Next Article in Journal
Immobilized Non-Precious Electrocatalysts for Advanced Energy Devices
Next Article in Special Issue
Performance of Modified Alumina-Supported Ruthenium Catalysts in the Reforming of Methane with CO2
Previous Article in Journal
Solar-Energy-Driven Cu-ZnO/TiO2 Nanocomposite Photocatalyst for the Rapid Degradation of Congo Red Azo Dye
Previous Article in Special Issue
Water Gas Shift Reaction Activity on Fe (110): A DFT Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ni/CeO2 Catalyst Prepared via Microimpinging Stream Reactor with High Catalytic Performance for CO2 Dry Reforming Methane

1
School of Light Industry, Beijing Technology and Business University (BTBU), Beijing 100048, China
2
SUSTech Engineering Innovation Center, School of Environmental Science and Engineering, Southern University of Science and Technology, Shenzhen 518055, China
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(6), 606; https://doi.org/10.3390/catal12060606
Submission received: 29 April 2022 / Revised: 26 May 2022 / Accepted: 31 May 2022 / Published: 2 June 2022
(This article belongs to the Special Issue Catalysts in C1 Chemistry)

Abstract

:
Methane reforming with carbon dioxide (DRM) is one promising way to achieve carbon neutrality and convert methane to syngas for high-value chemical production. Catalyst development with better performance is the key to its potential large-scale industrial application due to its deactivation caused by carbon deposition and metal sintering. Hence, a Ni/CeO2 catalyst (Ni/CeO2-M) with higher CO2 conversion and better stability is prepared, supported on CeO2 precipitated via a novel microimpinging stream reactor. A series of ex-situ or in-situ characterizations, such as CO titration measurements, two-step transient surface reaction (two-step TSR), CO2 and CH4 temperature-programmed surface reaction (CO2-TPSR and CH4-TPSR), X-ray absorption fine structure (XAFS), and in-situ Raman spectroscopy study, were used to investigate its structure and mechanism. In contrast to Ni supported on commercial CeO2 (Ni/CeO2-C), the Ni/CeO2-M catalyst with stronger lattice oxygen mobility and higher oxygen storage capacity enhances its CO2 activation ability and carbon deposition. The Ni particle size of the Ni/CeO2-M catalyst decreased, and a higher oxidation state was obtained due to the strong metal–support interaction. Besides the reaction performance improvement of the Ni/CeO2-M catalyst, the novel microimpinging stream reactor could achieve catalyst continuous production with a high preparation efficiency. This work provides a novel method for the high-performance catalyst preparation for DRM reaction and its mechanism study gives a deep insight into high-performance catalyst development via bottom-up study.

1. Introduction

The high concentration of CO2 and CH4, which are identified as the world’s most abundant greenhouse gases (GHG), have caused severe global climate change and ocean acidification [1]. The dry reforming of methane (DRM, Equation (1)) is regarded as one of the most promising routes to mitigate the environmental challenges associated with GHG emissions [2,3]. On one hand, DRM could utilize CO2 and CH4 to reduce GHG emissions. On the other hand, the syngas produced from DRM is the platform chemical for subsequent products via Fischer−Tropsch (FT) synthesis, methanol synthesis [4,5], et al.
CH 4 + CO 2 2 CO + 2 H 2
Despite its considerable environmental potential, DRM is not an industrially mature process. The rapid deactivation of catalysts owing to carbon deposition and metal sintering hinders its potential large-scale industrial application [6,7,8]. Therefore, the catalysts’ development with high performance is still a great challenge for their commercial and industrial implementation. It is generally accepted that CO and H2 are produced via two steps: (i) the C–H bond of CH4 is directly dissociated to form surface carbon intermediates on metal active sites; (ii) the carbon intermediates are oxidized [9,10]. In the process, the oxygen vacancy in the catalyst could effectively assist the CO2 activation and carbon intermediates removal via reversible redox cycling [11,12,13]. Furthermore, the metal particles tend to stabilize with oxygen vacancy to reduce the sintering [14]. Thus, developing a catalyst with abundant oxygen vacancies for better catalytic performance is a logical direction.
CeO2 with high concentrations of oxygen vacancy has been extensively applied to the catalyst support for the dry reforming of methane [15,16,17,18,19,20] and Ni-based catalysts are a promising substitute for precious metal catalysts for DRM. The Ni supported on CeO2 is believed to involve the steps shown in Equations (2)–(4) [21], where the oxygen vacancy involves activating CO2, forming absorbed oxygen species, which could improve the elimination of coke deposition with the assistance of Ni species:
CO 2 + CeO 2 x CO + CeO 2
CH 4 C ( ads ) + 2 H 2
C ( ads ) + CeO 2 CO + CeO 2 x
Thus, the increase in the concentration of oxygen vacancies over the Ni/CeO2 catalyst could greatly enhance the catalytic performance. Previous studies have devoted a lot of efforts to the oxygen vacancy density increasing, such as modification of ceria with aliovalent cations [22,23,24], morphology control [25,26], and solid solutions [12,27], et al. However, most of the strategies are focused on mechanism studies over model catalysts without large-scale application potential. Thus, a simple method is still desired for preparing the high-performance Ni/CeO2 catalyst with a higher density of oxygen vacancies at a large scale for the industrial DRM process.
The microimpinging stream reactor could significantly enhance the micromixing efficiency [28,29,30], easily controlling the crystallization process for the physical and chemical properties modification of the CeO2 support. Hence, the Ni/CeO2 catalyst (labeled as Ni/CeO2-M) prepared supported on CeO2 precipitated via a novel microimpinging stream reactor with higher CO2 conversion and better stability is prepared in this work. A series of characterizations, including X-ray diffraction (XRD), transmission electron microscope (TEM), and CO titration measurements, are carried out for its structural investigation. The CO2 and CH4 temperature-programmed surface reaction (CO2-TPSR and CH4-TPSR), two-step transient surface reaction (two-step TSR), X-ray absorption fine structure (XAFS), and in-situ Raman spectroscopy study, are used to establish the “structure–catalytic performance–reaction mechanism” relationship. This work provides a novel method for high-performance catalyst preparation for DRM reaction and its mechanism study gives a deep insight into high-performance catalyst development via bottom-up study.

