Google Scholar] [CrossRef]
  18. Ullah, N.; Qu, J.; Li, Z. Enhanced sulfur-resistant methanation performance over MoO3–ZrO2 catalyst prepared by solution combustion method. Appl. Organomet. Chem. 2019, 33, e5022. [Google Scholar]
  19. Jiang, Y.; Huang, T.; Dong, L.; Qin, Z.; Ji, H. Ni/bentonite catalysts prepared by solution combustion method for CO2 methanation. Chin. J. Chem. Eng. 2018, 26, 2361–2367. [Google Scholar] [CrossRef]
  20. Abbasian, A.R.; Mahvary, A.; Alirezaei, S. Salt-assisted solution combustion synthesis of NiFe2O4: Effect of salt type. Ceram. Int. 2021, 47, 23794–23802. [Google Scholar] [CrossRef]
  21. Chen, W.; Li, F.; Yu, J.; Liu, L. A facile and novel route to high surface area ceria-based nanopowders by salt-assisted solution combustion synthesis. Mater. Sci. Eng. B 2006, 133, 151–156. [Google Scholar] [CrossRef]
  22. Abbasian, A.R.; Rahmani, M. Salt-assisted solution combustion synthesis of nanostructured ZnFe2O4-ZnS powders. Inorg. Chem. Commun. 2020, 111, 107629. [Google Scholar] [CrossRef]
  23. Zhao, Y.; Jiang, N.; Zhang, X.; Guo, J.; Yang, Q.; Gao, L.; Li, Y.; Ma, T. One-step salt-assisted solution combustion synthesis of Ni-based composites for use as supercapacitor electrodes. J. Alloy. Compd. 2018, 765, 396–404. [Google Scholar] [CrossRef]
  24. Zhou, G.; Liu, H.; Cui, K.; Jia, A.; Hu, G.; Jiao, Z.; Liu, Y.; Zhang, X. Role of surface Ni and Ce species of Ni/CeO2 catalyst in CO2 methanation. Appl. Surf. Sci. 2016, 383, 248–252. [Google Scholar] [CrossRef]
  25. Guo, Y.; Zou, J.; Shi, X.; Rukundo, P.; Wang, Z.-j. A Ni/CeO2–CDC-SiC catalyst with improved coke resistance in CO2 reforming of methane. ACS Sustain. Chem. Eng. 2017, 5, 2330–2338. [Google Scholar] [CrossRef]
  26. Zhou, G.; Liu, H.; Cui, K.; Xie, H.; Jiao, Z.; Zhang, G.; Xiong, K.; Zheng, X. Methanation of carbon dioxide over Ni/CeO2 catalysts: Effects of support CeO2 structure. Int. J. Hydrog. Energy 2017, 42, 16108–16117. [Google Scholar] [CrossRef]
  27. 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]
  28. Tada, S.; Ikeda, S.; Shimoda, N.; Honma, T.; Takahashi, M.; Nariyuki, A.; Satokawa, S. Sponge Ni catalyst with high activity in CO2 methanation. Int. J. Hydrog. Energy 2017, 42, 30126–30134. [Google Scholar] [CrossRef]
  29. Ye, R.-P.; Li, Q.; Gong, W.; Wang, T.; Razink, J.J.; Lin, L.; Qin, Y.-Y.; Zhou, Z.; Adidharma, H.; Tang, J. High-performance of nanostructured Ni/CeO2 catalyst on CO2 methanation. Appl. Catal. B Environ. 2020, 268, 118474. [Google Scholar] [CrossRef]
  30. Rui, N.; Zhang, X.; Zhang, F.; Liu, Z.; Cao, X.; Xie, Z.; Zou, R.; Senanayake, S.D.; Yang, Y.; Rodriguez, J.A. Highly active Ni/CeO2 catalyst for CO2 methanation: Preparation and characterization. Appl. Catal. B Environ. 2021, 282, 119581. [Google Scholar] [CrossRef]
  31. Tada, S.; Nagase, H.; Fujiwara, N.; Kikuchi, R. What are the best active sites for CO2 methanation over Ni/CeO2? Energy Fuels 2021, 35, 5241–5251. [Google Scholar] [CrossRef]
  32. Lee, S.M.; Lee, Y.H.; Moon, D.H.; Ahn, J.Y.; Nguyen, D.D.; Chang, S.W.; Kim, S.S. Reaction mechanism and catalytic impact of Ni/CeO2–x catalyst for low-temperature CO2 methanation. Ind. Eng. Chem. Res. 2019, 58, 8656–8662. [Google Scholar] [CrossRef]
  33. Atzori, L.; Cutrufello, M.G.; Meloni, D.; Monaci, R.; Cannas, C.; Gazzoli, D.; Sini, M.F.; Deiana, P.; Rombi, E. CO2 methanation on hard-templated NiO-CeO2 mixed oxides. Int. J. Hydrog. Energy 2017, 42, 20689–20702. [Google Scholar] [CrossRef]
  34. Kloetstra, K.; Zandbergen, H.; Van Koten, M.; Van Bekkum, H. Thermoporometry as a new tool in analyzing mesoporous MCM-41 materials. Catal. Lett. 1995, 33, 145–156. [Google Scholar] [CrossRef]
  35. Ocampo, F.; Louis, B.; Roger, A.-C. Methanation of carbon dioxide over nickel-based Ce0.72Zr0.28O2 mixed oxide catalysts prepared by sol–gel method. Appl. Catal. A Gen. 2009, 369, 90–96. [Google Scholar] [CrossRef]
