Next Article in Journal
Crystalline Derivatives of Dipyrazolo-1,5-diazocine and Dipyrazolopyrimidine: A Case of Unexpected Synthesis and Isostructural Polymorphism
Previous Article in Journal
Elucidation of the Conformational Transition of Oligopeptidase B by an Integrative Approach Based on the Combination of X-ray, SAXS, and Essential Dynamics Sampling Simulation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ni-CeO2/SBA-15 Catalyst Prepared by Glycine-Assisted Impregnation Method for Low-Temperature Dry Reforming of Methane

1
Zhejiang Key Laboratory of Petrochemical Pollution Control, Zhejiang Ocean University, Zhoushan 316022, China
2
School of Food and Pharmaceutical, Zhejiang Ocean University, Zhoushan 316022, China
3
Department of Chemical Engineering, School of Petrochemical Engineering & Environment, Zhejiang Ocean University, Zhoushan 316022, China
4
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: 22 April 2022 / Revised: 12 May 2022 / Accepted: 13 May 2022 / Published: 17 May 2022

Abstract

:
Developing low-temperature nickel-based catalysts with good resistance to coking and sintering for dry reforming of methane (DRM) is of great significance. In this work, Ni (5 wt%) and CeO2 (5 wt%) were supported on SBA-15 porous material by glycine-assisted impregnation method to obtain Ni-CeO2/SBA-15-G catalyst. XRD and TEM results showed that the addition of glycine can effectively promote the dispersion of NiO and CeO2 in the pores of SBA-15. H2-TPR and XPS results confirmed the formation of stronger metal-support interaction. In addition, after the addition of glycine, the NixCe1−xOy solid solution content was increased significantly, meanwhile, the Ce3+ concentration was increased from 31% to 49%, accompanied by more oxygen vacancies and generation of active oxygen species. For the above reasons, Ni-CeO2/SBA-15-G had better catalytic performance in the low-temperature DRM test (20 h, 600 °C) with high GHSV (600,000 mL/gcat/h), its CH4 conversion after reaction of 20 h was 2 times that of Ni-CeO2/SBA-15-C catalyst prepared by a conventional impregnation method. TGA-DTA test also proved that Ni-CeO2/SBA-15-G almost completely eliminated carbon deposition. The above advantages of the Ni-CeO2/SBA-15-G catalyst may have originated from the complexation of glycine with metal cations and can prevent them from gathering.

1. Introduction

Due to the over-exploitation and use of fossil fuels, large amounts of methane (CH4) and carbon dioxide (CO2) have been released into the atmosphere in recent decades, causing serious environmental problems [1]. The conversion and utilization of CH4 and CO2 by the CO2 (dry) reforming of CH4 (DRM, CH4 + CO2 ↔ 2H2 + 2CO) as the most feasible way to solve these greenhouse gases has attracted great attention [2] for the following reasons: ① DRM can realize the simultaneous conversion of CH4 and CO2 to syngas (H2 and CO) in one step. ② The theoretical H2/CO ratio of DRM reaction is approximately 1, which is an ideal raw material for Fischer–Tropsch synthesis [2,3,4,5].
The active metal, Ni, is favored by researchers due to its cheapness and stronger C-H bond breaking ability than other transition metals [6]. Due to the strong endothermic feature of the DRM reaction, Ni-based catalysts operating at high temperatures for a long time face two major challenges: carbon deposition and Ni nanoparticles sintering [7,8]. Carbon deposition will destroy the active sites and cover the catalyst surface to hinder mass transfer or even block the reactor completely [9]. Meanwhile, the sintering of Ni particles will significantly reduce the active sites [4], which will lead to the loss of catalytic activity [10]. Great efforts have been made to solve these two problems [11,12,13]. However, most of the reported working temperatures are usually higher than 700 °C; considering the expensive operating cost, researchers tried to improve the thermal conductivity of the reforming processes by microwave-assisted method [14,15], using structured catalysts [16,17], and integrated joule heating [18,19], etc., so as to achieve the purpose of energy saving. Another possible direction is to reduce the working temperature of the DRM reaction. Generally speaking, lowering the working temperature of DRM can alleviate the sintering of Ni nanoparticles but significantly aggravate carbon deposition of the catalyst [20], which is more demanding for Ni-based catalysts.
Research shows that the structure and distribution of Ni nanoparticles on the support can significantly affect the performance of the catalyst [21]. First, highly dispersed Ni nanoparticles can expose more active metal surfaces and metal-support contact interfaces, which are generally considered as active sites for DRM reaction [22]. Second, it has been reported that small-size (<5 nm) Ni nanoparticles can significantly suppress carbon deposition [6,23], which originates from the weakened carbon diffusion rate in Ni crystals. Furthermore, the strong metal-support interaction was shown to effectively reduce the migration and agglomeration of Ni nanoparticles, thereby inhibiting sintering [24]. A variety of materials can be used as supports for Ni-based catalysts, such as Al2O3 [25,26], ZrO2 [27,28], CeO2 [29,30], SiO2 [23,31], hydrotalcite [32,33], and SBA-15 [3,34], etc. Among them, SBA-15 is widely used as a support for Ni-based catalysts because SBA-15 possesses an ordered high-thermal stability mesoporous structure (pore size around 5–30 nm), the high specific area and pore volume, which can limit the migration and agglomeration of Ni particles at high temperature, and increase the high dispersion of Ni nanoparticles, respectively [35]. In addition, CeO2 has attracted much attention because of its special redox (Ce4+ ↔ Ce3+) properties and is widely used as a catalyst support and promoter. Moreover, CeO2 also provides abundant active oxygen and helps the dispersion of active metals, which can not only improve the metal-support interaction, but also can effectively inhibit carbon accumulation by the presence of reactive oxygen species [36,37]. However, only relying on supports and promoters is not enough to greatly limit the sintering and severe carbon deposition of Ni-based catalysts during low-temperature DRM reaction.
The organic acid-assisted impregnation method [31,38] has been reported to prepare highly dispersed Ni-based catalysts. However, to the best of our knowledge, there are no reports on Ni-based catalysts prepared by this method that completely eliminate carbon deposition for low-temperature DRM. In this paper, the catalysts with Ni (5 wt%) and CeO2 (5 wt%) loaded on the SBA-15 support were prepared by using the glycine-assisted impregnation method. The structure of the catalysts was detected by various characterization methods and the catalytic performance for low-temperature DRM reaction was investigated.