2. Results and Discussion

2.1. Structural Characterizations of the Ni/CeO2 Catalysts

The structures of Ni/CeO2 catalysts prepared by different methods are shown in Figure 1a and some structure characterizations of CeO2 (CeO2-M and CeO2-C), including XRD and Raman, are shown in Figure S1. The main diffraction peaks at 2θ = 28.55, 33.08, 47.48, 56.33, 59.09, 69.40, 76.70, and 79.07°, corresponding to (111), (200), (220), (311), (222), (400), (331) and (420) crystal planes, clearly indicate the presence of cubic CeO2 crystal phase (Joint Committee on Powder Diffraction File No. 34-0394). Besides the patterns of CeO2 support, diffractions of NiO appear at 37.29° and 43.28° in the Ni/CeO2 catalysts, indicating that Ni is successfully supported on the CeO2 support. Compared to the Ni/CeO2-C catalyst, the intensity of the CeO2 diffraction peaks of the Ni/CeO2-M catalyst is attenuated, indicating that the CeO2 prepared via the microimpinging stream reactor tends to have a smaller particle size of the CeO2 support and lattice disorder. Figure 1b shows the particle size of Ni/CeO2-C and Ni/CeO2-M catalysts calculated via the Scherrer equation [31]. The average particle size of Ni/CeO2-M catalysts is 109 nm, which is significantly smaller than that of Ni/CeO2-C catalysts (210 nm). Furthermore, the BET specific surface areas of Ni/CeO2-M catalysts are 112 m2/g, increasing 3.1-times more than that of Ni/CeO2-C catalysts (36 m2/g), as shown in Table 1. The high surface area would contribute to the catalytic performance. Generally, the surface area would increase its active site and, consequently, lead to higher activity [32]. Previous researchers found that the high surface area would enhance the oxygen mobility and vacancies generation [33,34]. Compared with the commercial CeO2 support, the microimpinging stream reactor provides a relatively uniform and rapid path for the formation of CeO2 with a larger surface area due to its strong mixing efficiency and high mass transfer performance.
The TEM and EDX characterizations of Ni/CeO2-C and Ni/CeO2-M catalysts are shown in Figure 2. From the TEM images, the Ni/CeO2-M catalyst has a smaller grain size, whereas larger and aggregated particles were obtained in the reduced Ni/CeO2-C catalyst. The higher dispersion of CeO2 support would maximize the Ni-CeO2 interface, which may benefit its reaction performance. From the EDX result, it is clear that the Ni species is more uniformly supported on the Ni/CeO2-M catalyst than that of the Ni/CeO2-C catalyst. In contrast with the large Ni particles over the commercial CeO2, the particle size of the CeO2 is remarkably smaller. Moreover, there are no particles > 15 nm found over the Ni/CeO2-M catalyst, which agrees with the CO chemisorption result in Table 1. The highly dispersed Ni indicates a stronger metal–support interaction over the CeO2 prepared via the microreactor synthesis method, and the increase in metal dispersion would provide more active sites for activation of CH4 and CO2, beneficial to the enhancement of a dry reforming reaction.
To further reveal the chemical and physical properties of the reduced catalysts, Raman spectroscopy study was performed, as shown in Figure 3. In the Raman spectra, the strength of the Raman peak near 460 cm−1 was attributed to that of the first-order F2g peak in CeO2. The OD peak near 570 cm−1 was associated with oxygen vacancies in CeO2 [35,36]. The concentration of oxygen vacancies can be represented by the ratio of IOD/IF2g [37]. As shown in Figure 3a, the intensity of the D peak in the Ni/CeO2-M catalyst is stronger and has a higher ratio (0.469) than that of the Ni/CeO2-C catalyst (0.037), demonstrating that higher concentrations of oxygen vacancies can be obtained by the Ni/CeO2-M catalyst. Furthermore, the presence of Ce3+ in the CeO2 support can be proved by the peak shift of the F2g peak. As shown in Figure 3b, the first-order F2g peak in the Ni/CeO2-M catalyst has lower Raman shifts, which is due to the lattice expansion and mode softening when two Ce4+ ions (0.970 Å) are replaced by two Ce3+ ions (ionic radius 1.143 Å) for oxygen vacancy generation [36,38]. The high concentration of oxygen vacancies in the Ni/CeO2-M catalyst enables the interaction between Ni and CeO2 to drive the metal dispersion via Ni-Ov-CeOx [39], hindering Ni particle sintering, which is expected to improve the stability.

2.2. Reaction Performance

The catalytic performance in DRM of the as-prepared catalysts was evaluated at 800 °C under a relatively high space velocity of 90,000 mL/g/h. In this work, the blank tube and pure CeO2 support with different preparation methods were also tested, as shown in Figure S2. As shown in Figure 4a, the initial CO2 conversion of the Ni/CeO2-M catalyst is 14% higher than that of the Ni/CeO2-C catalyst with higher CH4 conversion of 23% and both of the catalysts could significantly improve their catalytic performance compared with blank tube and pure CeO2 support. As is shown in Figure S3, the apparent activation energy (Ea) of Ni/CeO2-C and Ni/CeO2-M catalysts was measured at low CO2 and CH4 conversions (below 10%) to minimize the limitations of heat and mass transport. The lower apparent activation barriers for both CO2 and CH4 are obtained over the Ni/CeO2-M catalyst, consistent with its higher DRM activity. The Ni species possesses the turnover frequency of CH4 (TOF), ~3.1 s−1, for the DRM over the Ni/CeO2-M catalyst. The catalytic performance based on TOF is compared with previously reported ceria catalysts in the literature for DRM (Table S1). The Ni/CeO2-M catalyst exhibits a good DRM activity compared to the reported Ni-based catalysts. The better catalytic performance may be attributed to the synergistic effect of the higher density of oxygen vacancies and dispersed Ni species on CeO2; Ni species assist the lower dissociation temperature and surface oxygen species promote the conversion of CO2.
The stability of the Ni/CeO2-M catalyst is also excellent, as shown in Figure 4b,c. Under conditions of WHSV = 90,000 mL/g/h at 800 °C, it maintains 86% of CO2 conversion and 90% of CH4 conversion after 50 h. In comparison, the Ni/CeO2-C catalyst loses more than 63% of its CO2 conversion (from 58% to 21%) with the same trends of CH4 conversion (from 69% to 37%) under the same reaction conditions. The better stability indicates that CeO2 prepared via the microimpinging stream reactor could effectively hinder Ni sintering and coke deposition. It has been reported that the stronger interaction over Ni and oxygen vacancy could improve the Ni dispersion and the resistance to sintering under high reaction temperature [12,15,40]. Furthermore, the ratio of H2/CO over the Ni/CeO2-M catalyst is closer to 1, exceeding that of the Ni/CeO2-C catalyst (Figure 4d), indicating the superior ability to suppress the reverse water–gas shift reaction (RWGS, CO 2 + H 2 CO + H 2 O ). Thus, the microimpinging stream reactor could prepare the Ni/CeO2-M catalyst with higher conversion and stability with a higher ratio of H2 to CO.

2.3. Mechanism Investigation

2.3.1. TGA and Raman Spectroscopy Characterizations

The amount of coke deposition in the spent catalysts after the stability test at 800 °C was measured using TGA−DTG (Figure 5a). A weight loss of approximately 3.2% is observed for the used Ni/CeO2-C catalyst in a temperature range of 400−700 °C, with a DTG peak with the maximum weight loss rate [41], indicating the combustion of carbonaceous deposits on the catalyst. On the other hand, the spent Ni/CeO2-M catalyst undergoes no obvious weight loss and no obvious DTG peak was observed. The difference in coke formation on the two catalysts is also confirmed via Raman spectra, as shown in Figure 5b. The vibrational peak centered at around 1345 cm−1 corresponds to the D band, and the G band is at around 1580 cm−1, respectively [42]. The D band and G band mainly contribute to the amorphous carbon and the well-graphitized and ordered carbon species, respectively. The greater intensity in both the D band and G band occur in the spent Ni/CeO2-C catalyst, suggesting a higher amount of coke, which is consistent with the TGA−DTG results. Compared with the well-graphitized and ordered carbon species (G band), the amorphous carbon (D band) tends to be oxidized and removed easily by the oxygen species in the catalyst. A higher peak ratio of ID/IG (1.04) over the Ni/CeO2-M catalyst demonstrates that the amorphous carbon over catalysts is dominant and it would easily be removed with the assistance of oxygen species. Thus, the Ni/CeO2-M catalyst could significantly suppress the coke formation on the catalyst.

