MATERIALS TRANSACTIONS
Online ISSN : 1347-5320
Print ISSN : 1345-9678
ISSN-L : 1345-9678
Mechanics of Materials
Positive Effects of CeO2 Promoter and Co-Reactant/CO on Methanation of CO2-Rich Gas over Ni/SBA-15 Catalyst
Nguyen T. T. VanLuu C. LocNguyen P. AnhHoang T. CuongNguyen Tri
Author information
JOURNAL FREE ACCESS FULL-TEXT HTML

2020 Volume 61 Issue 7 Pages 1332-1338

Details
Abstract

NiO/SBA-15 catalysts modified by CeO2 were synthesized and studied for the hydrogenation of CO2-rich gas. Effect of CeO2 promoter on the physicochemical properties of the catalysts was investigated by BET, XRD, Raman spectroscopy, SEM, TEM, H2-TPR and CO2-TPD methods. The 4 mass% CeO2 addition improved reduction and dispersion of Ni, resulting in its higher activity in methanation, reaching CO2 conversion of 90% at 350°C. Furthermore, the performance of the prepared catalysts in methanation reaction was enhanced by adding small amount of CO (1 mol%) in the feedstock.

Fig. 11 CO2 conversion in methanation of mixture of 19 mol% CO2 with 1 mol% CO (case I - dashed lines) and with 1 mol% N2 (case II - solid lines) on Ni4Ce/SBA catalyst.

1. Introduction

It was observed that global temperature has increased in parallel with the atmospheric concentration of carbon dioxide that has been constantly rising, from 375 ppm in 2001 to 390.5 ppm in 2011, 403.53 ppm in 2016, and 500 ppm within the next 50 years.1,2) As a result, the highest temperature which was about 1.1°C higher than the pre-industrial level was recorded in 2016 since 1880. Therefore, if greenhouse gas emissions cannot be properly reduced, the global temperature increase will soon reach the red line at which sea level rises more rapidly and extreme weather will happen more frequently.2) The elimination of CO2 has been attracting interest from the environmental perspective with various methods.3)

Methanation of CO2, which is abundant in the exhaust gas and a “zero-calorific value” greenhouse gas,4) for on-site production of CH4 fuel could be a promising and effective strategy for CO2 emissions mitigation.5,6) Currently, nickel-based catalysts are good candidates because of its lower cost and comparable catalytic activity for methanation6) despite exhibiting lower stability and resistance against coke deposition compared to the noble metal catalysts.7,8) Moreover, using various traditional supports, such as γ-Al2O3, SiO2, TiO2 has been shown to be an effective way to improve the activity of Ni-based catalysts.9) Basically, high nickel loading is always desirable, however, the optimal value has been found to be strongly dependent on the nature of supports since excessive loading can cause blockage of their channels.10) It has been reported that NiO can be loaded up to 10 mass% on ZnxMg1−xAl2O4,6) 20 mass% on Al2O3,11) 20.7 mass% on ZrO212) and especially up to 50 mass% on ordered mesoporous SBA-1513) thanks to its excellent features such as high surface area, large pore volume, thermal stability and uniform pore size distribution. Nevertheless, it has been reported that Ni particles could not stay stable inside the mesoporous channels of the silica support at elevated temperatures due to their naturally weak interaction.14,15) Fortunately, this interaction could be strengthened by using another oxide such as TiO2, ZrO2, or CeO2.16,17) Among them, CeO2 is used as a catalyst support in CO oxidation, NOx purification reaction, and water gas shift reaction.18) Besides that, CeO2 has been widely studied as a promising promoter due to its excellent capability of: (i) stabilizing metal particles against thermal sintering,19) (ii) modifying the structural and electronic properties of Ni,20) (iii) enhancing the interaction between Ni particles and support,21) thus, (iv) controlling the growth of the active metal.22) In addition, it was reported that CeO2 could promote effectively the adsorption and activation of CO2 molecules on the MgAl2O4 supported NiCeO2ZrO2 catalyst due to its oxygen exchange capacity.23) Besides, CeO2-modified NiO catalysts are generally less prone to coke deposition.24) The interaction of Ni2+ with CeO2 leads to the formation of Ce2O3 and oxygen vacancies, enhancing dissociative adsorption of CO2. This, in turn, improves the coke resistance of CeO2-doped catalysts, as confirmed by Li et al.25)

In this study, the effect of CeO2 promoter on the physico-chemical properties and catalytic performance of the Ni-based catalysts supported on SBA-15 was particularly investigated. Besides, adding a small amount of co-reactant/CO into the reaction feedstock to improve the process efficiency of CO2 methanation in terms of catalytic performance was also studied.

2. Experimental

SBA-15 was synthesized by the combination of sol–gel and hydrothermal methods according to the report by authors:26) 4 gram of Poly(ethylene glycol)-block-poly(propylene glycol)-block-poly(ethylene glycol) (P123) (Sigma-Aldrich) was diluted by deionized water to form 105 mL solution; afterthat, 8.8 mL of Tetraethyl orthosilicate (TEOS) (≥99.9%, Merck) was added drop by drop to the above solution with vigorous stirring, followed by adding 24.2 mL of HCl (≥99.9%, Merck); after further stirring for 30 minutes, the mixture was transferred to a Teflon autoclave for hydrothermal treatment at 60°C for 24 hours. After that, the solid was filtered off, washed many times with deionized water and ethanol, and calcined at 550°C for 10 hours to obtain SBA-15.