2. Materials and Methods

2.1. Catalyst Preparation

Materials: P123 was purchased from Sigma-Aldrich, St. Louis, MO, USA, and all other reagents were purchased from Sinopharm Chemical Reagent Co., Ltd., Shanghai, China.
Preparation of SBA-15 support: First, 16.387 g P123 was dissolved in 2 mol/L hydrochloric acid at 40 °C, stirring until the P123 was completely dissolved, at which point, the solution was light blue. Next, 36.1 mL (34 g) of tetraethyl orthosilicate (TEOS) was added dropwise to the above solution, continually stirring in a 40 °C water bath for 4 h, at which time the solution was milky white. The solution was poured into a hydrothermal kettle and placed in an oven at 100 °C for 48 h. After being cooled to room temperature, the slurry was washed with water and ethanol until no foaming appeared during the suction filtration process. The thus obtained white solid sample was then dried in an oven at 80 °C overnight and calcined at 600 °C (heating rate of 1 °C/min) for 4 h.
Preparation of Ni-CeO2/SBA-15-G and Ni-CeO2/SBA-15-C catalyst: In a typical preparation process, 0.50 g Ni(NO3)2·6H2O, 0.25 g Ce(NO3)3·6H2O, and 0.27 g glycine (n(Glycine)/n(NO3) = 0.7) were dissolved in 3 mL deionized water. After ultrasonic dispersion for 20 min, the above mixture solution was added dropwise to 1.80 g SBA-15 powder, continually stirring for 20 min. Then, the sample stayed for 24 h at room temperature. Finally, the above samples were calcined at 600 °C (heating rate of 1 °C/min) for 4 h, and cooled to room temperature to obtain the Ni-CeO2/SBA-15-G catalyst. The Ni-CeO2/SBA-15-C was prepared by a conventional impregnation method without adding glycine. The theoretical content of Ni and CeO2 in these two catalysts were equal and both were 5 wt%.

2.2. Characterization of Catalysts

X-ray diffraction (XRD) measurements of all catalysts were measured on a DX-2700 diffractometer (Fangyuan Instruments, Wenzhou, China) using Cu Kα (λ = 0.15418 nm) radiation (30 kV, 40 mA) with a scan range of 2θ at 10° to 80° at a rate of 1.2°/min, and the particle size of NiO and CeO2 was calculated by the Scherrer formula. Transmission electron microscopy (TEM) images were acquired on an FEI Tecnai G2 F20 device. X-ray photoelectron spectroscopy (XPS) with a Thermo Scientific K-Alpha spectrometer was employed to detect the chemical states of the surface elements using Al Kα rays (hν = 1486.6 eV, 12 kV, 6 mA). The charge correction was carried out with the binding energy of C1s = 284.80 eV as the standard. The TGA-DTA test of the reacted samples was determined using HCT-1 (Hengjiu Instrument, Beijing, China) with a test temperature range of 20 °C to 800 °C (heating rate of 10 °C/min and with an air flow rate of 30 mL/min). The H2-temperature programmed reduction (H2-TPR) test was carried out in a 5% H2/Ar (30 mL/min) environment with a TP-5080 instrument (Tianjin Xianquan, Tianjin, China) equipped with a thermal conductivity detector (TCD). After in situ pretreatment in Ar (20 mL/min) at 400 °C for 5 min, the catalyst was cooled down to room temperature. Finally, the catalyst was heated to 900 °C at a heating rate of 10 °C/min and the hydrogen consumption curve was recorded.

2.3. Catalytic Test

The DRM reaction performance test of the two samples were carried out in a quartz tube reactor with an inner diameter of 8 mm at 600 °C under atmospheric pressure. For each test, 10 mg of the catalyst was packed on a quartz wool bed. Firstly, the catalyst was pre-reduced by flowing 40 mL/min of H2/N2 (1/3) at 600 °C for 90 min. Then, the reaction was carried out at 600 °C with the mixture gases (100 mL/min) of CH4, CO2, and N2 in a molar ratio of 3:3:14. The feed and product gas streams were analyzed by an online gas chromatograph (GC-7900) equipped with a TDX-01 packed column and a thermal conductivity detector (TCD). The CO2 and CH4 conversion and H2/CO ratio were calculated by the formula as follows:
CO 2   conversion = F ( CO 2 in ) F ( CO 2 out ) F ( CO 2 in )   ×   100 %
CH 4   conversion = F ( CH 4 in ) F ( CH 4 out ) F ( CH 4 in )   ×   100 %
H 2 / CO   mole   ratio = F ( H 2 out ) F ( CO out )
where F refers to the molar flow rate of CO2, CH4, H2, and CO.