2.3.2. CO2 and CH4-TPSR Characterizations

The small amount of carbon deposits are mainly composed of active carbon species formed over the Ni/CeO2-M catalysts. It has been reported that the active carbon intermediates from methane could be oxidized directly by the oxygen vacancies via a Mars−van Krevelen-type redox mechanism [43]. The CO2-TPSR was carried out to investigate the density of oxygen vacancies via monitoring the concentration of CO ( CO 2 + O vac CO ). From the profiles, as shown in Figure 6a, CO production starts at about 193 °C over the Ni/CeO2-M catalyst. Compared with the higher temperature of CO2 dissociation over the Ni/CeO2-C catalyst at 450 °C, the lower temperature indicated that the abundant oxygen vacancy on the Ni/CeO2-M catalyst accelerates the CO2 activation ability. The same trend could be found over the CO2-TPSR of CeO2 in Figure S4a. Furthermore, the higher intensity of CO over Ni/CeO2-M catalyst indicates a higher density of oxygen vacancies for CO2 activation.
Then, the CH4-TPSR experiments were employed to further investigate the coke deposition behaviors (Figure 6b and Figure S4b). During the process of CH4-TPSR, several simultaneous processes occur, namely: (i) generation of carbon species on the catalyst surface as a result of the methane decomposition and (ii) oxidation of these species to CO2, and then to CO due to the mobile oxygen in the lattice of CeO2 [44]. The oxygen vacancies could assist the dissociation of C-H bonds and oxidation of carbon deposits. From Figure 6b, it is clear that the CO peaks of the Ni/CeO2-M catalyst shift to lower temperatures compared with the Ni/CeO2-C catalyst (255 °C to 215 °C), indicating that the high density of surface oxygen species in the Ni/CeO2-M catalyst is more active to the conversion of CH4 to CO, which hinders the carbon deposition and increases its stability.

2.3.3. H2-TPR and XAFS Characterizations

A higher dispersion of Ni over CeO2 prepared via the microimpinging stream reactor, confirmed via TEM and CO chemisorption, was observed. A stronger interaction between Ni and CeO2 is expected to tune the Ni dispersion and electronic state. The H2-TPR profiles, indicating the interaction between metal and support, are shown in Figure 7a. Peaks between 150 and 300 °C are attributed to a reduction in different Ni species: the peak between 150 and 220 °C is assigned to the reduction in free surface NiO and the next peak, around 220–300 °C, is assigned to the reduction in bulk NiO [45,46]. The reduction peak occurring at a higher temperature usually means that it is hard to reduce with stronger metal–support interactions [47,48]. For the Ni/CeO2-M catalysts, stronger metal−support interactions are shown, which verified that the reduction in various Ni species occurred at a higher temperature, as shown in Figure 7a. The stronger interaction would benefit maintaining the Ni dispersion and hindering its agglomeration under 800 °C.
In addition to the H2-TPR experiment, the electronic state of the catalysts is also detected by X-ray absorption near the edge structure (XANES) and extended X-ray absorption fine structure (EXAFS) studies of Ni K-edge with a near air-free method, shown in Figure 7b,c. The Ni K-edge XANES spectra over CeO2 remains in a reduced state. However, a higher Ni2+ content is found over the Ni/CeO2-M catalysts at 64.0%, according to the results of linear combination fitting (LCF, Figure S5). The higher oxidation state of Ni (Niδ+) originates from the electron transfer between Ni and CeO2 support. The higher “white line” of the Ni/CeO2-M catalyst further hints a stronger interaction between the Ni and CeO2 prepared by the microimpinging stream reactor. The Fourier transforms of k3-weighted Ni K extended X-ray absorption fine structure (EXAFS) signals of all the samples are shown in Figure 7c. For both fresh samples, the shapes of the Fourier transformed scattering resemble that of the Ni-Ni shell with some differences in the intensity of the backscattering peaks. The main backscattering peak around 2.2 Å corresponds to a Ni−Ni shell in Ni foil, while the peak around 1.7 Å corresponds to a Ni−O shell and the peak around 2.6 Å is assigned to a Ni-Ni shell in NiO [49]. Over the Ni/CeO2-C and Ni/CeO2-M catalysts, the main peak is aligned with the Ni-Ni shell of Ni foil, indicating the reduced state. The EXAFS has been fitted as shown in Figure 7c and the details are listed in Table 2. The coordination number (CN) of Ni-Ni in the Ni/CeO2-M catalysts (9.80 ± 0.30) is significantly smaller than that of the Ni/CeO2-C catalysts (11.40 ± 0.70). The faintish intensity of Ni-Ni over the Ni/CeO2-M catalysts suggests a smaller Ni particle size, respectively, and in the sample of the Ni/CeO2-M catalyst, the intensity of Ni-O and Ni-Ni (NiO) is stronger than that of the Ni/CeO2-C catalyst, respectively. With the fitting results, the Ni/CeO2-M catalyst has a larger coordination number of Ni-Ni (NiO) of 6.9. Combined with linear combination fitting, the higher oxidation state of Ni (Niδ+) over the Ni/CeO2-M catalyst results from a stronger metal–support interaction [50], contributing to the Ni dispersion and hindering its agglomeration. The same trends could also be found over the Ni-ceria catalysts in the literature for DRM with X-ray Photoelectron Spectroscopy (XPS) characterization, where the peak area assigned oxygen vacancies and Niδ+ is in a stronger intensity [51,52].

