Next Article in Journal
Aerobic Methanol Oxidation over Unsupported Nanoporous Gold: The Influence of an Added Base
Previous Article in Journal
Catalyst Performance in the HDPE Pyrolysis-Reforming under Reaction-Regeneration Cycles
Previous Article in Special Issue
Structure-Sensitivity of CO2 Methanation over Nanostructured Ni Supported on CeO2 Nanorods
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

CO and CO2 Co-Methanation on Ni/CeO2-ZrO2 Soft-Templated Catalysts

Dipartimento di Scienze Chimiche e Geologiche, Università di Cagliari, 09124 Cagliari, Italy
*
Author to whom correspondence should be addressed.
Catalysts 2019, 9(5), 415; https://doi.org/10.3390/catal9050415
Submission received: 30 March 2019 / Revised: 20 April 2019 / Accepted: 1 May 2019 / Published: 2 May 2019
(This article belongs to the Special Issue Supported Catalysts for Carbon Oxides Methanation)

Abstract

:
Supported nickel catalysts were synthesized, characterized, and employed in the carbon oxides co-methanation process. Five NiO/CeO2-ZrO2 mixed oxides, with the same Ni content and different Ce/Zr molar ratios, were prepared by the soft-template method. They were characterized through ICP-AES, N2 adsorption, XRD, and TPR. Reduced Ni/CeO2-ZrO2 catalysts were obtained by submitting the oxide systems to reduction treatment in H2 at 400 °C. They were characterized by XRD, H2-TPD, and CO2 adsorption microcalorimetry and their catalytic performances in the carbon oxides co-methanation were investigated. Catalytic tests were performed in a fixed-bed continuous-flow microreactor at atmospheric pressure. The effect of experimental conditions (reaction temperature, space velocity, reactants molar ratio) was also studied. Almost complete CO conversion was obtained on any catalyst, whereas CO2 conversion was much lower and increased with Ce content, at least up to Ce/Zr = 1. The beneficial effect of the Ce content could be related to the increased NiO reducibility and to the higher ability to adsorb and activate CO2. However, at high Ce/Zr ratios, it is probably counterbalanced by an interplay of reactions involving CO and CO2.

1. Introduction

Due to the depleting nature of oil and natural gas reserves, the production of a valuable fuel such as synthetic natural gas (SNG) has been attracting increasing attention. SNG can be obtained by hydrogenation of CO and/or CO2 [1]. When syngas or hydrogen is obtained from renewable resources, carbon oxides methanation presents the additional advantage of reducing greenhouse gases emissions. Generally, reformed gases deriving from the steam reforming of oil fractions or from coal gasification contain both CO and CO2. Thus, carbon oxides can be simultaneously hydrogenated to methane (co-methanation). Several studies have been devoted to the methanation of either CO [2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20] or CO2 [5,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42] on different supported metal catalysts. In some cases methanation of syngas containing both carbon oxides has been investigated with the aim of selectively hydrogenating CO [7,10,13,17,18,19]. However, a still limited number of papers deal with catalytic co-methanation of carbon oxides [43,44,45,46,47,48,49,50,51,52].
Supported nickel-based catalysts are generally very active and selective in carbon oxides hydrogenation reactions; the additional advantage of being reasonably inexpensive makes them promising alternatives to noble metals-based catalysts (e.g., supported Ru or Rh). As for the co-methanation process, several oxidic systems have been used as a support for nickel in combination with another metal [45,47,51,52]. For supporting undoped nickel, besides Al2O3 [47,51], CeO2-Al2O3 [50], and Gd2O3-CeO2 [52] systems, zirconia—pure [43,46] or in combination with other oxides [43,46,48]—has been studied. When compared to pure zirconia, the presence of a second oxide has usually proved to be beneficial, especially for the thermal stability of the catalytic system [43,46]. On the other hand, when zirconia has been used in combination with ceria [48], only one composition has been studied, and no comparison with the pure oxides has been carried out. Recently, in the present authors’ laboratory a series of undoped nickel catalysts, supported on pure and mixed cerium and zirconium oxides, has been synthesized and tested in the CO2 hydrogenation to SNG, indicating the role of the CeO2 component of the support in activating CO2 [42].
In the present work a series of nanostructured NiO-CeO2-ZrO2 systems with different Ce/Zr ratios—including pure ZrO2 and pure CeO2—has been prepared by the unconventional soft-template synthetic method [53]. The mixed oxides have been characterized as for their structural, textural, and redox properties. The supported nickel catalysts have been obtained by in situ reduction of the synthesized systems. Their structure and their H2 and CO2 adsorption properties have also been studied and related to their catalytic performance in the carbon oxides co-methanation process, for which the effect of both the support composition and the reaction conditions (reaction temperature, space velocity, reactants molar ratio) was investigated.

2. Results

2.1. Characterization of Fresh NiO-CeO2-ZrO2 Mixed Oxides

NiO-CeO2-ZrO2 mixed oxides with a nominal NiO content of 30 wt% and different Ce/Zr molar ratios were synthesized by means of the soft-template method and will hereafter be referred to as NiO/CexZr1−xO2, where x is the nominal CeO2 molar fraction in the (CeO2)x-(ZrO2)1−x support. Their chemical composition was determined by inductively coupled plasma atomic emission spectroscopy (ICP-AES). Data reported in Table 1 show that the experimental composition is close to the nominal one for all the samples.
Textural analysis was carried out by determining the nitrogen adsorption-desorption isotherms at −196 °C, which are shown in Figure 1. All the samples are characterized by a type IVa N2 adsorption isotherm [54]; however, increasing Zr content results in a higher microporous contribution (at low p/p0) and a more linear rise at high relative pressure (p/p0 > 0.6). The pore size distribution curves, obtained by applying the BJH method to the isotherm adsorption branch [54,55], are displayed in the insets in Figure 1. They show a general decrease in pore size with increasing Zr content. In particular, at low Ce contents in the (CeO2)x-(ZrO2)1−x support (x ≤ 0.25) the contribution of pores wider than 10 nm becomes negligible. Specific surface area calculated using the BET equation (SBET) and specific pore volume (Vp) are reported in Table 1. The pure zirconia-supported sample is characterized by a high surface area, which generally decreases with increasing Ce content.
The X-ray diffraction (XRD) patterns of the fresh NiO-CeO2-ZrO2 systems are shown in Figure 2. In all samples, NiO crystalline phase (PDF Card 441159) can be identified, and from the peak width (by means of the Scherrer equation [56]) an average crystallite size of ca. 4 nm can be calculated. In the XRD patterns of NiO/ZrO2 and NiO/CeO2, signals ascribable to the corresponding pure oxide supports are also present. However, while the typical signals of nanocrystalline cubic CeO2 phase (PDF Card 750151) can be identified, from which an average crystalline size of ca. 3 nm can be calculated, only one broad band ascribable to amorphous zirconia can be observed. In all the samples containing both Ce and Zr, besides the pure oxides (nanocrystalline ceria and amorphous zirconia), a ceria-zirconia solid solution is possibly present. In fact, signals corresponding to a cubic CeO2-like phase are visible. With increasing Zr content, they become wider and shift towards higher angles, suggesting the (at least partial) incorporation of the small zirconium ions into the ceria lattice, which would lead to the formation of CexZr1−xO2 solid solutions characterized by smaller crystallite size and lattice parameter.