3. Results and Discussion

3.1. Characterization of Fresh Catalysts

Figure 1 shows the XRD patterns of the fresh catalysts. The broad peak of the two samples at 20–25° is attributed to amorphous SiO2 (SBA-15) [39]. The Ni-CeO2/SBA-15-C sample showed distinct diffraction peaks at 2θ = 37.2°, 43.3°, 62.9°, and 75.2°, which were attributed to the larger particles of NiO (111), (200), (220), and (311), respectively [37]. In addition, diffraction peaks attributed to CeO2 were recorded at 28.5°, 33.1°, 47.5°, and 56.3° [37]. The crystal sizes of NiO and CeO2, calculated using the Scherrer formula, were 9.5 nm and 5.3 nm (Table 1), respectively. For the Ni-CeO2/SBA-15-G catalyst, only a weak diffraction peak of NiO (200) was observed near 43.3°, where the crystal planes of NiO and the associated presence of Ce species were difficult to identify, implying that NiO and Ce species were highly dispersed on the SBA-15 support. This means that the adding of glycine during the impregnation process can increase the dispersion of NiO and Ce species remarkably. Furthermore, the highly dispersed NiO and Ce species have a better chance to enter into the mesoporous channels of SBA-15, thereby suppressing carbon deposition and sintering during DRM reaction.
In order to observe the microscopic appearance and structure of the catalysts, the samples were characterized by transmission electron microscopy (TEM), and the results are shown in Figure 2. The hexagonal ordered pore structure of a typical SBA-15 molecular sieve (light color) was still very obvious after loading NiO and CeOx, with uniform pore walls (thickness of about 5 nm) and pore size (about 6 nm). As shown in Figure 2a, a large number of NiO and CeOx particles (deep color) of different sizes were observed on the support surface of the Ni-CeO2/SBA-15-C sample, with a particle size distribution ranging from 15 to 30 nm. Figure 2b is a partial enlarged view of Figure 2a, in which CeO2 (111) was observed with particle size of about 5.0 nm, which was consistent with the XRD results. In addition, a large number of NiO (200) crystal planes were observed near the CeO2 (111) crystal plane, indicating that NiO and CeO2 were in close contact, and have a certain interaction, which may also be the reason why the average particle size (23.1 nm) of TEM was larger than the XRD result. In contrast, the Ni-CeO2/SBA-15-G catalyst could not observe obvious NiO particles under the same magnification (Figure 2c,d), but showed a large dark area on the SBA-15 support. A similar phenomenon was also observed by Liu et al. [40] on the Ni/SBA-15 catalyst prepared by cyclodextrin-assisted impregnation method, which was interpreted as the NiO particles were too small to be observed by TEM. It can be speculated that NiO entered the SBA-15 mesoporous channels and was highly dispersed. This means that the particle size of the metallic oxide should be at least smaller than the pore size of the SBA-15 (about 6 nm).
As we speculated, glycine and cyclodextrin may have a similar mechanism of action, that is, metal cations (Ni2+ and Ce3+) were wrapped by glycine, and NO3 was connected to the hydroxyl group of glycine through hydrogen bonds to achieve charge balance. During the catalyst preparation process, since Ni2+ and Ce3+ migrated to the SBA-15 channels together with glycine, Ni2+ and Ce3+ were not easy to aggregate during heat treatment and thus formed highly dispersed NiO and CeOx [40].
The metal-support interaction and reducibility of the catalysts were analyzed using H2-TPR; the results are shown in Figure 3. The reduction profile of the Ni-CeO2/SBA-15-C sample mainly showed a broad asymmetric peak at 340–600 °C, which should be attributed to the reduction of NiO [41]. Among them, the reduction in the lower temperature range (340–500 °C) was mainly due to the bulk NiO located on the surface of SBA-15 or weakly interacting with SBA-15, and this part of NiO dominated the catalyst. The consumption of hydrogen between 500–600 °C was attributed to the reduction of a small amount of NiO entering the mesopores of SBA-15 or the strong interaction with the SBA-15 support [42]. That is to say, the particle size distribution of NiO in catalyst Ni-CeO2/SBA-15-C was uneven, and most of them had a weak metal-support interaction.
After the addition of glycine, the reduction process of the Ni-CeO2/SBA-15-G catalyst was mainly divided into three parts. It is worth noting that a new weak peak appeared between 200–350 °C, which proved that part of the Ni2+ ions entered and distorted the CeO2 lattice, leading to the formation of NixCe1-xOy solid solution, accompanied by the rise of oxygen vacancies and Ce3+ content [36,43]. The broad peak between 350–700 °C was divided into two parts. Similar to Ni-CeO2/SBA-15-C, the small reduction peak in the range of 350–500 °C attributed to the reduction of NiO with weak metal-support interaction. The hydrogen consumption peaks above 500 °C occupied the majority of the total peak area, indicating that most of the NiO in the Ni-CeO2/SBA-15-G catalyst existed in the form of strong metal-support interaction or entered the channels of SBA-15. Thereby, the NiO particles were highly dispersed on the support, which was consistent with the XRD results.
XPS spectroscopy was used to characterize the chemical states of the surface elements. The Ni 2p3/2 region spectrum is shown in Figure 4. For the Ni-CeO2/SBA-15-C catalyst, the typical doublet structure, with shake-up satellite, of NiO is detected at binding energies of 854.4 eV (yellow curve) and 856.4 eV (red curve) [44]. According to the literature [21,45], the Ni 2p3/2 peak centered at 854.4 eV can be attributed to the bulk NiO weakly interacting with the SBA-15 support, and the larger peak area indicated that this NiO species dominated the sample. The characteristic peak of binding energy at 856.4 eV is attributed to a small amount of NiO, which has a strong interaction with the SBA-15 support, and this part of Ni species may be attached to the surface of SBA-15 or the inner wall of the pore in the form of small particles. However, for the Ni-CeO2/SBA-15-G catalyst, the intensity of the peak centered at 856.4 eV of Ni 2p3/2 was larger than that centered at 854.4 eV, suggesting that most Ni species were strongly interacted with SBA-15 support.
The sintering resistance of Ni particles was directly related to the metal-support interaction, which also depended on the Ce3+ concentration to some extent [36]. The XPS spectrum of the Ce 3d of the Ce element on the surface Ce element of the catalyst is shown in Figure 5. Ce 3d spectra can be deconvoluted into ten peaks, containing five bands each for Ce 3d5/2 (u) and Ce 3d3/2 (v). Among them, the u0, u′, u‴, v0, v′, and v‴ peaks (purple curves) corresponded to Ce4+ species, while the u, u″, v, and v″ peaks (red curves) were attributed to the contribution of Ce3+ species [46]. Obviously, both Ni-CeO2/SBA-15-C and Ni-CeO2/SBA-15-G catalysts contained Ce4+ and Ce3+ species, and the relative Ce3+ concentration on the catalyst surface can be determined by integrating the ratio of Ce3+ peak area to the sum of Ce3+ and Ce4+ peak areas (Ce3+/(Ce3+ + Ce4+)) [44], The Ce3+ concentration of Ni-CeO2/SBA-15-C and Ni-CeO2/SBA-15-G were 30.9% and 48.5%, respectively. This result indicated that the addition of glycine can significantly increase the concentration of Ce3+, whereas the high relative content of Ce3+ corresponded to the formation of numerous oxygen vacancies and the spontaneous presence of reactive oxygen species [47,48]. Overall, the XPS and H2-TPR results corresponded well.