2.3.4. Two-Step TSR and In-Situ Raman Spectroscopy Characterizations

As is shown in Figure 8a, two-step TSR experiments were performed to evaluate the DRM reaction behaviors of the Ni/CeO2-C and Ni/CeO2-M catalysts at a reaction temperature of 800 °C [53,54]. After switching the gas from H2/Ar to CH4/Ar, CO is detected immediately for all reduced catalysts when the oxygen vacancies in the catalyst react with the carbon intermediates from CH4 dissociation. The CO peak area of the Ni/CeO2-M catalyst (0.76) is higher by 31% than that of the Ni/CeO2-M catalyst (0.58), as shown in Figure 8b, agreeing with the abundant oxygen vacancies in the Ni/CeO2-M catalyst. When the gas is switched to CO2/Ar, more CO is generated as follows:
C x + x CO 2 2 x CO
The amount of CO could be the descriptor for the CO2 activation ability. It is clear that the Ni/CeO2-M catalyst could produce more CO in Figure 8b. Therefore, the involvement of more surface oxygen species in the DRM reaction could be helpful to the activation of CO2 and the elimination of the surface coke deposition.
To further investigate the reaction mechanism in the catalysts and demonstrate the important roles the oxygen vacancies play in activating CO2 and carbon deposition, in-situ Raman spectroscopy study was performed under different atmospheres, as shown in Figure 9. The peak near 460 cm−1 is assigned to CeO2 F2g mode, and another one near 600 cm−1 is attributed to the oxygen defects. As the results of in-situ Raman reveal, the exposure to H2/Ar causes a downshift in F2g Raman band frequency over the Ni/CeO2-M catalyst (Figure 9a) as follows [55]:
CeO 2 + x H 2 CeO 2 x + x H 2 O
However, there is no shift under H2/Ar over the Ni/CeO2-C catalyst, which indicated that the Ni/CeO2-M catalyst generates more oxygen vacancies than the Ni/CeO2-C catalyst at the same temperature and atmosphere. When the gas is switched to CH4/Ar, the T2g Raman band tends to shift to high frequency, respectively. Furthermore, the D and G bands are generated simultaneously, indicating that the oxygen vacancies participate in the CH4 dissociation over all of the catalysts. For the Ni/CeO2-M catalyst, the D and G bands quickly vanish when the CO2 is introduced to the reactor and oxidize the carbon deposition. However, there are still obvious D and G bands over the Ni/CeO2-C catalyst, indicating a weaker carbon oxidation ability. Based on the results of two-step TSR and in-situ Raman spectroscopy study, more oxygen vacancies could be generated over the Ni/CeO2-M catalyst and more mobile oxygen could be transferred, leading to an enhancement in the activation of CO2 and oxidization of surface coke.

3. Experimental Section

3.1. Catalyst Preparation

The commercial CeO2 support was supplied from Sigma. The improved CeO2 support was prepared by co-precipitation continuous synthesis via a microimpinging stream reactor, as shown in Figure S6. The microimpinging stream reactor system is composed of two pumps (2PB, SZWEICO, Beijing, China) and a mini T-junction (SS-2MTF, XIONGCHUAN, China). The materials of the reactor and tube were 316L stainless steel. The details of the system are shown in Figures S6 and S7 and listed in Table S2.
Typically, 0.01mol Ce(NO3)3·6H2O (Aladdin, Shanghai, China) was dissolved into 250 mL of ethyl alcohol as the precursor solution. Ammonia (28%, Beijing Chemical Plant, Beijing, China) was diluted with deionized water to 0.2 mol/L as the precipitating agent. Two solutions were pumped into the mini T-junction with a flow rate of 100 mL/min in which they could mix homogeneously and produce precipitation. The resultant precipitation was washed with 99.7% ethyl alcohol and centrifugated three times and dried at 100 °C for 12 h. The dried solid was calcined at 500 °C with a heating rate of 10 °C/min for 5 h under air.
The Ni/CeO2-C and Ni/CeO2-M were prepared by the impregnation method with 5 wt% of Ni. The CeO2 support and Ni(NO3)2·6H2O (Alfa Aesar, 98%) were dissolved into ethyl alcohol and vigorously stirred at 80 °C until the ethyl alcohol was evaporated completely. The samples were dried at 100 °C for 12 h and calcined at 400 °C with a heating rate of 10 °C/min for 2 h under air.

3.2. Catalyst Characterization

The catalyst crystal structures were confirmed by an X-ray diffractometer (XRD) with a Cu Kα X-ray source (D8 Advance, Bruker, Billerica, MA, USA). The range of scanning 2θ degrees was from 10° to 80° with a 2 °/min scanning rate. The surface area measurements were analyzed via a N2 adsorption instrument (ASAP 3020, Micromeritics, Norcross (Atlanta), GA, USA) with the BET analysis method. The samples were degassed under a vacuum at 300 °C for 6 h and the BET area was calculated with the region of P/P0 between 0.05 and 0.35. The electron microscope (FEI Tecnai 30, FEI, Houston, TX, USA) equipped with energy-dispersive X-ray spectroscopy (EDX) was used for the morphology investigation of both reduced catalysts. X-ray absorption fine structure (XAFS) of the spent catalyst with a near air-free method as described in Supporting Information was carried out in the fluorescence mode at National Synchrotron Radiation Research Center (NSRRC). The XAFS was analyzed and fitted with the assistance of Artemis software [56]. The absorption edge (K edge) energy was selected at 8333 eV. The R range between 1 and 3 was used for the extended X-ray absorption fine structure (EXAFS) fitting with R-factor below 0.0009.
The H2 temperature-programmed reduction (H2-TPR) and CO chemisorption characterization were conducted using a chemisorption analyzer (AutoChem II 2920, Micromeritics, Norcross (Atlanta), GA, USA) with a thermal conductivity detector (TCD). For the H2-TPR, 50 mg of samples was reduced in the 10% H2/He at 50 mL/min. The temperature was set from 20 to 800 °C with a 10 °C min−1 heating rate. The CO chemisorption characterization was carried out as follows: 100 mg of sample was reduced in 10% H2/He at 50 mL/min flow rate for 2 h at 800 °C. After Ar flow being purged for 10 min, 50 pulses of CO via a 50 μL gas loop were used to evaluate the metal dispersion. The turnover frequency (TOF), Ni dispersion (DNi), and Ni particle size (dNi) are calculated based on the values of CO chemisorption uptakes. The calculation formula is as follows:
TOF = F reactant in × X reactant C CO × m catalyst
D N i = C CO × M Ni m catalyst × w Ni
d N i = 6 M Ni ρ D Ni N a S a
where F reactant in is the flow rate of reactant, mol/s; X reactant is the conversion of the reactant, %; CCO is the CO uptake value of each catalyst based on CO chemisorption, mol CO/g; mcatalyst is the weight of catalyst used, g; wNi is the Ni loading of catalyst, wt%, which is verified by inductively coupled plasma-atomic emission spectrometry (ICP-OES, PerkinElmer 8300, PerkinElmer, Norwalk, CT, USA); MNi is the atomic weight of Ni, 58.69 g/mol; ρ is the Ni density, 8.902 g/cm3; Na is the Avogadro’s number, 6.02 × 1023; Sa is the area of each surface Ni atom, 6.5 × 10−20 m2/atom.
A Raman spectrometer (LabRam HR Evolution, Horiba, Paris, France) with a laser wavelength of 532 nm was used to characterize the structure of the coke on the spent catalysts. The thermogravimetric and differential thermal analysis (TGA/DTG, EVO2, Thermo, Waltham, MA, USA) was performed to evaluate the amount of coke deposited on the spent catalysts. The spent catalysts were firstly degassed with Ar at 200 °C for 30 min. Then the catalysts were heated from 30 to 800 °C at a 10 °C/min heating rate with flowing air.