2.2. Reducibility of NiO-CeO2-ZrO2 Mixed Oxides And Characterization of the Reduced Samples

Reducibility of NiO-CeO2-ZrO2 mixed oxides was investigated by means of temperature-programmed reduction (TPR) analyses, whose results are reported in Figure 3. The TPR profile of NiO/ZrO2 shows a single signal (maximum at ca. 590 °C) which—taking into account that soft-templated pure zirconia did not give rise to any significant hydrogen consumption [42]—can be ascribed to the reduction of the NiO species present in the sample. However, when compared with the position of the maximum (ca. 500 °C) in the TPR profile of a soft-templated pure NiO sample [42], such signal appears shifted to higher temperature, thus indicating the presence of strong ZrO2-NiO interactions, which make the nickel oxide species more difficult to reduce. The TPR curves of the cerium-containing samples show a low-temperature (150–350 °C) small signal and a major one at higher temperatures. Both signals result from the overlapping of different contributions, which indicates the presence of different reducible species. Although also CeO2 species can be reduced, most hydrogen consumption is expected to be due to the reduction of NiO species [42]. The features of the NiO/CeO2 TPR profile can be described in the light of the literature on NiO-CeO2 systems [19,42,57,58,59]. The low-temperature feature can be ascribed to the loss of oxygen species adsorbed on defective sites at the NiO-CeO2 interface, where Ce and Ni are in close interaction. The high-temperature signal clearly results from the partial overlapping of at least two contributions, which can be assigned to the reduction of nickel oxide particles dispersed on the ceria surface and to NiO species strongly interacting with CeO2. Contributions similar to those distinguished for the pure ceria-supported sample are also present in the TPR curves of the systems containing both Ce and Zr, although with different relative areas.
The synthetized systems were also characterized after being submitted to the same reduction pretreatment that they undergo prior to the catalytic tests (flowing H2 at 400 °C for 1 h). The reduced samples will hereafter be referred to as Ni/CexZr1−xO2. The XRD patterns of the hydrogen-treated systems are reported in Figure 4. As expected, the signals ascribable to the oxidic supports (amorphous zirconia, cubic ceria, and cubic solid solutions) are the same as those of the fresh NiO/CexZr1−xO2 samples (cf. Figure 2). For all the reduced systems, signals of Ni0 phase (PDF Card 040850) are present. The metallic nickel average crystal size, calculated by means of the Scherrer equation, is ca. 6 nm for all the Ce-containing samples, and lower than 5 nm for Ni/ZrO2. Such small values indicate that the strong interaction with the support can hinder the sintering of the metal nanocrystals. The asymmetry of the main signals in the Ni/ZrO2 pattern is probably due to some unreduced NiO, whose presence also in the other samples cannot be excluded.
The H2 adsorption properties where studied through hydrogen temperature-programmed desorption (H2-TPD) experiments. It resulted to be 55 μmol g−1 for Ni/ZrO2 and to progressively increase with the Ce content up to 115 μmol g−1 for Ni/CeO2. Probably the simple adsorption on the metallic Ni0 surface is not the only cause for the total amount of hydrogen adsorbed, a contribution of spillover being expected. However, since the nickel content and the size of metal nanocrystals are similar for all the samples, the increasing trend of hydrogen adsorption with Ce content could mainly be explained by an increase in the NiO reduction degree at increasing Ce content. This would be in accordance with the TPR results and would imply that, even though no signals are clearly detectable in the XRD patterns, NiO phase is still present in the hydrogen-treated samples.
CO2 adsorption properties of the Ni/CexZr1−xO2 samples were studied through adsorption microcalorimetry. The differential heat of adsorption (Qdiff) is an indication of the strength of the adsorbent-adsorbate interaction, whereas CO2 uptake is a measure of the number of adsorbing sites. Therefore, reporting Qdiff as a function of CO2 uptake (Figure 5) simultaneous information on the concentration of the adsorbing sites and on their strength distribution is obtained. Whatever the sample, strong sites (Qdiff ≥ 150 kJ mol−1) are present in low concentration. As CO2 uptake increases Qdiff decreases down to 50 kJ mol−1 (below which adsorption cannot be considered specific). For Ni/ZrO2 such decrease is very steep and the total amount of CO2 adsorbed is ca. 35 μmol g−1. On the other hand, Ni/CeO2 is able to adsorb a much higher amount of CO2 (ca. 230 μmol g−1), most of which on medium-strength sites (75 kJ mol−1Qdiff ≤ 150 kJ mol−1). These results, which are in agreement with those obtained on the pure oxidic phases [42], indicate the superior ability of CeO2 – in comparison with ZrO2 – in adsorbing and activating CO2. The systems containing both Ce and Zr show an intermediate behavior, in terms of both concentration of adsorbing sites and strength of interaction. The general enhancement in the adsorption of CO2 at increasing Ce content is not followed by Ni/Ce0.50Zr0.50O2 and Ni/Ce0.75Zr0.25O2, whose curves practically overlap. However, it has to be taken into account that the adsorption capacity also depends on the surface area, which for the reduced samples is expected to follow the same trend observed for the parent fresh oxide samples (Table 1). Compared to the sample with the same content of Ce and Zr, the Ce-richer sample is most probably characterized by a significantly lower surface area, which would counterbalance the higher capacity of the ceria component to adsorb CO2.