3.2. DRM Performances over Catalysts

The low-temperature DRM reaction performance tests were carried out at 600 °C with a space velocity of 6 × 105 mL/gcat/h, CH4/CO2/N2 = 3/3/14, and 1 atm. As displayed in Figure 6, the conversion of CO2 was generally higher than that of CH4 due to the occurrence of reverse water gas shift reaction [1].
Figure 6a presented the CH4 conversion of the catalysts; for the Ni-CeO2/SBA-15-C catalyst, the CH4 conversion dropped from the initial 24.6% to 16.1% within a reaction time of 20 h. In contrast, for the Ni-CeO2/SBA-15-G catalyst, the CH4 conversion decreased from 35.3% to 32.5%, about 2 times that of the Ni-CeO2/SBA-15-C sample at 20 h. It was generally believed that the exposed active metal on the support was the active site for the adsorption and dissociation of CH4 [22]. In addition, it was also reported that the oxygen vacancies provided by CeO2 also contributed to the adsorption and activation of CO2 [49]. For the Ni-CeO2/SBA-15-G catalyst, the XRD and TEM results have demonstrated the existence of highly dispersed Ni nanoparticles. Meanwhile, the H2-TPR and XPS results have confirmed that more oxygen vacancies appeared than Ni-CeO2/SBA-15-C. Thereby, the Ni-CeO2/SBA-15-G catalyst had better adsorption and activation ability of CH4 and CO2, which further exhibited much higher DRM reactivity than the Ni-CeO2/SBA-15-C catalyst. In addition, the Ni-CeO2/SBA-15-G catalyst was more stable than the Ni-CeO2/SBA-15-C catalyst. The specific rate of CH4 of the catalysts were also calculated and compared with those reported in the literature; as shown in Table 2, it can be seen that the Ni-CeO2/SBA-15-G had good catalytic activity for DRM reaction with high GHSV at 600 °C.
The H2/CO ratio is displayed in Figure 6c, and the difference in the H2/CO ratio of the catalysts usually originates from the difference in the active site. The results revealed that the H2/CO ratio of the two catalysts were relatively stable: the Ni-CeO2/SBA-15-C sample was stable at 0.55–0.57, while the H2/CO ratio of the Ni-CeO2/SBA-15-G catalyst was between 0.64 and 0.66. Therefore, the Ni-CeO2/SBA-15-G catalyst was more favorable for the DRM reaction, and suppresses side reactions such as RWGS.

3.3. Characterization of Spent Catalysts

The TGA-DTA test effectively characterized the carbon deposition of the catalyst after the DRM reaction, and the results are shown in Figure 7. As can be seen, the spent Ni-CeO2/SBA-15-C catalyst showed significant weight loss (48.5%) at 540–690 °C, accompanied by an exothermic peak in the DTA curve, indicating that the weight loss originated from the oxidation of carbon deposits. The spent Ni-CeO2/SBA-15-G catalyst showed almost no weight loss in the test, and no exotherm was detected, suggesting that the Ni-CeO2/SBA-15-G catalyst had significant resistance to carbon deposition. The excellent carbon deposition resistance of Ni-CeO2/SBA-15-G catalyst should be attributed to the following two reasons: Firstly, the smaller Ni particles suppressed the coke produced by CH4 cracking. Moreover, the higher Ce3+ and active oxygen concentration of the Ni-CeO2/SBA-15-G catalyst eliminated the coke quickly during the DRM reaction. The excellent resistance to carbon deposition was crucial to the higher stability of the Ni-CeO2/SBA-15-G catalyst for the low-temperature DRM reaction.
The XRD patterns of the spent catalysts are presented in Figure 8. For the spent Ni-CeO2/SBA-15-C, distinct graphite diffraction peaks were observed at 2θ = 26.1° and 42.0° [55,56], which were consistent with the TGA-DTA results. In addition, the peaks of metallic Ni were also shown at 2θ = 44.5° and 51.8° [23]. Compared with the fresh catalyst, the diffraction peaks of CeO2 were broader and weaker, which was consistent with the report of Das et al. [50]. However, no diffraction peaks attributable to Ni, Ce or carbon species were observed in the XRD pattern of the Ni-CeO2/SBA-15-G catalyst after the DRM test, which indicated that the dispersion of Ni and Ce species was too high to be detected. This result meant that the catalyst of Ni-CeO2/SBA-15-G had excellent resistance to sintering and carbon deposition. Since no carbon deposition was formed, the slight loss of activity on Ni-CeO2/SBA-15-G may have originated in the agglomeration of a small amount of Ni particles.

4. Conclusions

In conclusion, the glycine-assisted impregnation method is a simple and effective method to prepare highly active and stable Ni-based catalysts for low-temperature DRM reaction. The characterization results showed that the Ni-CeO2/SBA-15-G catalyst had high Ni dispersion and strong metal-support interaction with SBA-15, and that Ni-CeO2/SBA-15-G showed excellent catalytic performance in the DRM reaction at low temperature (600 °C) for 20 h. Metal sintering was significantly suppressed and carbon deposition was almost completely eliminated. Moreover, XPS and H2-TPR results confirmed that the addition of glycine significantly promoted the generation of Ce3+, accompanied by the generation of more oxygen vacancies and reactive oxygen species, which increased CO2 conversion and carbon removal.