3.3. Reaction Performance and Mechanism Studies

A continuous fixed-bed flow reactor was used to evaluate the activity of dry reforming of methane (DRM). The catalyst (40 mg, 40–60 mesh) was pretreated at a flow of 10 mL/min H2 at 800 °C for 2 h. After being reduced, the gas flow was adjusted to the required reactants (CH4:CO2:N2 = 15 mL/min:15 mL/min:30 mL/min) and the reaction temperature was raised to 800 °C for 50 h. The reactor outlet gases were analyzed online by the gas chromatograph (2010PLUS, Shimadzu, Kyoto, Japan) equipped with FID and TCD detectors. The conversions of CH4 ( X CH 4 ) and CO2 ( X CO 2 ) were calculated with the mole flow rate at the inlet and outlet of the reactor:
X CH 4 = F CH 4 , in F CH 4 , out F CH 4 , in × 100 %
X CO 2 = F CO 2 , in F CO 2 , out F CO 2 , in × 100 %
The ability to activate CH4 was investigated by CO2-assisted CH4-TPSR. The catalyst was firstly reduced in a 10 mL/min H2 flow at 800 °C for 2 h. After reduction, the catalyst was cooled down to 30 °C, and exposed to 10 mL/min CO2 for 20 min. Then the sample was heated to 800 °C with a heating rate of 5 °C/min at a 25 mL/min 10% CH4/Ar flow. The ability to activate CO2 was evaluated by CO2-TPSR. The catalyst was reduced consistent with the CH4-TPSR. After being cooled to 30 °C, the reduced sample was heated to 750 °C with a 5 °C/min heating rate during the flow of 25 mL/min 20% CO2/Ar. The production of CO2-TPSR and CH4-TPSR were detected by mass spectrometry.
The in-situ Raman spectroscopy study was carried out under reaction conditions. The reactor was equipped from Harrick Scientific Products Inc with different gases (H2, CH4, or CO2). The spectra were obtained with an Argon laser (532 nm) excitation at 4 scans and 30 s acquisition time.
The two-step transient surface reaction (two-step TSR) was used to study the reaction mechanism as shown in Figure S8 [53,54]. The catalyst of 100 mg was firstly reduced in a 10 mL/min H2 flow at 800 °C for 2 h. After purging with Ar flow for 10 min, CH4/Ar was switched into the reactor. For 30 min, the gas was switched from CH4/Ar to CO2/Ar. The production was detected by an FTIR (Nexus 670, Thermo, Waltham, MA, USA) with a transmission cell.

4. Conclusions

A novel microimpinging stream reactor was used to prepare CeO2-M support and the Ni/CeO2-M catalyst could achieve higher CO2 conversion and better stability compared with Ni supported on commercial CeO2 for DRM. The improved conversion and stability of the Ni/CeO2-M catalyst are attributed to the higher density of oxygen vacancies over CeO2-M due to the better micromixing performance. CO2 and CH4 could be effectively activated with the assistance of oxygen vacancies and the carbon deposition could be oxidized to CO, avoiding carbon deposition. Furthermore, Ni could be stabilized and drove the dispersion via Ni-Ov-CeOx, hindering Ni particles sintering. This work provides a simple method for high-performance catalysts preparation for DRM and its mechanism study guides the design of new Ni with high stability.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12060606/s1, Figure S1. (a) XRD patterns and (b) Raman profile of CeO2-C and CeO2-M catalysts. Figure S2. (a) CO2 conversion during 50 h, (b) CH4 conversion during 50 h, and (c) H2/CO ratio during 50 h of the blank tube and pure CeO2 support with different preparation methods. Figure S3. Apparent activation for: (a) CO2 and (b) CH4 of CeO2-C and CeO2-M catalysts. Table S1. Comparisons of TOF for typical Ni-based catalysts. Figure S4. (a) CO2-TPSR profile; and (b) CH4-TPSR profile of CeO2-C and CeO2-M. Figure S5. Linear combination fitting (LCF) of XANES spectra for the reduced catalysts. Figure S6. Scheme picture of microimpinging stream reactor system. Figure S7. Photo and structure of the mini T-junction. Table S2. Details of the microimpinging stream reactor system. Figure S8. Schematic diagram of two-step transient surface reaction. Figure S9. Diagrammatic sketch of the tube with values at the ends. Figure S10. (a) CO2 conversion (b) CH4 conversion, and (c) H2/CO ratio of Ni/CeO2-B and Ni/CeO2-M catalysts. References [21,57,58,59] are cited in the Supplementary Material.

Author Contributions

Formal analysis, writing—original draft preparation, Y.W. (Yadong Wang); conceptualization, supervision, Q.H.; investigation, writing—original draft preparation, X.W.; writing—original draft preparation, Y.H.; writing—review and editing, supervision, Y.W. (Yuanhao Wang); writing—review and editing, supervision, F.W. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by Fundamental Research Funding of Beijing Technology and Business University (PXM2020_014213_000017).