2.3. CO and CO2 Co-Methanation Catalytic Tests

Simultaneous CO and CO2 hydrogenation to SNG was carried out at 300 °C and atmospheric pressure on the Ni/CexZr1−xO2 samples, obtained in the reduced form by in situ pretreatment (flowing H2 at 400 °C for 1 h). A reactant gas mixture CO/CO2/H2 of molar composition of 1/1/5 was fed with a space velocity (SV) of 150,000 cm3 h−1 gcat−1. The results, in terms of CO and CO2 conversions and methane selectivity, are summarized in Table 2. Since, during the 6-h runs, catalytic activity was found to be stable with time-on-stream (t.o.s) for all the samples, average values are reported.
Irrespective of the catalyst composition, a CH4 selectivity higher than 99 mol% is always observed. All the catalysts exhibit very high CO conversion, which seems to be slightly favored by high Ce contents. CO2 conversion is much lower and the beneficial effect of Ce is apparent up to an equimolar content of Ce and Zr, a further enrichment in Ce not producing any significant effect.
On Ni/Ce0.50Zr0.50O2, a long-time run at variable temperature was carried out, to study the catalytic stability under thermal stress. As shown in Figure 6, both CO and CO2 conversions decrease at increasing temperature, because of the exothermic character of both the CO and CO2 methanation reactions. However, when the reaction temperature is set back at 300 °C, the catalytic activity is restored, indicating a very good thermal stability of the catalyst. Methane selectivity (not reported) was always higher than 99 mol%.
On Ni/Ce0.50Zr0.50O2 the influence of both feed composition and space velocity was also investigated, by performing co-methanation at 300 °C with two different feed compositions (CO/CO2/H2 = 1/1/5 or 1/1/7) and three different space velocities (SV = 50,000, 150,000, or 450,000 cm3 h−1 gcat−1). Results reported in Figure 7 show that when a higher H2 concentration is used CO2 conversion significantly increases, whereas only a very slight improvement is observed for the already very high CO conversion. Whatever the feed composition, both XCO and XCO2 decrease at increasing SV. However, whereas only small decreases in CO conversion are observed, the effect of SV on CO2 conversion is particularly important, indicating that kinetic limitations are far more significant for CO2 methanation. Noteworthy, for XCO2 the relative decrease is practically the same (ca. 37%) with either feed compositions. Also in these runs, methane selectivity (not reported) was always higher than 99 mol%.

3. Discussion

The NiO-CeO2-ZrO2 mixed oxides samples synthetized by the soft-template method are characterized by high surface areas and nanometric sizes of pores and crystallites. The strong interaction between the dispersed NiO phase and the ZrO2 support makes the NiO nanoparticles reduction more difficult than for the unsupported nickel oxide. When Ce is present different reducible species are formed and reducibility generally increases. XRD results indicate the reduction of NiO nanocrystals under the experimental pretreatment conditions. However, H2-TPD experiments suggest that such reduction is not complete, though favored by increasing Ce content.
In the carbon oxides co-methanation, a competition between CO and CO2 hydrogenation reactions occurs. In the catalytic runs an almost complete CO conversion is obtained, whereas CO2 is converted to a much smaller extent. The prevailing hydrogenation of CO rather than of CO2 has already been reported [48,52] and is supported by thermodynamics [1]. Both reactions (Equations (1) and (2) in Scheme 1) are reversible and exothermic. Although both equilibrium constants decrease with temperature, at 300 °C still quite high values can be calculated, especially for CO hydrogenation [1].
On the present catalysts, CO hydrogenation is probably favored also from a kinetic point of view. In the co-methanation process, both the dispersed metallic nickel and the oxide support play a role. H2 and CO are adsorbed and activated on metallic Ni0 species [8]. On the other hand, adsorption and activation of CO2 occur on the oxide support, through the intermediate formation of either CO or carbonate and hydrogencarbonate species [27]. Since CO is activated on the same sites as H2, its hydrogenation is likely to be favored over that of CO2, which requires the migration of adsorbed hydrogen from the metallic to the oxide phase (hydrogen spillover).
It has also to be taken into account that several other reactions might take place in the carbon oxides co-methanation system [1]. For the present catalytic tests, calculation of methane selectivity and carbon balance allow excluding the occurrence of reactions leading to C2+ hydrocarbons or to carbon deposits. However, CO and CO2 are possibly involved also in reverse methane dry reforming and water-gas shift reactions (Equations (3) and (4) in Scheme 1). Although they are reversible reactions, CO conversion into CO2 is thermodynamically favored; in particular, the reaction of reverse methane dry reforming is characterized by an equilibrium constant higher than those of the methanation reactions [1]. The combination of these reactions, or of the corresponding reverse reactions, with CO2 or CO hydrogenation reactions would in the end lead to the production of methane. However, their occurrence needs to be considered when comparing the almost complete CO conversion and the low values obtained for CO2 conversion. Most likely, not only CO2 hydrogenation is repressed by the competitive CO hydrogenation, but CO2 is also formed from CO. This is particularly true when the feed composition in the co-methanation runs is CO/CO2/H2 = 1/1/5, i.e., hydrogen concentration is lower than the stoichiometric amount. When hydrogen is fed in higher concentration (CO/CO2/H2 = 1/1/7; i.e., stoichiometric amount) CO2 conversion significantly increases. However it does not reach the equilibrium value of 93 mol% [1], indicating the occurrence of kinetic limitations, as also suggested by the significant effect of SV.
The general beneficial effect of the Ce content on the catalytic activity of the present samples can be partly related to the increased reducibility of NiO species and the consequently higher amount of hydrogen adsorbed and activated. However, the role of the ceria component itself has to be taken into account. Besides favoring CO activation on metallic Ni0 [5,20], ceria is the main responsible for CO2 adsorption and activation, as confirmed by the microcalorimetric results. CO2 would preferentially adsorb on the ceria component of the support, most probably forming first carbonate and hydrogencarbonate species, which would then be hydrogenated to formates and finally to methane by the hydrogen previously adsorbed and activated by the metallic Ni0 nanoparticles [27,36,38,42]. Therefore, the beneficial effect of the Ce content in the hydrogenation of CO2 is most likely due also to the superior ability of CeO2 in activating CO2. However, a Ce molar fraction x > 0.50 in the support does not produce any further increase in XCO2. In the light of the above description, this could be explained by the balance of different effects. As Ce content increases, increasing amounts of both CO2 and H2 can be adsorbed and activated. Therefore, increasing the Ce content up to Ce/Zr = 1 results in an increase in CO2 conversion. A possible explanation for the observed behavior at Ce molar fraction x > 0.50 could be that the greater amount of activated hydrogen would favor the reverse methane dry reforming reaction. Thus, at high Ce contents the increased capacity in activating CO2 would be counterbalanced by the increased amount of CO2 produced through the reverse methane dry reforming reaction, resulting in a practically unchanged overall XCO2. Further experiments should be performed in order to confirm such hypothesis.
Moreover, optimization of the co-methanation system, with the aim of obtaining higher CH4 yields, could be achieved by analyzing the effect of the catalyst composition at different experimental conditions and better understanding the role of the catalytic active phases in the complex interplay of the different reactions involving carbon oxides.