Author Contributions

Investigation, X.L. and C.H.; Writing—original draft, J.Y.; Writing—review and editing, D.G., H.L. 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 the Open Research Subject of Zhejiang Key Laboratory of Petrochemical Environmental Pollution Control (No. 2021Z01).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available upon reasonable request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Marinho, A.L.A.; Toniolo, F.S.; Noronha, F.B.; Epron, F.; Duprez, D.; Bion, N. Highly active and stable Ni dispersed on mesoporous CeO2-Al2O3 catalysts for production of syngas by dry reforming of methane. Appl. Catal. B 2021, 281, 119459. [Google Scholar] [CrossRef]
  2. Zhang, G.; Liu, J.; Xu, Y.; Sun, Y. A review of CH4-CO2 reforming to synthesis gas over Ni-based catalysts in recent years (2010–2017). Int. J. Hydrogen Energy 2018, 43, 15030–15054. [Google Scholar] [CrossRef]
  3. Gálvez, M.E.; Albarazi, A.; da Costa, P. Enhanced catalytic stability through non-conventional synthesis of Ni/SBA-15 for methane dry reforming at low temperatures. Appl. Catal. A 2015, 504, 143–150. [Google Scholar] [CrossRef]
  4. Jing, J.-y.; Wei, Z.-h.; Zhang, Y.-b.; Bai, H.-C.; Li, W.-y. Carbon dioxide reforming of methane over MgO-promoted Ni/SiO2 catalysts with tunable Ni particle size. Catal. Today 2020, 356, 589–596. [Google Scholar] [CrossRef]
  5. Wang, Y.; Yao, L.; Wang, S.; Mao, D.; Hu, C. Low-Temperature catalytic CO2 dry reforming of methane on Ni-based catalysts: A review. Fuel Process. Technol. 2018, 169, 199–206. [Google Scholar] [CrossRef]
  6. Akri, M.; Zhao, S.; Li, X.; Zang, K.; Lee, A.F.; Isaacs, M.A.; Xi, W.; Gangarajula, Y.; Luo, J.; Ren, Y.; et al. Atomically dispersed nickel as coke-resistant active sites for methane dry reforming. Nat. Commun. 2019, 10, 5181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Al-Fatesh, A.S.; Fakeeha, A.H.; Ibrahim, A.A.; Abasaeed, A.E. Ni supported on La2O3 + ZrO2 for dry reforming of methane: The impact of surface adsorbed oxygen species. Int. J. Hydrogen Energy 2021, 46, 3780–3788. [Google Scholar] [CrossRef]
  8. Yang, E.; Nam, E.; Lee, J.; Lee, H.; Park, E.D.; Lim, H.; An, K. Al2O3-Coated Ni/CeO2 nanoparticles as coke-resistant catalyst for dry reforming of methane. Catal. Sci. Technol. 2020, 10, 8283–8294. [Google Scholar] [CrossRef]
  9. De la Cruz-Flores, V.G.; Martinez-Hernandez, A.; Gracia-Pinilla, M.A. Deactivation of Ni-SiO2 catalysts that are synthetized via a modified direct synthesis method during the dry reforming of methane. Appl. Catal. A 2020, 594, 117455. [Google Scholar] [CrossRef]
  10. Huang, F.; Wang, R.; Yang, C.; Driss, H.; Chu, W.; Zhang, H. Catalytic performances of Ni/mesoporous SiO2 catalysts for dry reforming of methane to hydrogen. J. Energy Chem. 2016, 25, 709–719. [Google Scholar] [CrossRef]
  11. Abdullah, B.; Abd Ghani, N.A.; Vo, D.-V.N. Recent advances in dry reforming of methane over Ni-based catalysts. J. Clean. Prod. 2017, 162, 170–185. [Google Scholar] [CrossRef] [Green Version]
  12. Kurlov, A.; Deeva, E.B.; Abdala, P.M.; Lebedev, D.; Tsoukalou, A.; Comas-Vives, A.; Fedorov, A.; Müller, C.R. Exploiting two-dimensional morphology of molybdenum oxycarbide to enable efficient catalytic dry reforming of methane. Nat. Commun. 2020, 11, 4920. [Google Scholar] [CrossRef] [PubMed]
  13. Joo, S.; Kim, K.; Kwon, O.; Oh, J.; Kim, H.J.; Zhang, L.; Zhou, J.; Wang, J.-Q.; Jeong, H.Y.; Han, J.W.; et al. Enhancing Thermocatalytic Activities by Upshifting the d-Band Center of Exsolved Co-Ni-Fe Ternary Alloy Nanoparticles for the Dry Reforming of Methane. Angew. Chem. Int. Ed. 2021, 60, 15912–15919. [Google Scholar] [CrossRef] [PubMed]
  14. Marin, C.M.; Popczun, E.J.; Nguyen-Phan, T.-D.; Tafen, D.N.; Alfonso, D.; Waluyo, I.; Hunt, A.; Kauffman, D.R. Designing perovskite catalysts for controlled active-site exsolution in the microwave dry reforming of methane. Appl. Catal. B 2021, 284, 119711. [Google Scholar] [CrossRef]
  15. Hamzehlouia, S.; Jaffer, S.A.; Chaouki, J. Microwave Heating-Assisted Catalytic Dry Reforming of Methane to Syngas. Sci. Rep. 2018, 8, 8940. [Google Scholar] [CrossRef] [Green Version]
  16. Renda, S.; Cortese, M.; Iervolino, G.; Martino, M.; Meloni, E.; Palma, V. Electrically driven SiC-based structured catalysts for intensified reforming processes. Catal. Today 2022, 383, 31–43. [Google Scholar] [CrossRef]
  17. Meloni, E.; Martino, M.; Ricca, A.; Palma, V. Ultracompact methane steam reforming reactor based on microwaves susceptible structured catalysts for distributed hydrogen production. Int. J. Hydrogen Energy 2021, 46, 13729–13747. [Google Scholar] [CrossRef]