  36. Holzwarth, U.; Gibson, N. The Scherrer equation versus the ‘Debye-Scherrer equation’. Nat. Nanotechnol. 2011, 6, 534. [Google Scholar] [CrossRef]
  37. Kang, W.; Varma, A. Hydrogen generation from hydrous hydrazine over Ni/CeO2 catalysts prepared by solution combustion synthesis. Appl. Catal. B Environ. 2018, 220, 409–416. [Google Scholar] [CrossRef]
  38. Shan, W.; Luo, M.; Ying, P.; Shen, W.; Li, C. Reduction property and catalytic activity of Ce1− XNiXO2 mixed oxide catalysts for CH4 oxidation. Appl. Catal. A Gen. 2003, 246, 1–9. [Google Scholar] [CrossRef]
  39. Liu, H.; Zhao, C.; Wang, L. Mesoporous Ni–CeO2 Catalyst with Enhanced Selectivity and Stability for Reverse Water–Gas Shift Reaction. J. Chem. Eng. Jpn. 2018, 49, 161–165. [Google Scholar] [CrossRef]
Figure 1. (a) N2 adsorption/desorption isotherms and pore size distributions (b) of Ni-CeO2(SSC) and Ni-CeO2(SC) catalysts.
Figure 1. (a) N2 adsorption/desorption isotherms and pore size distributions (b) of Ni-CeO2(SSC) and Ni-CeO2(SC) catalysts.
Crystals 12 00702 g001
Figure 2. XRD patterns of reduced Ni-CeO2(SC) and Ni-CeO2(SSC) catalysts.
Figure 2. XRD patterns of reduced Ni-CeO2(SC) and Ni-CeO2(SSC) catalysts.
Crystals 12 00702 g002
Figure 3. TEM images particle size distributions of reduced Ni-CeO2(SC) (a,b,e) and Ni-CeO2(SSC) (c,d,f) catalysts.
Figure 3. TEM images particle size distributions of reduced Ni-CeO2(SC) (a,b,e) and Ni-CeO2(SSC) (c,d,f) catalysts.
Crystals 12 00702 g003
Figure 4. H2-TPR of Ni-CeO2(SC) and Ni-CeO2(SSC) catalysts.
Figure 4. H2-TPR of Ni-CeO2(SC) and Ni-CeO2(SSC) catalysts.
Crystals 12 00702 g004
Figure 5. CO2 methanation over Ni-CeO2(SC) and Ni-CeO2(SSC) catalysts: (a) CO2 conversion, (b) CH4 selectivity (reaction condition: P = 1.0 bar, WHSV = 300,000 mL g−1 h−1, n(CO2):n(H2):n(N2) = 1:4:5).
Figure 5. CO2 methanation over Ni-CeO2(SC) and Ni-CeO2(SSC) catalysts: (a) CO2 conversion, (b) CH4 selectivity (reaction condition: P = 1.0 bar, WHSV = 300,000 mL g−1 h−1, n(CO2):n(H2):n(N2) = 1:4:5).
Crystals 12 00702 g005
Figure 6. The stability of the Ni-CeO2(SSC) catalyst (reaction condition: T = 350 °C, P = 1.0 bar, WHSV = 300,000 mL g−1 h−1, n(CO2):n(H2):n(N2) = 1:4:5).
Figure 6. The stability of the Ni-CeO2(SSC) catalyst (reaction condition: T = 350 °C, P = 1.0 bar, WHSV = 300,000 mL g−1 h−1, n(CO2):n(H2):n(N2) = 1:4:5).
Crystals 12 00702 g006
Table 1. Textural parameters of Ni-CeO2(SSC) and Ni-CeO2(SC) catalysts.
Table 1. Textural parameters of Ni-CeO2(SSC) and Ni-CeO2(SC) catalysts.
SampleSBET (m2/g)Pore Volume (cm3/g)Pore Size (nm)
Ni-CeO2(SC)11.20.0723.8
Ni-CeO2(SSC)77.50.189.9
Table 2. Crystal sizes and SNi of reduced Ni-CeO2(SC) and Ni-CeO2(SSC) catalysts.
Table 2. Crystal sizes and SNi of reduced Ni-CeO2(SC) and Ni-CeO2(SSC) catalysts.
SamplesNi (nm) aCeO2 (nm) aSNi (m2/gcat) b
Ni-CeO2(SC)21.418.85.87
Ni-CeO2(SSC)7.211.211.45
a Calculated from XRD results based on Scherrer’s equation. b Calculated from H2-TPD results.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Han, C.; Cao, Z.; Yang, J.; Lu, X.; Liu, H.; Jin, Z.; Zhang, Y.; Yang, S.; Zheng, X.; Wang, L. Ni-CeO2 Catalyst with High Ni Loading Prepared by Salt-Assisted Solution Combustion for CO2 Methanation. Crystals 2022, 12, 702. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst12050702

AMA Style

Han C, Cao Z, Yang J, Lu X, Liu H, Jin Z, Zhang Y, Yang S, Zheng X, Wang L. Ni-CeO2 Catalyst with High Ni Loading Prepared by Salt-Assisted Solution Combustion for CO2 Methanation. Crystals. 2022; 12(5):702. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst12050702

Chicago/Turabian Style

Han, Cui, Zhongqi Cao, Jiliang Yang, Xinkang Lu, Hui Liu, Zheyu Jin, Ying Zhang, Shuqing Yang, Xianmin Zheng, and Luhui Wang. 2022. "Ni-CeO2 Catalyst with High Ni Loading Prepared by Salt-Assisted Solution Combustion for CO2 Methanation" Crystals 12, no. 5: 702. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst12050702

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