4. Materials and Methods

4.1. Synthesis of Materials

NiO-CeO2-ZrO2 mixed oxides were synthesized by means of the soft-template method [53,60], using cetyl-trimethyl-ammonium bromide (CTAB, Aldrich, St. Louis, MI, USA, ≥98%) as templating agent, Ni(NO3)2·6H2O (Aldrich, 99.999%), Ce(NO3)3·6H2O (Aldrich, St. Louis, MI, USA, 99%), and ZrO(NO3)2·6H2O (Aldrich, St. Louis, MI, USA, 99.999%) as precursors, and NaOH (Aldrich, St. Louis, MI, USA, 97%) as precipitating agent. For the synthesis, appropriate amounts of the template and the nitrate precursors (CTAB/precursors: 0.62 mol mol−1; total precursors concentration: 0.048 M) were dissolved in distilled water at room temperature under stirring. After 30 min, a 0.15 M solution of NaOH was added dropwise until a pH value of 13 was reached; the mixture was then stirred for 15 h. After digestion at 90 °C for 3 h, the resulting solid was separated by filtration and washed with hot water (70 °C). It was then treated at 60 °C for 1.5 h, ground, dried at 110 °C for 6 h, and finally calcined at 450 °C for 4 h.

4.2. Characterization

Inductively coupled plasma atomic emission spectroscopy (ICP-AES) analyses were performed with a 5110 ICP-OES spectrophotometer (Agilent Technologies, Santa Clara, CA, USA) to determine the Ni, Ce, and Zr contents. Sample solutions were obtained by treating the solids with a H2O2 (35%) – HNO3 (70%) mixture (1:1 by volume), stirring at 80 °C for 2 h, adding a HCl (37%) – HNO3 (70%) mixture (3:1 by volume), and finally—after 16 h at RT—diluting with Milli-Q water.
Textural analysis was carried out with an ASAP 2020 system (Micromeritics), by determining the nitrogen adsorption-desorption isotherms at −196 °C. Before analysis, the sample was heated overnight under vacuum up to 250 °C (heating rate, 1 °C min−1). Surface area values were calculated by the BET equation. The pore size distribution profiles were determined by applying the BJH method to the isotherm adsorption branch [55,56].
Structural characteristics of the fresh and reduced samples were investigated by X-ray diffraction (XRD) using a X3000 diffractometer (Seifert) with θ-θ Bragg-Brentano geometry with Cu-Kα wavelength and a graphite monochromator before the detector. The average crystallite sizes were estimated by the Scherrer equation [56].
The studies on the reducibility of NiO-CeO2-ZrO2 mixed oxides and on the hydrogen adsorption properties of the reduced samples were performed by means of temperature-programmed reduction (TPR) and hydrogen temperature-programmed desorption (H2-TPD) analyses, respectively, by using a TPD/R/O 1100 apparatus (ThermoQuest, Waltham, MA, USA) equipped with a thermal conductivity detector (TCD). Prior to the TPR experiment, the sample (typically 0.020 g) was pretreated in air (15 cm3 min−1) at 450 °C for 1 h; reduction profiles were then recorded under flowing H2 (5 vol% in N2; flow rate, 30 cm3 min−1) while heating (20 °C min−1) from 40 to 950 °C. As for H2-TPD analyses, the sample (ca. 0.100 g) was first reduced with pure H2 (15 cm3 min−1) at 400 °C for 1 h. Then, after purge with N2 at 400 °C for 1 h, pulses of H2 (5 vol% in N2) were admitted in the reactor at 50 °C until the area of the peaks was observed constant; finally, the desorption profile was collected under flowing N2 (20 cm3 min−1) while heating (10 °C min−1) from 40 to 500 °C.
CO2 adsorption properties of the reduced samples were studied through adsorption microcalorimetry, using a Tian-Calvet heat flow calorimeter (Setaram, Caluire, France) connected to a volumetric vacuum line. Prior the analysis, each sample (ca. 0.150 g, 40 mesh) was reduced with pure H2 at 400 °C for 1 h; then it was pretreated overnight at 250 °C under vacuum (10−3 Pa) before adsorption. CO2 adsorption was carried out at 80 °C by admitting successive doses of the probe gas; for each dose, the equilibrium pressure, the amount of gas adsorbed, and the corresponding thermal effect were recorded. The run was stopped at the final pressure of 133.3 Pa.

4.3. CO and CO2 Co-Methanation Catalytic Tests

Simultaneous CO and CO2 hydrogenation to SNG was carried out in a tubular (i.d. = 0.8 cm) quartz-glass fixed-bed continuous-flow microreactor at atmospheric pressure. Prior to the reaction, the appropriate amount of catalyst powder was placed inside the reactor and pretreated in situ: It was first treated in flowing air (30 cm3 min−1) at 400 °C overnight; then, after purging in He (60 cm3 min−1), it was reduced under H2 flow (15 cm3 min−1) at the same temperature for 1 h, and finally cooled to the reaction temperature in flowing He (60 cm3 min−1). The whole series of Ni/CexZr1−xO2 catalysts was tested for 6 h at 300 °C, with a reactant gas mixture CO/CO2/H2 with molar composition of 1/1/5 (CO, 10 mol%; CO2, 10 mol%; H2, 50 mol%; balance N2, used as internal standard) and space velocity (SV) 150,000 cm3 h−1 gcat−1.
On Ni/Ce0.50Zr0.50O2 additional runs were performed, also at different temperature (400 or 500 °C), with different CO/CO2/H2 molar composition (CO/CO2/H2 = 1/1/7: CO, 8 mol%; CO2, 8 mol%; H2, 56 mol%; balance N2, used as internal standard), and/or at different space velocities (50,000 or 450,000 cm3 h−1 gcat−1).
For each run, after the first hour on stream, on-line analysis of the reactor effluent was performed every hour (after removing water through an ice trap and a 3A molecular sieves trap, where no significant CO2 adsorption proved to occur, placed between the reactor outlet and the injection valve) with a GC 6890 (Agilent), equipped with a HP Poraplot Q capillary column and a TCD. The results of the quantitative analysis of the carbon-containing components were used for checking the carbon mass balance and for calculating CO and CO2 conversion and products selectivity.