  18. Rieks, M.; Bellinghausen, R.; Kockmann, N.; Mleczko, L. Experimental study of methane dry reforming in an electrically heated reactor. Int. J. Hydrogen Energy 2015, 40, 15940–15951. [Google Scholar] [CrossRef]
  19. Wismann Sebastian, T.; Engbæk Jakob, S.; Vendelbo Søren, B.; Bendixen Flemming, B.; Eriksen Winnie, L.; Aasberg-Petersen, K.; Frandsen, C.; Chorkendorff, I.; Mortensen Peter, M. Electrified methane reforming: A compact approach to greener industrial hydrogen production. Science 2019, 364, 756–759. [Google Scholar] [CrossRef] [Green Version]
  20. Sokolov, S.; Radnik, J.; Schneider, M.; Rodemerck, U. Low-Temperature CO2 reforming of methane over Ni supported on ZnAl mixed metal oxides. Int. J. Hydrogen Energy 2017, 42, 9831–9839. [Google Scholar] [CrossRef]
  21. Károlyi, J.; Németh, M.; Evangelisti, C.; Sáfrán, G.; Schay, Z.; Horváth, A.; Somodi, F. Carbon dioxide reforming of methane over Ni–In/SiO2 catalyst without coke formation. J. Ind. Eng. Chem. 2018, 58, 189–201. [Google Scholar] [CrossRef] [Green Version]
  22. Abdulrasheed, A.; Jalil, A.A.; Gambo, Y.; Ibrahim, M.; Hambali, H.U.; Shahul Hamid, M.Y. A review on catalyst development for dry reforming of methane to syngas: Recent advances. Renew. Sustain. Energy Rev. 2019, 108, 175–193. [Google Scholar] [CrossRef]
  23. Wang, F.; Han, B.; Zhang, L.; Xu, L.; Yu, H.; Shi, W. CO2 reforming with methane over small-sized Ni@SiO2 catalysts with unique features of sintering-free and low carbon. Appl. Catal. B 2018, 235, 26–35. [Google Scholar] [CrossRef]
  24. Wang, F.; Han, K.; Xu, L.; Yu, H.; Shi, W. Ni/SiO2 Catalyst Prepared by Strong Electrostatic Adsorption for a Low-Temperature Methane Dry Reforming Reaction. Ind. Eng. Chem. Res. 2021, 60, 3324–3333. [Google Scholar] [CrossRef]
  25. Da Silva, F.D.A.R.; dos Santos, R.C.R.; Nunes, R.S.; Valentini, A. Role of tin on the electronic properties of Ni/Al2O3 catalyst and its effect over the methane dry reforming reaction. Appl. Catal. A 2021, 618, 118129. [Google Scholar] [CrossRef]
  26. Cichy, M.; Pańczyk, M.; Słowik, G.; Zawadzki, W.; Borowiecki, T. Ni–Re alloy catalysts on Al2O3 for methane dry reforming. Int. J. Hydrogen Energy 2022, 47, 16528–16543. [Google Scholar] [CrossRef]
  27. Zhang, M.; Zhang, J.; Zhang, X.; Zhou, Z.; Song, F.; Zhang, Q.; Tan, Y.; Han, Y. How the reflux treatment stabilizes the metastable structure of ZrO2 and improves the performance of Ni/ZrO2 catalyst for dry reforming of methane? Energy Convers. Manag. 2020, 216, 112950. [Google Scholar] [CrossRef]
  28. Hu, X.; Jia, X.; Zhang, X.; Liu, Y.; Liu, C.-j. Improvement in the activity of Ni/ZrO2 by cold plasma decomposition for dry reforming of methane. Catal. Commun. 2019, 128, 105720. [Google Scholar] [CrossRef]
  29. Liu, Z.; Grinter, D.C.; Lustemberg, P.G.; Nguyen-Phan, T.-D.; Zhou, Y.; Luo, S.; Waluyo, I.; Crumlin, E.J.; Stacchiola, D.J.; Zhou, J.; et al. Dry Reforming of Methane on a Highly-Active Ni-CeO2 Catalyst: Effects of Metal-Support Interactions on C−H Bond Breaking. Angew. Chem. Int. Ed. 2016, 55, 7455–7459. [Google Scholar] [CrossRef] [Green Version]
  30. Tang, Y.; Wei, Y.; Wang, Z.; Zhang, S.; Li, Y.; Nguyen, L.; Li, Y.; Zhou, Y.; Shen, W.; Tao, F.F.; et al. Synergy of Single-Atom Ni1 and Ru1 Sites on CeO2 for Dry Reforming of CH4. J. Am. Chem. Soc. 2019, 141, 7283–7293. [Google Scholar] [CrossRef] [Green Version]
  31. Xu, Y.; Wu, Y.; Li, J.; Wei, S.; Gao, X.; Wang, P. Combustion-Impregnation preparation of Ni/SiO2 catalyst with improved low-temperature activity for CO2 methanation. Int. J. Hydrogen Energy 2021, 46, 20919–20929. [Google Scholar] [CrossRef]
  32. Dębek, R.; Zubek, K.; Motak, M.; Galvez, M.E.; da Costa, P.; Grzybek, T. Ni–Al hydrotalcite-like material as the catalyst precursors for the dry reforming of methane at low temperature. C. R. Chim. 2015, 18, 1205–1210. [Google Scholar] [CrossRef]
  33. Liu, H.; Wierzbicki, D.; Debek, R.; Motak, M.; Grzybek, T.; da Costa, P.; Gálvez, M.E. La-Promoted Ni-hydrotalcite-derived catalysts for dry reforming of methane at low temperatures. Fuel 2016, 182, 8–16. [Google Scholar] [CrossRef]
  34. Omoregbe, O.; Danh, H.T.; Nguyen-Huy, C.; Setiabudi, H.D.; Abidin, S.Z.; Truong, Q.D.; Vo, D.-V.N. Syngas production from methane dry reforming over Ni/SBA-15 catalyst: Effect of operating parameters. Int. J. Hydrogen Energy 2017, 42, 11283–11294. [Google Scholar] [CrossRef]