Data Availability Statement

Data available on request from the authors. The data that support the findings of this study are available from the corresponding author, [Yuanhao Wang], upon reasonable request.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Tollefson, J. World’s carbon emissions set to spike by 2% in 2017. Nature 2017, 551, 283. [Google Scholar] [CrossRef] [Green Version]
  2. Bradford, M.; Vannice, M.J.C.R. CO2 reforming of CH4. Catal. Rev. 1999, 41, 1–42. [Google Scholar] [CrossRef]
  3. Song, Y.; Ozdemir, E.; Ramesh, S.; Adishev, A.; Subramanian, S.; Harale, A.; Albuali, M.; Fadhel Bandar, A.; Jamal, A.; Moon, D.; et al. Dry reforming of methane by stable Ni–Mo nanocatalysts on single-crystalline MgO. Science 2020, 367, 777–781. [Google Scholar] [CrossRef] [PubMed]
  4. Gallego, G.S.; Batiot-Dupeyrat, C.; Barrault, J.; Florez, E.; Mondragón, F. Dry reforming of methane over LaNi1−yByO3±δ (B = Mg, Co) perovskites used as catalyst precursor. Appl. Catal. A Gen. 2008, 334, 251–258. [Google Scholar] [CrossRef]
  5. De Medeiros, F.G.; Lopes, F.W.; Rego de Vasconcelos, B. Prospects and Technical Challenges in Hydrogen Production through Dry Reforming of Methane. Catalysts 2022, 12, 363. [Google Scholar] [CrossRef]
  6. Navarro, R.M.; Álvarez-Galván, M.C.; Rosa, F.; Fierro, J.L.G. Hydrogen production by oxidative reforming of hexadecane over Ni and Pt catalysts supported on Ce/La-doped Al2O3. Appl. Catal. A Gen. 2006, 297, 60–72. [Google Scholar] [CrossRef]
  7. Bradford, M.C.J.; Vannice, M.A. Catalytic reforming of methane with carbon dioxide over nickel catalysts II. Reaction kinetics. Appl. Catal. A Gen. 1996, 142, 97–122. [Google Scholar] [CrossRef]
  8. Pakhare, D.; Spivey, J. A review of dry (CO2) reforming of methane over noble metal catalysts. Chem. Soc. Rev. 2014, 43, 7813–7837. [Google Scholar] [CrossRef]
  9. Trimm, D.L. Catalysts for the control of coking during steam reforming. Catal. Today 1999, 49, 3–10. [Google Scholar] [CrossRef]
  10. Djinović, P.; Batista, J.; Pintar, A. Efficient catalytic abatement of greenhouse gases: Methane reforming with CO2 using a novel and thermally stable Rh–CeO2 catalyst. Int. J. Hydrogen Energy 2012, 37, 2699–2707. [Google Scholar] [CrossRef]
  11. Al-Fatesh, A.S.; Arafat, Y.; Kasim, S.O.; Ibrahim, A.A.; Abasaeed, A.E.; Fakeeha, A.H. In situ auto-gasification of coke deposits over a novel Ni-Ce/W-Zr catalyst by sequential generation of oxygen vacancies for remarkably stable syngas production via CO2-reforming of methane. Appl. Catal. B Environ. 2021, 280, 12. [Google Scholar] [CrossRef]
  12. Marinho, A.L.A.; Rabelo-Neto, R.C.; Epron, F.; Bion, N.; Toniolo, F.S.; Noronha, F.B. Embedded Ni nanoparticles in CeZrO2 as stable catalyst for dry reforming of methane. Appl. Catal. B Environ. 2020, 268, 118387. [Google Scholar] [CrossRef]
  13. Lyu, Y.; Jocz, J.; Xu, R.; Stavitski, E.; Sievers, C. Nickel Speciation and Methane Dry Reforming Performance of Ni/CexZr1–xO2 Prepared by Different Synthesis Methods. ACS Catal. 2020, 10, 11235–11252. [Google Scholar] [CrossRef]
  14. Farmer, J.A.; Campbell, C.T. Ceria Maintains Smaller Metal Catalyst Particles by Strong Metal-Support Bonding. Science 2010, 329, 933–936. [Google Scholar] [CrossRef] [PubMed]
  15. Kambolis, A.; Matralis, H.; Trovarelli, A.; Papadopoulou, C. Ni/CeO2-ZrO2 catalysts for the dry reforming of methane. Appl. Catal. A Gen. 2010, 377, 16–26. [Google Scholar] [CrossRef]
  16. Wang, N.; Qian, W.; Chu, W.; Wei, F. Crystal-plane effect of nanoscale CeO2 on the catalytic performance of Ni/CeO2 catalysts for methane dry reforming. Catal. Sci. Technol. 2016, 6, 3594–3605. [Google Scholar] [CrossRef]
  17. 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]
  18. Xue, Y.; Xu, L.; Chen, M.; Wu, C.-e.; Cheng, G.; Wang, N.; Hu, X. Constructing Ni-based confinement catalysts with advanced performances toward the CO2 reforming of CH4: State-of-the-art review and perspectives. Catal. Sci. Technol. 2021, 11, 6344–6368. [Google Scholar] [CrossRef]
  19. Kim, S.B.; Eissa, A.A.; Kim, M.-J.; Goda, E.S.; Youn, J.-R.; Lee, K. Sustainable Synthesis of a Highly Stable and Coke-Free Ni@CeO2 Catalyst for the Efficient Carbon Dioxide Reforming of Methane. Catalysts 2022, 12, 423. [Google Scholar] [CrossRef]
  20. Manan, W.N.; Wan Isahak, W.N.; Yaakob, Z. CeO2-Based Heterogeneous Catalysts in Dry Reforming Methane and Steam Reforming Methane: A Short Review. Catalysts 2022, 12, 452. [Google Scholar] [CrossRef]
  21. Wang, Y.; Zhang, R.; Yan, B. Ni/Ce0.9Eu0.1O1.95 with enhanced coke resistance for dry reforming of methane. J. Catal. 2022, 407, 77–89. [Google Scholar] [CrossRef]
  22. Bueno-López, A.; Such-Basáñez, I.; Salinas-Martínez de Lecea, C. Stabilization of active Rh2O3 species for catalytic decomposition of N2O on La-, Pr-doped CeO2. J. Catal. 2006, 244, 102–112. [Google Scholar] [CrossRef]
  23. Rangaswamy, A.; Sudarsanam, P.; Reddy, B.M. Rare earth metal doped CeO2-based catalytic materials for diesel soot oxidation at lower temperatures. J. Rare Earths 2015, 33, 1162–1169. [Google Scholar] [CrossRef]
  24. Luisetto, I.; Tuti, S.; Romano, C.; Boaro, M.; Di Bartolomeo, E.; Kesavan, J.K.; Kumar, S.S.; Selvakumar, K. Dry reforming of methane over Ni supported on doped CeO2: New insight on the role of dopants for CO2 activation. J. CO2 Util. 2019, 30, 63–78. [Google Scholar] [CrossRef]
  25. Meng, F.; Fan, Z.; Zhang, C.; Hu, Y.; Guan, T.; Li, A. Morphology-Controlled Synthesis of CeO2 Microstructures and Their Room Temperature Ferromagnetism. J. Mater. Sci. Technol. 2017, 33, 444–451. [Google Scholar] [CrossRef]
  26. Wang, L.; Yu, Y.; He, H.; Zhang, Y.; Qin, X.; Wang, B. Oxygen vacancy clusters essential for the catalytic activity of CeO2 nanocubes for o-xylene oxidation. Sci. Rep. 2017, 7, 12845. [Google Scholar] [CrossRef]