5. Conclusions

NiO-CeO2-ZrO2 mixed oxides with narrow mesopores and high specific surface areas were prepared by the soft-template method. NiO was dispersed in the form of small crystallites on the CeO2-ZrO2 supports and its reducibility increased with the Ce content. Upon in situ reduction, active catalysts for the carbon oxides co-methanation were obtained. Whereas CO was almost completely converted, CO2 conversion was much lower and resulted to depend on the catalyst composition. The increase in CO2 conversion with Ce content up to Ce/Zr = 1 could be explained with the higher ability of the CeO2 component of the support to adsorb and activate CO2 and to promote NiO reduction, thus also favoring H2 adsorption and activation. However, at high Ce/Zr ratios, such effects are probably counterbalanced by the competition among the reactions involving H2, CO, and CO2.

Author Contributions

Conceptualization, L.A., E.R., and M.G.C.; Investigation, L.A., D.M., and M.F.S.; Resources, R.M.; Data curation, L.A.; Writing—Original draft preparation, M.G.C.; Writing—review and editing, L.A., E.R., and M.G.C.; Visualization, L.A. and M.G.C.; Supervision, E.R.; Project administration, E.R.

Funding

This research received no external funding.

Acknowledgments

The authors thank Fondazione di Sardegna (Progetti Biennali di Ateneo, Annualità 2017; CUP: F71I17000170002).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gao, J.; Wang, Y.; Ping, Y.; Hu, D.; Xu, G.; Gu, F.; Su, F. A thermodynamic analysis of methanation reactions of carbon oxides for the production of synthetic natural gas. RSC Adv. 2012, 2, 2358–2368. [Google Scholar] [CrossRef]
  2. Ishihara, T.; Eguchi, K.; Arai, H. Hydrogenation of Carbon Monoxide over SiO2-Supported Fe-Co, Co-Ni and Ni-Fe Bimetallic Catalysts. Appl. Catal. 1987, 30, 225–238. [Google Scholar] [CrossRef]
  3. Aksoylu, A.E.; Önsan, Z.İ. Interaction between Nickel and Molybdenum in Ni-Mo/Al2O3 catalysts: II. CO hydrogenation. Appl. Catal. A Gen. 1998, 168, 399–407. [Google Scholar] [CrossRef]
  4. Xavier, K.O.; Sreekala, R.; Rashid, K.; Yusuff, K.; Sen, B. Doping effects of cerium oxide on Ni/Al2O3 catalysts for methanation. Catal. Today 1999, 49, 17–21. [Google Scholar] [CrossRef]
  5. Znak, L.; Stołecki, K.; Zieliński, J. The effect of cerium; lanthanum and zirconium on nickel/alumina catalysts for the hydrogenation of carbon oxides. Catal. Today 2005, 101, 65–71. [Google Scholar] [CrossRef]
  6. Kustov, A.L.; Frey, A.M.; Larsen, K.E.; Johannessen, T.; Nørskov, J.K.; Christensen, C.H. CO methanation over supported bimetallic Ni-Fe catalysts: From computational studies towards catalyst optimization. Appl. Catal. A Gen. 2007, 320, 98–104. [Google Scholar] [CrossRef]
  7. Jimenez, V.; Sanchez, P.; Panagiotopoulou, P.; Valverde, J.L.; Romero, A. Methanation of CO, CO2 and selective methanation of CO, in mixtures of CO and CO2, over ruthenium carbon nanofibers catalysts. Appl. Catal. A Gen. 2010, 390, 35–44. [Google Scholar] [CrossRef]
  8. Senanayake, S.D.; Evans, J.; Agnoli, S.; Barrio, L.; Chen, T.-L.; Hrbek, J.; Rodriguez, J.A. Water-Gas Shift and CO Methanation Reactions over Ni-CeO2(111) Catalysts. Top Catal. 2011, 54, 34–41. [Google Scholar] [CrossRef]
  9. Hu, D.; Gao, J.; Ping, Y.; Jia, L.; Gunawan, P.; Zhong, Z.; Xu, G.; Gu, F.; Su, F. Enhanced Investigation of CO Methanation over Ni/Al2O3 Catalysts for Synthetic Natural Gas Production. Ind. Eng. Chem. Res. 2012, 51, 4875–4886. [Google Scholar] [CrossRef]
  10. Zhao, A.; Ying, W.; Zhang, H.; Ma, H.; Fang, D. Ni-Al2O3 catalysts prepared by solution combustion method for syngas methanation. Catal. Commun. 2012, 17, 34–38. [Google Scholar] [CrossRef]
  11. Hwang, S.; Lee, J.; Hong, U.G.; Jung, J.C.; Koh, D.J.; Lim, H.; Byun, C.; Song, I.K. Hydrogenation of carbon monoxide to methane over mesoporous nickel-M-alumina (M = Fe, Ni, Co, Ce, and La) xerogel catalysts. J. Ind. Eng. Chem. 2012, 18, 243–248. [Google Scholar] [CrossRef]
  12. Yan, X.; Liu, Y.; Zhao, B.; Wang, Z.; Wang, Y.; Liu, C.-J. Methanation over Ni/SiO2: Effect of the catalyst preparation methodologies. Int. J. Hydrogen Energy 2013, 38, 2283–2291. [Google Scholar] [CrossRef]
  13. Zyryanova, M.M.; Snytnikov, P.V.; Gulyaev, R.V.; Amosov, Y.I.; Boronin, A.I.; Sobyanin, V.A. Performance of Ni/CeO2 catalysts for selective CO methanation in hydrogen-rich gas. Chem. Eng. J. 2014, 238, 189–197. [Google Scholar] [CrossRef]
  14. Shinde, V.M.; Madras, G. CO Methanation Toward the Production of Synthetic Natural Gas over Highly Active Ni/TiO2 Catalyst. AIChE J. 2014, 60, 1027–1035. [Google Scholar] [CrossRef]
  15. Meng, F.; Li, Z.; Liu, J.; Cui, X.; Zheng, H. Effect of promoter Ce on the structure and catalytic performance of Ni/Al2O3 catalyst for CO methanation in slurry-bed reactor. J. Nat. Gas Sci. Eng. 2015, 23, 250–258. [Google Scholar] [CrossRef]