  35. Panduro, E.A.C.; Granlund, H.; Sztucki, M.; Konovalov, O.; Breiby, D.W.; Gibaud, A. Using Three-Dimensional 3D Grazing-Incidence Small-Angle X-ray Scattering (GISAXS) Analysis To Probe Pore Deformation in Mesoporous Silica Films. ACS Appl. Mater. Interfaces 2014, 6, 2686–2691. [Google Scholar] [CrossRef] [PubMed]
  36. Li, L.; Liu, D.; Guo, Z.; Xi, S.; Chu, W.; Liu, Y. Insights into Ni and (Ce)SBA-15-CTA interaction and syngas formation rate. Mol. Catal. 2021, 514, 111850. [Google Scholar] [CrossRef]
  37. Yan, X.; Hu, T.; Liu, P.; Li, S.; Zhao, B.; Zhang, Q.; Jiao, W.; Chen, S.; Wang, P.; Lu, J.; et al. Highly efficient and stable Ni/CeO2-SiO2 catalyst for dry reforming of methane: Effect of interfacial structure of Ni/CeO2 on SiO2. Appl. Catal. B 2019, 246, 221–231. [Google Scholar] [CrossRef]
  38. Ren, H.-P.; Hao, Q.-Q.; Ding, S.-Y.; Zhao, Y.-Z.; Zhu, M.; Tian, S.-P.; Ma, Q.; Song, W.-Q.; Miao, Z.; Liu, Z.-T. A High-Performance Ni/SiO2 Prepared by the Complexed-Impregnation Method with Citric Acid for Carbon Dioxide Reforming of Methane. Ind. Eng. Chem. Res. 2018, 57, 16257–16263. [Google Scholar] [CrossRef]
  39. Bian, L.; Zhang, L.; Zhu, Z.; Li, Z. Methanation of carbon oxides on Ni/Ce/SBA-15 pretreated with dielectric barrier discharge plasma. Mol. Catal. 2018, 446, 131–139. [Google Scholar] [CrossRef]
  40. Liu, H.; Li, Y.; Wu, H.; Liu, J.; He, D. Effects of α- and γ-cyclodextrin-modified impregnation method on physicochemical properties of Ni/SBA-15 and its catalytic performance in CO2 reforming of methane. Chin. J. Catal. 2015, 36, 283–289. [Google Scholar] [CrossRef]
  41. Yang, Y.; Ochoa-Hernández, C.; Pizarro, P.; de la Peña O’shea, V.A.; Coronado, J.M.; Serrano, D.P. Ce-Promoted Ni/SBA-15 catalysts for anisole hydrotreating under mild conditions. Appl. Catal. B 2016, 197, 206–213. [Google Scholar] [CrossRef]
  42. Tapia-Parada, K.; Valverde-Aguilar, G.; Mantilla, A.; Valenzuela, M.A.; Hernández, E. Synthesis and characterization of Ni/Ce–SiO2 and Co/Ce–TiO2 catalysts for methane decomposition. Fuel 2013, 110, 70–75. [Google Scholar] [CrossRef]
  43. Setiabudi, H.D.; Chong, C.C.; Abed, S.M.; Teh, L.P.; Chin, S.Y. Comparative study of Ni-Ce loading method: Beneficial effect of ultrasonic-assisted impregnation method in CO2 reforming of CH4 over Ni-Ce/SBA-15. J. Environ. Chem. Eng. 2018, 6, 745–753. [Google Scholar] [CrossRef]
  44. Emamdoust, A.; la Parola, V.; Pantaleo, G.; Testa, M.L.; Farjami Shayesteh, S.; Venezia, A.M. Partial oxidation of methane over SiO2 supported Ni and NiCe catalysts. J. Energy Chem. 2020, 47, 1–9. [Google Scholar] [CrossRef]
  45. Yi, H.; Xue, Q.; Lu, S.; Wu, J.; Wang, Y.; Luo, G. Effect of pore structure on Ni/Al2O3 microsphere catalysts for enhanced CO2 methanation. Fuel 2022, 315, 123262. [Google Scholar] [CrossRef]
  46. Ashok, J.; Ang, M.L.; 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]
  47. 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]
  48. Kang, W.; Guo, H.; Varma, A. Noble-metal-free NiCu/CeO2 catalysts for H2 generation from hydrous hydrazine. Appl. Catal. B 2019, 249, 54–62. [Google Scholar] [CrossRef]
  49. Guo, D.; Lu, Y.; Ruan, Y.; Zhao, Y.; Zhao, Y.; Wang, S.; Ma, X. Effects of extrinsic defects originating from the interfacial reaction of CeO2-x-nickel silicate on catalytic performance in methane dry reforming. Appl. Catal. B 2020, 277, 119278. [Google Scholar] [CrossRef]
  50. Das, S.; Ashok, J.; Bian, Z.; Dewangan, N.; Wai, M.H.; Du, Y.; Borgna, A.; Hidajat, K.; Kawi, S. Silica–Ceria sandwiched Ni core–shell catalyst for low temperature dry reforming of biogas: Coke resistance and mechanistic insights. Appl. Catal. B 2018, 230, 220–236. [Google Scholar] [CrossRef]
  51. Albarazi, A.; Beaunier, P.; da Costa, P. Hydrogen and syngas production by methane dry reforming on SBA-15 supported nickel catalysts: On the effect of promotion by Ce0.75Zr0.25O2 mixed oxide. Int. J. Hydrogen Energy 2013, 38, 127–139. [Google Scholar] [CrossRef]
  52. Yao, L.; Zhu, J.; Peng, X.; Tong, D.; Hu, C. Comparative study on the promotion effect of Mn and Zr on the stability of Ni/SiO2 catalyst for CO2 reforming of methane. Int. J. Hydrogen Energy 2013, 38, 7268–7279. [Google Scholar] [CrossRef]
  53. Wang, F.; Han, K.; Yu, W.; Zhao, L.; Wang, Y.; Wang, X.; Yu, H.; Shi, W. Low Temperature CO2 Reforming with Methane Reaction over CeO2-Modified Ni@SiO2 Catalysts. ACS Appl. Mater. Interfaces 2020, 12, 35022–35034. [Google Scholar] [CrossRef] [PubMed]
  54. Li, Z.; Sibudjing, K. Facile Synthesis of Multi-Ni-Core@Ni Phyllosilicate@CeO2 Shell Hollow Spheres with High Oxygen Vacancy Concentration for Dry Reforming of CH4. ChemCatChem 2018, 10, 2994–3001. [Google Scholar] [CrossRef]