  27. Niu, G.; Hildebrandt, E.; Schubert, M.A.; Boscherini, F.; Zoellner, M.H.; Alff, L.; Walczyk, D.; Zaumseil, P.; Costina, I.; Wilkens, H.; et al. Oxygen Vacancy Induced Room Temperature Ferromagnetism in Pr-Doped CeO2 Thin Films on Silicon. ACS Appl. Mater. Interfaces 2014, 6, 17496–17505. [Google Scholar] [CrossRef]
  28. Zhang, Q.-C.; Cheng, K.-P.; Wen, L.-X.; Guo, K.; Chen, J.-F. A study on the precipitating and aging processes of CuO/ZnO/Al2O3 catalysts synthesized in micro-impinging stream reactors. RSC Adv. 2016, 6, 33611–33621. [Google Scholar] [CrossRef]
  29. Zhang, Q.-C.; Liu, Z.-W.; Zhu, X.-H.; Wen, L.-X.; Zhu, Q.-F.; Guo, K.; Chen, J.-F. Application of microimpinging stream reactors in the preparation of CuO/ZnO/Al2O3 catalysts for methanol synthesis. Ind. Eng. Chem. Res. 2015, 54, 8874–8882. [Google Scholar] [CrossRef]
  30. Gu, R.; Cheng, K.; Wen, L. Application of the engulfment model in assessing micromixing time of a micro-impinging stream reactor based on the determination of impinging zone with CFD. Chem. Eng. J. 2021, 409, 128248. [Google Scholar] [CrossRef]
  31. Holzwarth, U.; Gibson, N. The Scherrer equation versus the ‘Debye-Scherrer equation’. Nat. Nanotechnol. 2011, 6, 534. [Google Scholar] [CrossRef] [PubMed]
  32. Li, H.-F.; Zhang, N.; Chen, P.; Luo, M.-F.; Lu, J.-Q. High surface area Au/CeO2 catalysts for low temperature formaldehyde oxidation. Appl. Catal. B Environ. 2011, 110, 279–285. [Google Scholar] [CrossRef]
  33. Galkin, A.A.; Kostyuk, B.G.; Lunin, V.V.; Poliakoff, M. Continuous Reactions in Supercritical Water: A New Route to La2CuO4 with a High Surface Area and Enhanced Oxygen Mobility. Angew. Chem. Int. Ed. 2000, 39, 2738–2740. [Google Scholar] [CrossRef]
  34. López, J.M.; Gilbank, A.L.; García, T.; Solsona, B.; Agouram, S.; Torrente-Murciano, L. The prevalence of surface oxygen vacancies over the mobility of bulk oxygen in nanostructured ceria for the total toluene oxidation. Appl. Catal. B Environ. 2015, 174, 403–412. [Google Scholar] [CrossRef] [Green Version]
  35. Taniguchi, T.; Watanabe, T.; Sugiyama, N.; Subramani, A.; Wagata, H.; Matsushita, N.; Yoshimura, M. Identifying defects in ceria-based nanocrystals by UV resonance Raman spectroscopy. J. Phys. Chem. C 2009, 113, 19789–19793. [Google Scholar] [CrossRef]
  36. Lee, Y.; He, G.; Akey, A.J.; Si, R.; Flytzani-Stephanopoulos, M.; Herman, I.P. Raman analysis of mode softening in nanoparticle CeO2-δ and Au-CeO2-δ during CO oxidation. J. Am. Chem. Soc. 2011, 133, 12952–12955. [Google Scholar] [CrossRef]
  37. Guo, M.; Lu, J.; Wu, Y.; Wang, Y.; Luo, M. UV and Visible Raman Studies of Oxygen Vacancies in Rare-Earth-Doped Ceria. Langmuir 2011, 27, 3872–3877. [Google Scholar] [CrossRef]
  38. Hong, S.J.; Virkar, A.V. Lattice Parameters and Densities of Rare-Earth Oxide Doped Ceria Electrolytes. J. Am. Ceram. Soc. 1995, 78, 433–439. [Google Scholar] [CrossRef]
  39. Xiao, Z.; Li, Y.; Hou, F.; Wu, C.; Pan, L.; Zou, J.; Wang, L.; Zhang, X.; Liu, G.; Li, G. Engineering oxygen vacancies and nickel dispersion on CeO2 by Pr doping for highly stable ethanol steam reforming. Appl. Catal. B Environ. 2019, 258, 117940. [Google Scholar] [CrossRef]
  40. Li, S.; Gong, J. Strategies for improving the performance and stability of Ni-based catalysts for reforming reactions. Chem. Soc. Rev. 2014, 43, 7245–7256. [Google Scholar] [CrossRef]
  41. Hirata, T.; Kashiwagi, T.; Brown, J.E. Thermal and oxidative degradation of poly(methyl methacrylate): Weight loss. Macromolecules 1985, 18, 1410–1418. [Google Scholar] [CrossRef]
  42. Kim, S.M.; Abdala, P.M.; Margossian, T.; Hosseini, D.; Foppa, L.; Armutlulu, A.; van Beek, W.; Comas-Vives, A.; Copéret, C.; Müller, C. Cooperativity and Dynamics Increase the Performance of NiFe Dry Reforming Catalysts. J. Am. Chem. Soc. 2017, 139, 1937–1949. [Google Scholar] [CrossRef] [PubMed]
  43. Das, S.; Bhattar, S.; Liu, L.; Wang, Z.; Xi, S.; Spivey, J.J.; Kawi, S. Effect of Partial Fe Substitution in La0.9Sr0.1NiO3 Perovskite-Derived Catalysts on the Reaction Mechanism of Methane Dry Reforming. ACS Catal. 2020, 10, 12466–12486. [Google Scholar] [CrossRef]
  44. Zakrzewski, M.; Shtyka, O.; Ciesielski, R.; Kedziora, A.; Maniukiewicz, W.; Arcab, N.; Maniecki, T. Effect of Ruthenium and Cerium Oxide (IV) Promotors on the Removal of Carbon Deposit Formed during the Mixed Methane Reforming Process. Materials 2021, 14, 7581. [Google Scholar] [CrossRef] [PubMed]
  45. Guo, J.; Lou, H.; Zhao, H.; Chai, D.; Zheng, X. Dry reforming of methane over nickel catalysts supported on magnesium aluminate spinels. Appl. Catal. A Gen. 2004, 273, 75–82. [Google Scholar] [CrossRef]
  46. Akiki, E.; Akiki, D.; Italiano, C.; Vita, A.; Abbas-Ghaleb, R.; Chlala, D.; Drago Ferrante, G.; Laganà, M.; Pino, L.; Specchia, S. Production of hydrogen by methane dry reforming: A study on the effect of cerium and lanthanum on Ni/MgAl2O4 catalyst performance. Int. J. Hydrogen Energy 2020, 45, 21392–21408. [Google Scholar] [CrossRef]
  47. Parastaev, A.; Muravev, V.; Huertas Osta, E.; van Hoof, A.J.F.; Kimpel, T.F.; Kosinov, N.; Hensen, E.J.M. Boosting CO2 hydrogenation via size-dependent metal–support interactions in cobalt/ceria-based catalysts. Nat. Catal. 2020, 3, 526–533. [Google Scholar] [CrossRef]
  48. Lin, S.S.Y.; Daimon, H.; Ha, S.Y. Co/CeO2-ZrO2 catalysts prepared by impregnation and coprecipitation for ethanol steam reforming. Appl. Catal. A Gen. 2009, 366, 252–261. [Google Scholar] [CrossRef]
  49. Zhao, Y.; Zhao, B.; Liu, J.; Chen, G.; Gao, R.; Yao, S.; Li, M.; Zhang, Q.; Gu, L.; Xie, J.; et al. Oxide-Modified Nickel Photocatalysts for the Production of Hydrocarbons in Visible Light. Angew. Chem. Int. Ed. 2016, 55, 4215–4219. [Google Scholar] [CrossRef]
  50. Gonzalez-DelaCruz, V.M.; Holgado, J.P.; Pereñíguez, R.; Caballero, A. Morphology changes induced by strong metal–support interaction on a Ni–ceria catalytic system. J. Catal. 2008, 257, 307–314. [Google Scholar] [CrossRef]