  16. Zheng, Y.; Ma, H.; Zhang, H.; Ying, W.; Fang, D. Ni-Ce-Al composite oxide catalysts synthesized by solution combustion method: Enhanced catalytic activity for CO methanation. Fuel 2015, 162, 16–22. [Google Scholar] [CrossRef]
  17. Konishcheva, M.V.; Potemkin, D.I.; Snytnikov, P.V.; Zyryanova, M.M.; Pakharukova, V.P.; Simonov, P.A.; Sobyanin, V.A. Selective CO methanation in H2-rich stream over Ni-, Co- and Fe/CeO2: Effect of metal and precursor nature. Int. J. Hydrogen Energy 2015, 40, 14058–14063. [Google Scholar] [CrossRef]
  18. Nematollahi, B.; Rezaei, M.; Nemati Lay, E. Preparation of highly active and stable NiO-CeO2 nanocatalysts for CO selective methanation. Int. J. Hydrogen Energy 2015, 40, 8539–8547. [Google Scholar] [CrossRef]
  19. Nematollahi, B.; Rezaei, M.; Nemati Lay, E. Selective methanation of carbon monoxide in hydrogen rich stream over Ni/CeO2 nanocatalysts. J. Rare Earth 2015, 33, 619–628. [Google Scholar] [CrossRef]
  20. Rombi, E.; Cutrufello, M.G.; Atzori, L.; Monaci, R.; Ardu, A.; Gazzoli, D.; Deiana, P.; Ferino, I. CO methanation on Ni-Ce mixed oxides prepared by hard template method. Appl. Catal. A Gen. 2016, 515, 144–153. [Google Scholar] [CrossRef]
  21. Yamasaki, M.; Habazaki, H.; Asami, K.; Izumiya, K.; Hashimoto, K. Effect of tetragonal ZrO2 on the catalytic activity of Ni/ZrO2 catalyst prepared from amorphous Ni-Zr alloys. Catal. Commun. 2006, 7, 24–28. [Google Scholar] [CrossRef]
  22. Ocampo, F.; Louis, B.; Roger, A.-C. Methanation of carbon dioxide over nickel-based Ce0.72Zr0.28O2 mixed oxide catalysts prepared by sol-gel method. Appl. Catal. A Gen. 2009, 369, 90–96. [Google Scholar] [CrossRef]
  23. Ocampo, F.; Louis, B.; Kiwi-Minsker, L.; Roger, A.-C. Effect of Ce/Zr composition and noble metal promotion on nickel based CexZr1−xO2 catalysts for carbon dioxide methanation. Appl. Catal. A Gen. 2011, 392, 36–44. [Google Scholar] [CrossRef]
  24. Takano, H.; Izumiya, K.; Kumagai, N.; Hashimoto, K. The effect of heat treatment on the performance of the Ni/(Zr-Sm oxide) catalysts for carbon dioxide methanation. Appl. Surf. Sci. 2011, 257, 8171–8176. [Google Scholar] [CrossRef]
  25. Tada, S.; Shimizu, T.; Kameyama, H.; Haneda, T.; Kikuchi, R. Ni/CeO2 catalysts with high CO2 methanation activity and high CH4 selectivity at low temperatures. Int. J. Hydrogen Energy 2012, 37, 5527–5531. [Google Scholar] [CrossRef]
  26. Cai, W.; Zhong, Q.; Zhao, Y. Fractional-hydrolysis-driven formation of non-uniform dopant concentration catalyst nanoparticles of Ni/CexZr1−xO2 and its catalysis in methanation of CO2. Catal. Commun. 2013, 39, 30–34. [Google Scholar] [CrossRef]
  27. Ussa Aldana, P.A.; Ocampo, F.; Kobl, K.; Louis, B.; Thibault-Starzyk, F.; Daturi, M.; Bazin, P.; Thomas, S.; Roger, A.-C. Catalytic CO2 valorization into CH4 on Ni-based ceria-zirconia. Reaction mechanism by operando IR spectroscopy. Catal. Today 2013, 215, 201–207. [Google Scholar] [CrossRef]
  28. Liu, J.; Li, C.; Wang, F.; He, S.; Chen, H.; Zhao, Y.; Wei, M.; Evans, D.G.; Duan, X. Enhanced low-temperature activity of CO2 methanation over highly-dispersed Ni/TiO2 catalyst. Catal. Sci. Technol. 2013, 3, 2627–2633. [Google Scholar] [CrossRef]
  29. Pan, Q.; Peng, J.; Sun, T.; Gao, D.; Wang, S.; Wang, S.D. CO2 methanation on Ni/Ce0.5Zr0.5O2 catalysts for the production of synthetic natural gas. Fuel Process. Technol. 2014, 123, 166–171. [Google Scholar] [CrossRef]
  30. Takano, H.; Shinomiya, H.; Izumiya, K.; Kumagai, N.; Habazaki, H.; Hashimoto, K. CO2 methanation of Ni catalysts supported on tetragonal ZrO2 doped with Ca2+ and Ni2+ ions. Int. J. Hydrogen Energy 2015, 40, 8347–8355. [Google Scholar] [CrossRef]
  31. Aziz, M.A.A.; Jalil, A.A.; Triwahyono, S.; Ahmad, A. CO2 methanation over heterogeneous catalysts: Recent progress and future prospects. Green Chem. 2015, 17, 2647–2663. [Google Scholar] [CrossRef]
  32. Zhou, G.; Liu, H.; Cui, K.; Jia, A.; Hu, G.; Jiao, Z.; Liu, Y.; Zhang, X. Role of surface Ni and Ce species of Ni/CeO2 catalyst in CO2 methanation. Appl. Surf. Sci. 2016, 383, 248–252. [Google Scholar] [CrossRef]
  33. Danaci, S.; Protasova, L.; Lefevere, J.; Bedel, L.; Guilet, R.; Marty, P. Efficient CO2 methanation over Ni/Al2O3 coated structured catalysts. Catal. Today 2016, 273, 234–243. [Google Scholar] [CrossRef]
  34. Frontera, P.; Macario, A.; Ferraro, M.; Antonucci, P.L. Supported Catalysts for CO2 Methanation: A Review. Catalysts 2017, 7, 59. [Google Scholar] [CrossRef]
  35. Nizio, M.; Albarazi, A.; Cavadias, S.; Amouroux, J.; Galvez, M.E.; Da Costa, P. Hybrid plasma-catalytic methanation of CO2 at low temperature over ceria zirconia supported Ni catalysts. Int. J. Hydrogen Energy 2016, 41, 11584–11592. [Google Scholar] [CrossRef]