  55. Wolfbeisser, A.; Sophiphun, O.; Bernardi, J.; Wittayakun, J.; Föttinger, K.; Rupprechter, G. Methane dry reforming over ceria-zirconia supported Ni catalysts. Catal. Today 2016, 277, 234–245. [Google Scholar] [CrossRef] [Green Version]
  56. Made Joni, I.; Vanith, M.; Camellia, P.; Balasubramanian, N. Augmentation of graphite purity from mineral resources and enhancing % graphitization using microwave irradiation: XRD and Raman studies. Diam. Relat. Mater. 2018, 88, 129–136. [Google Scholar]
Figure 1. XRD patterns of Ni-CeO2/SBA-15-C and Ni-CeO2/SBA-15-G catalysts.
Figure 1. XRD patterns of Ni-CeO2/SBA-15-C and Ni-CeO2/SBA-15-G catalysts.
Crystals 12 00713 g001
Figure 2. TEM images of fresh catalysts: (a,b) Ni-CeO2/SBA-15-C and (c,d) Ni-CeO2/SBA-15-G.
Figure 2. TEM images of fresh catalysts: (a,b) Ni-CeO2/SBA-15-C and (c,d) Ni-CeO2/SBA-15-G.
Crystals 12 00713 g002
Figure 3. H2-TPR profiles of Ni-CeO2/SBA-15-C and Ni-CeO2/SBA-15-G catalysts.
Figure 3. H2-TPR profiles of Ni-CeO2/SBA-15-C and Ni-CeO2/SBA-15-G catalysts.
Crystals 12 00713 g003
Figure 4. Ni 2p3/2 X-ray photoelectron spectra of Ni-CeO2/SBA-15-C and Ni-CeO2/SBA-15-G catalysts.
Figure 4. Ni 2p3/2 X-ray photoelectron spectra of Ni-CeO2/SBA-15-C and Ni-CeO2/SBA-15-G catalysts.
Crystals 12 00713 g004
Figure 5. Ce 3D X-ray photoelectron spectra of Ni-CeO2/SBA-15-C and Ni-CeO2/SBA-15-G catalysts.
Figure 5. Ce 3D X-ray photoelectron spectra of Ni-CeO2/SBA-15-C and Ni-CeO2/SBA-15-G catalysts.
Crystals 12 00713 g005
Figure 6. The catalysts’ stability on stream (a) CH4 conversion; (b) CO2 conversion; (c) H2/CO ratio, with conditions: GHSV = 600,000 mL/gcat/h, 600 °C, 1 atm, and CH4/CO2/N2 (15/15/70) with a flow rate of 100 mL/min.
Figure 6. The catalysts’ stability on stream (a) CH4 conversion; (b) CO2 conversion; (c) H2/CO ratio, with conditions: GHSV = 600,000 mL/gcat/h, 600 °C, 1 atm, and CH4/CO2/N2 (15/15/70) with a flow rate of 100 mL/min.
Crystals 12 00713 g006
Figure 7. TGA-DTA profiles for Ni-CeO2/SBA-15-C and Ni-CeO2/SBA-15-G catalysts.
Figure 7. TGA-DTA profiles for Ni-CeO2/SBA-15-C and Ni-CeO2/SBA-15-G catalysts.
Crystals 12 00713 g007
Figure 8. XRD patterns of spent Ni-CeO2/SBA-15-C and Ni-CeO2/SBA-15-G catalysts.
Figure 8. XRD patterns of spent Ni-CeO2/SBA-15-C and Ni-CeO2/SBA-15-G catalysts.
Crystals 12 00713 g008
Table 1. Textural properties of Ni-CeO2/SBA-15-C and Ni-CeO2/SBA-15-G.
Table 1. Textural properties of Ni-CeO2/SBA-15-C and Ni-CeO2/SBA-15-G.
CatalystsNiO Crystal Size 1 (nm)CeO2 Crystal Size 1 (nm)Ce3+-Ce4+ 2Carbon Deposited 3 (%)
Ni-CeO2/SBA-15-C9.55.331–69%48.5
Ni-CeO2/SBA-15-Gn.d.4n.d. 449–51%n.d. 4
1 Calculated by Scherrer formula from XRD. 2 Measured by XPS. 3 Calculated by TGA-DTA results. 4 Not detected.
Table 2. The CH4 specific rate of Ni-CeO2/SBA-15-C and Ni-CeO2/SBA-15-G compared with the former reported catalysts for DRM reaction at 600 °C.
Table 2. The CH4 specific rate of Ni-CeO2/SBA-15-C and Ni-CeO2/SBA-15-G compared with the former reported catalysts for DRM reaction at 600 °C.
CatalystsGHSV
(mL/gcat/h)
CH4:CO2:N2(Ar)Specific Rate 1
(×10−4 molCH4/gNi/s)
Specific Rate 1
(×10−5 molCH4/gcat/s)
Note
Ni-CeO2/SBA-15-C600,0003:3:1434.917.4This work
Ni-CeO2/SBA-15-G600,0003:3:1470.535.2This work
Ni-SiO2@CeO2200,0003:2:02010.6[50]
Ni/SBA-15-RM20,0001:1:012.19.7[3]
Ni/SBA-1520,0001:1:81.31.3[51]
Ni-Zr/SiO214,4001:1:03.63.4[52]
Ni@SiO2-CeO2-W60,0003:3:41.40.19[53]
Ni@SiO2-CeO2-E60,0003:3:41.10.14[53]
Ni@NiPhy36,0001:1:11.65.1[54]
Ni@NiPhy@CeO236,0001:1:12.14.0[54]
1 The results of this work were calculated at 20 h on catalysts.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yang, J.; Gong, D.; Lu, X.; Han, C.; Liu, H.; Wang, L. Ni-CeO2/SBA-15 Catalyst Prepared by Glycine-Assisted Impregnation Method for Low-Temperature Dry Reforming of Methane. Crystals 2022, 12, 713. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst12050713

AMA Style

Yang J, Gong D, Lu X, Han C, Liu H, Wang L. Ni-CeO2/SBA-15 Catalyst Prepared by Glycine-Assisted Impregnation Method for Low-Temperature Dry Reforming of Methane. Crystals. 2022; 12(5):713. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst12050713

Chicago/Turabian Style

Yang, Jiliang, Dandan Gong, Xinkang Lu, Cui Han, Hui Liu, and Luhui Wang. 2022. "Ni-CeO2/SBA-15 Catalyst Prepared by Glycine-Assisted Impregnation Method for Low-Temperature Dry Reforming of Methane" Crystals 12, no. 5: 713. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst12050713

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