  51. Anchieta, C.G.; Assaf, E.M.; Assaf, J.M. Syngas production by methane tri-reforming: Effect of Ni/CeO2 synthesis method on oxygen vacancies and coke formation. J. CO2 Util. 2022, 56, 101853. [Google Scholar] [CrossRef]
  52. Zhang, F.; Liu, Z.; Chen, X.; Rui, N.; Betancourt, L.E.; Lin, L.; Xu, W.; Sun, C.-j.; Abeykoon, A.M.M.; Rodriguez, J.A.; et al. Effects of Zr Doping into Ceria for the Dry Reforming of Methane over Ni/CeZrO2 Catalysts: In Situ Studies with XRD, XAFS, and AP-XPS. ACS Catal. 2020, 10, 3274–3284. [Google Scholar] [CrossRef]
  53. Zhang, J.; Wang, Y.; Tian, J.; Yan, B. Cu/LaFeO3 as an efficient and stable catalyst for CO2 reduction: Exploring synergistic effect between Cu and LaFeO3. AIChE J. 2022, 68, e17640. [Google Scholar] [CrossRef]
  54. Zhang, X.; Liu, Y.; Zhang, M.; Yu, T.; Chen, B.; Xu, Y.; Crocker, M.; Zhu, X.; Zhu, Y.; Wang, R.; et al. Synergy between β-Mo2C Nanorods and Non-thermal Plasma for Selective CO2 Reduction to CO. Chemisty 2020, 6, 3312–3328. [Google Scholar] [CrossRef]
  55. Silva, I.d.C.; Sigoli, F.A.; Mazali, I.O. Reversible Oxygen Vacancy Generation on Pure CeO2 Nanorods Evaluated by in Situ Raman Spectroscopy. J. Phys. Chem. C 2017, 121, 12928–12935. [Google Scholar] [CrossRef]
  56. Ravel, B.; Newville, M. Athena, Artemis, Hephaestus: Data analysis for X-ray absorption spectroscopy using IFEFFIT. J. Synchrotron Radiat. 2005, 12, 537–541. [Google Scholar] [CrossRef] [Green Version]
  57. Gonzalez-Delacruz, V.M.; Ternero, F.; Pereñíguez, R.; Caballero, A.; Holgado, J.P. Study of nanostructured Ni/CeO2 catalysts prepared by combustion synthesis in dry reforming of me-thane. Appl. Catal. A Gen. 2010, 384, 1–9. [Google Scholar] [CrossRef]
  58. Li, X.; Phornphimon, M.; Zhang, X.; Deng, J.; Zhang, D. Promoting Dry Reforming of Methane Catalysed by Atomically-Dispersed Ni over Ceria-Upgraded Boron Nitride. Chem. Asian J. 2022, 17, e202101428. [Google Scholar] [CrossRef]
  59. Zeng, F.; Zhang, J.; Xu, R.; Zhang, R.; Ge, J. Highly dispersed Ni/MgO-mSiO2 catalysts with excellent activity and stability for dry reforming of methane. Nano Res. 2022, 1–10. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns and (b) particle size of Ni/CeO2-C and Ni/CeO2-M catalysts.
Figure 1. (a) XRD patterns and (b) particle size of Ni/CeO2-C and Ni/CeO2-M catalysts.
Catalysts 12 00606 g001
Figure 2. TEM and EDX images of (a) reduced Ni/CeO2-C catalyst and (b) reduced Ni/CeO2-M catalyst.
Figure 2. TEM and EDX images of (a) reduced Ni/CeO2-C catalyst and (b) reduced Ni/CeO2-M catalyst.
Catalysts 12 00606 g002
Figure 3. (a) Raman profile and (b) enlarged profile of reduced Ni/CeO2-C and reduced Ni/CeO2-M catalysts.
Figure 3. (a) Raman profile and (b) enlarged profile of reduced Ni/CeO2-C and reduced Ni/CeO2-M catalysts.
Catalysts 12 00606 g003
Figure 4. (a) Catalytic performance comparation; (b) CO2 conversion during 50 h, (c) CH4 conversion during 50 h, and (d) H2/CO ratio during 50 h of Ni/CeO2-C and Ni/CeO2-M catalysts.
Figure 4. (a) Catalytic performance comparation; (b) CO2 conversion during 50 h, (c) CH4 conversion during 50 h, and (d) H2/CO ratio during 50 h of Ni/CeO2-C and Ni/CeO2-M catalysts.
Catalysts 12 00606 g004
Figure 5. (a) TGA−DTG result and (b) Raman spectra of Ni/CeO2-C and Ni/CeO2-M catalysts.
Figure 5. (a) TGA−DTG result and (b) Raman spectra of Ni/CeO2-C and Ni/CeO2-M catalysts.
Catalysts 12 00606 g005
Figure 6. (a) CO2-TPSR profileand (b) CH4-TPSR profile of Ni/CeO2-C and Ni/CeO2-M catalysts.
Figure 6. (a) CO2-TPSR profileand (b) CH4-TPSR profile of Ni/CeO2-C and Ni/CeO2-M catalysts.
Catalysts 12 00606 g006
Figure 7. (a) TPR profile; (b) XANES and (c) EXAFS of spent Ni/CeO2-C and spent Ni/CeO2-M catalysts.
Figure 7. (a) TPR profile; (b) XANES and (c) EXAFS of spent Ni/CeO2-C and spent Ni/CeO2-M catalysts.
Catalysts 12 00606 g007
Figure 8. (a) Two-step TSR profile; (b) peak area of CO of Ni/CeO2-C and Ni/CeO2-M catalysts.
Figure 8. (a) Two-step TSR profile; (b) peak area of CO of Ni/CeO2-C and Ni/CeO2-M catalysts.
Catalysts 12 00606 g008
Figure 9. In-situ Raman spectroscopy study result of (a) Ni/CeO2-M catalysts, and (b) Ni/CeO2-C catalysts.
Figure 9. In-situ Raman spectroscopy study result of (a) Ni/CeO2-M catalysts, and (b) Ni/CeO2-C catalysts.
Catalysts 12 00606 g009
Table 1. Characteristics of Ni/CeO2-C and Ni/CeO2-M catalysts.
Table 1. Characteristics of Ni/CeO2-C and Ni/CeO2-M catalysts.
CatalystsNi Loading (wt%) aSBET (m2/g) bNi Dispersion (DNi) (%) cNi Particle Size (dNi) (nm) d
Ni/CeO2-C4.96364.819.7
Ni/CeO2-M4.921128.911.3
a Determined via ICP. b Determined via N2 adsorption. c,d Determined via the CO chemisorption method.
Table 2. Fitting details of Ni/CeO2-C and Ni/CeO2-M catalysts.
Table 2. Fitting details of Ni/CeO2-C and Ni/CeO2-M catalysts.
SamplesNi-Ni (Metallic Ni)Ni-Ni (NiO)Ni-O (NiO)
R (Å)CNR (Å)CNR (Å)CN
Ni2.49 ± 0.00412.0----
NiO--2.95 ± 0.0312.02.09 ± 0.036.0
Ni/CeO2-M2.47 ± 0.019.80 ± 0.302.98 ± 0.0026.90 ± 1.102.12 ± 0.014.90 ± 0.40
Ni/CeO2-C2.48 ± 0.00111.40 ± 0.703.05 ± 0.070.30 ± 0.202.14 ± 0.042.20 ± 0.40
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, Y.; Hu, Q.; Wang, X.; Huang, Y.; Wang, Y.; Wang, F. Ni/CeO2 Catalyst Prepared via Microimpinging Stream Reactor with High Catalytic Performance for CO2 Dry Reforming Methane. Catalysts 2022, 12, 606. https://doi.org/10.3390/catal12060606

AMA Style

Wang Y, Hu Q, Wang X, Huang Y, Wang Y, Wang F. Ni/CeO2 Catalyst Prepared via Microimpinging Stream Reactor with High Catalytic Performance for CO2 Dry Reforming Methane. Catalysts. 2022; 12(6):606. https://doi.org/10.3390/catal12060606

Chicago/Turabian Style

Wang, Yadong, Qing Hu, Ximing Wang, Yanpeng Huang, Yuanhao Wang, and Fenghuan Wang. 2022. "Ni/CeO2 Catalyst Prepared via Microimpinging Stream Reactor with High Catalytic Performance for CO2 Dry Reforming Methane" Catalysts 12, no. 6: 606. https://doi.org/10.3390/catal12060606

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