  36. Atzori, L.; Cutrufello, M.G.; Meloni, D.; Monaci, R.; Cannas, C.; Gazzoli, D.; Sini, M.F.; Deiana, P.; Rombi, E. CO2 methanation on hard-templated NiO-CeO2 mixed oxides. Int. J. Hydrogen Energy 2017, 42, 20689–20702. [Google Scholar] [CrossRef]
  37. 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]
  38. Atzori, L.; Cutrufello, M.G.; Meloni, D.; Cannas, C.; Gazzoli, D.; Monaci, R.; Sini, M.F.; Rombi, E. Highly active NiO-CeO2 catalysts for synthetic natural gas production by CO2 methanation. Catal. Today 2018, 299, 183–192. [Google Scholar] [CrossRef]
  39. Ratchahat, S.; Sudoh, M.; Suzuki, Y.; Kawasaki, W.; Watanabe, R.; Fukuhara, C. Development of a powerful CO2 methanation process using a structured Ni/CeO2 catalyst. J. CO2 Util. 2018, 24, 210–219. [Google Scholar] [CrossRef]
  40. Yu, Y.; Bian, Z.; Song, F.; Wang, J.; Zhong, Q.; Kawi, S. Influence of Calcination Temperature on Activity and Selectivity of Ni-CeO2 and Ni-Ce0.8Zr0.2O2 Catalysts for CO2 Methanation. Top. Catal. 2018, 61, 1514–1527. [Google Scholar] [CrossRef]
  41. Shang, X.; Deng, D.; Wang, X.; Xuan, W.; Zou, X.; Ding, W.; Lu, X. Enhanced low-temperature activity for CO2 methanation over Ru doped the Ni/CexZr(1−x)O2 catalysts prepared by one-pot hydrolysis method. Int. J. Hydrogen Energy 2018, 43, 7179–7189. [Google Scholar] [CrossRef]
  42. Atzori, L.; Rombi, E.; Meloni, D.; Monaci, R.; Sini, M.F.; Cutrufello, M.G. Nanostructured Ni/CeO2-ZrO2 catalysts for CO2 conversion into SNG. J. Nanosci. Nanotechnol. 2019, 19, 3269–3276. [Google Scholar] [CrossRef]
  43. Habazaki, H.; Yamasaki, M.; Zhang, B.P.; Kawashima, A.; Kohno, S.; Takai, T.; Hashimoto, K. Co-methanation of carbon monoxide and carbon dioxide on supported nickel and cobalt catalysts prepared from amorphous alloys. Appl. Catal. A Gen. 1998, 172, 131–140. [Google Scholar] [CrossRef]
  44. Gogate, M.R.; Davis, R.J. Comparative study of CO and CO2 hydrogenation over supported Rh-Fe catalysts. Catal. Commun. 2010, 11, 901–906. [Google Scholar] [CrossRef]
  45. Kang, S.-H.; Ryu, J.-H.; Kim, J.-H.; Seo, S.-J.; Yoo, J.-D.; Sai Prasad, P.S.; Lim, H.-J.; Byun, C.-D. Co-methanation of CO and CO2 on the NiX-Fe1−X/Al2O3 catalysts; effect of Fe contents. Korean J. Chem. Eng. 2011, 28, 2282–2286. [Google Scholar] [CrossRef]
  46. Huang, Y.-H.; Wang, J.-J.; Liu, Z.-M.; Lin, G.-D.; Zhang, H.-B. Highly efficient Ni-ZrO2 catalyst doped with Yb2O3 for co-methanation of CO and CO2. Appl. Catal. A Gen. 2013, 466, 300–306. [Google Scholar] [CrossRef]
  47. Razzaq, R.; Zhu, H.; Jiang, L.; Muhammad, U.; Li, C.; Zhang, S. Catalytic Methanation of CO and CO2 in Coke Oven Gas over Ni-Co/ZrO2-CeO2. Ind. Eng. Chem. Res. 2013, 52, 2247–2256. [Google Scholar] [CrossRef]
  48. Razzaq, R.; Li, C.; Amin, N.; Zhang, S.; Suzuki, K. Co-methanation of Carbon Oxides over Nickel-Based CexZr1−xO2 Catalysts. Energy Fuels 2013, 27, 6955–6961. [Google Scholar] [CrossRef]
  49. Razzaq, R.; Li, C.; Usman, M.; Suzuki, K.; Zhang, S. A highly active and stable Co4N/γ-Al2O3 catalyst for CO and CO2 methanation to produce synthetic natural gas (SNG). Chem. Eng. J. 2015, 262, 1090–1098. [Google Scholar] [CrossRef]
  50. Li, Y.; Zhang, Q.; Chai, R.; Zhao, G.; Liu, Y.; Lu, Y. Structured Ni-CeO2-Al2O3/Ni-Foam Catalyst with Enhanced Heat Transfer for Substitute Natural Gas Production by Syngas Methanation. ChemCatChem 2015, 7, 1427–1431. [Google Scholar] [CrossRef]
  51. Zhao, K.; Li, Z.; Bian, L. CO2 methanation and co-methanation of CO and CO2 over Mn-promoted Ni/Al2O3 catalysts. Front. Chem. Sci. Eng. 2016, 10, 273–280. [Google Scholar] [CrossRef]
  52. Frontera, P.; Macario, A.; Malara, A.; Modafferi, V.; Mascolo, M.C.; Candamano, S.; Crea, F.; Antonucci, P. CO2 and CO hydrogenation over Ni-supported materials. Funct. Mater. Lett. 2018, 11, 1850061. [Google Scholar] [CrossRef]
  53. Xie, Y.; Kocaefe, D.; Chen, C.; Kocaefe, Y. Review of Research on Template Methods in Preparation of Nanomaterials. J. Nanomater. 2016, 2016, 2302595. [Google Scholar] [CrossRef]
  54. Rouquerol, F.; Rouquerol, J.; Sing, K.S.W.; Llewellyn, P.; Maurin, G. Adsorption by Powders and Porous Solids, Principles, Methodology and Applications, 2nd ed.; Elsevier: Amsterdam, The Netherlands, 2014. [Google Scholar]
  55. Groen, J.C.; Peffer, L.A.A. Perez-Ramirez, Pore size determination in modified micro- and mesoporous materials. Pitfalls and limitations in gas adsorption data analysis. J. Micropor. Mesopor. Mater. 2003, 60, 1–17. [Google Scholar] [CrossRef]
  56. Klug, H.P.; Alexander, L.E. X-Ray Diffraction Procedures; Wiley: New York, NY, USA, 1962. [Google Scholar]
  57. Jalowiecki-Duhamel, L.; Ponchel, A.; Lamonier, C.; D’Huysser, A.; Barbaux, Y. Relationship between Structure of CeNiXOY Mixed Oxides and Catalytic Properties in Oxidative Dehydrogenation of Propane. Langmuir 2001, 17, 1511–1517. [Google Scholar] [CrossRef]
  58. Wei, Y.; Wang, H.; Li, K.; Zhu, X.; Du, Y. Preparation and characterization of Ce1−xNixO2 as oxygen carrier for selective oxidation methane to syngas in absence of gaseous oxygen. J. Rare Earths 2010, 28, 357–361. [Google Scholar] [CrossRef]
  59. Sagar, T.V.; Sreelatha, N.; Hanmant, G.; Surendar, M.; Lingaiah, N.; Rama Rao, K.S.; Satyanarayana, C.V.V.; Reddy, I.A.K.; Sai Prasad, P.S. Influence of method of preparation on the activity of La-Ni-Ce mixed oxide catalysts for dry reforming of methane. RSC Adv. 2014, 4, 50226–50232. [Google Scholar] [CrossRef]
  60. Wang, Y.; Ma, J.; Luo, M.; Fang, P.; He, M. Preparation of High-Surface Area Nano-CeO2 by Template-Assisted Precipitation Method. J. Rare Earths 2007, 25, 58–62. [Google Scholar]
Figure 1. N2 adsorption (full symbols) and desorption (open symbols) isotherms and pore size distribution curves (insets) of the NiO/CexZr1−xO2 samples.
Figure 1. N2 adsorption (full symbols) and desorption (open symbols) isotherms and pore size distribution curves (insets) of the NiO/CexZr1−xO2 samples.
Catalysts 09 00415 g001
Figure 2. XRD patterns of the fresh NiO/CexZr1−xO2 samples.
Figure 2. XRD patterns of the fresh NiO/CexZr1−xO2 samples.
Catalysts 09 00415 g002
Figure 3. TPR curves of the NiO/CexZr1−xO2 samples.
Figure 3. TPR curves of the NiO/CexZr1−xO2 samples.
Catalysts 09 00415 g003
Figure 4. XRD patterns of the Ni/CexZr1−xO2 samples.
Figure 4. XRD patterns of the Ni/CexZr1−xO2 samples.
Catalysts 09 00415 g004
Figure 5. Differential heat of adsorption, Qdiff, as a function of CO2 uptake on the Ni/CexZr1−xO2 samples.
Figure 5. Differential heat of adsorption, Qdiff, as a function of CO2 uptake on the Ni/CexZr1−xO2 samples.
Catalysts 09 00415 g005
Figure 6. Thermal stability test for Ni/Ce0.50Zr0.50O2 in the COx co-methanation reaction: (a) Reactor temperature profile; (b) CO conversion (XCO) and CO2 conversion (XCO2) vs. time-on-stream (t.o.s.). CO/CO2/H2 = 1/1/5; SV = 150,000 cm3 h−1 gcat−1.
Figure 6. Thermal stability test for Ni/Ce0.50Zr0.50O2 in the COx co-methanation reaction: (a) Reactor temperature profile; (b) CO conversion (XCO) and CO2 conversion (XCO2) vs. time-on-stream (t.o.s.). CO/CO2/H2 = 1/1/5; SV = 150,000 cm3 h−1 gcat−1.
Catalysts 09 00415 g006
Figure 7. 6-h average CO conversion (XCO) and CO2 conversion (XCO2) vs. space velocity (SV) for Ni/Ce0.50Zr0.50O2 in the COx co-methanation reaction. T = 300 °C; red: CO/CO2/H2 = 1/1/5; blue: CO/CO2/H2 = 1/1/7.
Figure 7. 6-h average CO conversion (XCO) and CO2 conversion (XCO2) vs. space velocity (SV) for Ni/Ce0.50Zr0.50O2 in the COx co-methanation reaction. T = 300 °C; red: CO/CO2/H2 = 1/1/5; blue: CO/CO2/H2 = 1/1/7.
Catalysts 09 00415 g007
Scheme 1. Stoichiometric equations for: (1) CO methanation; (2) CO2 methanation; (3) reverse methane dry reforming; (4) water-gas shift [1].
Scheme 1. Stoichiometric equations for: (1) CO methanation; (2) CO2 methanation; (3) reverse methane dry reforming; (4) water-gas shift [1].
Catalysts 09 00415 sch001
Table 1. Chemical composition and textural features of the NiO/CexZr1−xO2 samples.
Table 1. Chemical composition and textural features of the NiO/CexZr1−xO2 samples.
SampleNiO Content
(wt%) a
CeO2 Molar Fraction in (CeO2)x-(ZrO2)1−x aSBET
(m2 g−1) b
Vp
(cm3 g−1) b
NiO/ZrO230.2-2820.27
NiO/Ce0.25Zr0.75O229.50.2632450.23
NiO/Ce0.50Zr0.50O229.00.5072450.31
NiO/Ce0.75Zr0.25O229.10.7371980.28
NiO/CeO229.0-1970.31
a ICP-AES results. b From N2 adsorption data.
Table 2. 6-h average CO conversion (XCO), CO2 conversion (XCO2), and CH4 selectivity (SCH4) for the Ni/CexZr1−xO2 samples in the COx co-methanation reaction. T = 300 °C; CO/CO2/H2 = 1/1/5; SV = 150,000 cm3 h−1 gcat−1.
Table 2. 6-h average CO conversion (XCO), CO2 conversion (XCO2), and CH4 selectivity (SCH4) for the Ni/CexZr1−xO2 samples in the COx co-methanation reaction. T = 300 °C; CO/CO2/H2 = 1/1/5; SV = 150,000 cm3 h−1 gcat−1.
SampleXCO (mol%)XCO2 (mol%)SCH4 (mol%)
NiO/ZrO29514>99
NiO/Ce0.25Zr0.75O29516>99
NiO/Ce0.50Zr0.50O29821>99
NiO/Ce0.75Zr0.25O29821>99
NiO/CeO29821>99

Share and Cite

MDPI and ACS Style

Atzori, L.; Rombi, E.; Meloni, D.; Sini, M.F.; Monaci, R.; Cutrufello, M.G. CO and CO2 Co-Methanation on Ni/CeO2-ZrO2 Soft-Templated Catalysts. Catalysts 2019, 9, 415. https://doi.org/10.3390/catal9050415

AMA Style

Atzori L, Rombi E, Meloni D, Sini MF, Monaci R, Cutrufello MG. CO and CO2 Co-Methanation on Ni/CeO2-ZrO2 Soft-Templated Catalysts. Catalysts. 2019; 9(5):415. https://doi.org/10.3390/catal9050415

Chicago/Turabian Style

Atzori, Luciano, Elisabetta Rombi, Daniela Meloni, Maria Franca Sini, Roberto Monaci, and Maria Giorgia Cutrufello. 2019. "CO and CO2 Co-Methanation on Ni/CeO2-ZrO2 Soft-Templated Catalysts" Catalysts 9, no. 5: 415. https://doi.org/10.3390/catal9050415

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