The CeO2-modified 50 mass% NiO/SBA-15 catalyst was prepared by the co-impregnation method from the precursor solutions using Ni(NO3)2·6H2O (99.8 mass%, Prolabo) and Ce(NO3)3·6H2O (99.8 mass%, Merck) at room temperature for 5 hours with ultrasound assistance. After stirring at 80°C and 100°C for 2 hours at each temperature, the sample was calcined in air at 600°C for 4 hours. The obtained catalysts were denoted as NixCe/SBA, where x = 4–5 mass%. The Ce/Si atomic ratio was in the range of 0.028–0.036, being below the limited value of 0.04.27)

The characteristics of obtained catalysts were determined by different methods. X-ray powder diffraction (XRD) using Bruker D2 Phaser powder diffractometer with CuKα radiation (λ = 0.15406 nm) was used to investigate the crystalline structure of the catalysts. Raman spectroscopy measurements were carried out at room temperature with a laser Raman spectrometer (Invia, Renishaw). Morphology of catalysts and particle size of the active phase was characterized by scanning electron microscopy (SEM) on the FE-SEM JEOL 7401 instrument and transmission electron microscopy (TEM) on the TEM JEM 1400 of Jeol USA instrument. The hydrogen temperature-programmed reduction (H2-TPR) measurement was performed in a fixed-bed reactor at atmospheric pressure by using a Gas Chromatograph GOW-MAC 69-350 with a TCD detector and the reduction gas of 10 mol% H2/N2 was introduced. The temperature of the reactor was raised linearly from 50 to 900°C at a heating rate of 10°C·min−1. The carbon dioxide temperature-programmed desorption (CO2-TPD) experiment was carried out to determine the basicity of the catalyst. This technique was carried out in a series of steps. Firstly, catalyst was reduced in H2 at 450°C for 1 hour and cooled down to room temperature, followed by CO2 adsorption for 1 hour. After purging with N2 for 1 hour at 50°C to remove gas-phase and physically adsorbed CO2, the sample was heated from 50 to 700°C at a rate of 10°C·min−1. The desorbed CO2 was then monitored online by using a Gas Chromatograph GOWMAC 69-350 with a thermal conductivity detector (TCD).

Methanation of CO2-rich gas was tested in a standard micro-flow reactor with and internal diameter of 10 mm under atmospheric pressure and a temperature range of 250–350°C. The loaded catalyst (0.2 grams) was mixed with quartz (0.8 grams, 99.95 mass%, Sigma-Aldrich) with a size range of 0.25–0.5 mm. The CO2/H2 molar ratio was 1/4 and gas hourly space velocity (GHSV) was 15,000 h−1. For common experiments, the CO2 concentration in feed was 20 mol%; for studying the effect of CO, the mixture of 19 mol% CO2 + 1 mol% CO (case I) and 19 mol% CO2 + 1 mol% N2 (case II) were used. Before carrying out the reaction, the catalyst was reduced in-situ in H2 flow at 450°C for 4 hours. The reaction mixture was analyzed using the Agilent 6890 Plus Gas Chromatograph equipped with a TCD detector (capillary column HP-PLOT MoleSieve 5A) and an FID detector (capillary column DB624).

The CO2 adsorption capacity at the typical reaction temperature of the samples reduced was also assessed. Firstly, 0.2 grams of the catalyst was reduced in H2 at 450°C for 4 hours and cooled down to a typical reaction temperature of 350°C. Then, the adsorption gas of 60,000 ppm CO2 in N2 was introduced. CO2 concentration with time is continuously recorded online by CO2 sensor (GMP251, Vaisala) connected with the computer using Senko Sensor Monitor. The adsorption process terminated when CO2 concentration in the output reached 60,000 ppm. The CO2 adsorption capacity ($\textit{mg}_{\textit{CO}_{2}}\cdot g_{\textit{cat}}^{ - 1}$) was calculated by eq. (1).   

\begin{equation} A_{\textit{CO${_{2}}$},\textit{ads}} = \frac{\biggl(C_{o}\cdot t - \displaystyle\int\limits_{0}^{t}C (t)\biggr)\cdot 10^{-3}\cdot \frac{V}{3600}\cdot M_{\textit{CO${_{2}}$}}}{22.4m_{\textit{cat}}} \end{equation} (1)
where: Co is the CO2 concentration in the input, 60,000 ppm; t is the saturated adsorption time, s; C(t) is the function CO2 concentration by time in the output; V is the flow of the adsorption gas, l·h−1; $M_{CO_{2}}$ is the CO2 molar mass of 44.01 g·mol−1; mcat is the loaded catalyst mass.

3. Results and Discussions

3.1 Physico-chemical characteristics

It could be observed from the low-angle XRD pattern for bare SBA-15 support (Fig. 1) that there were two-dimensional SBA-15 an intensive main diffraction peak 2θ = 0.9° and two small peaks at 2θ of 1.61° and 1.84°, indexed as the corresponding (100), (110) and (200) reflections of two dimensional, ordered hexagonal meso-structure.28) These demonstrated that a two-dimensional SBA-15 was successfully synthesized. The low-angle diffraction pattern of the as-synthesized Ni/SBA-15 catalyst exhibited the main peak corresponding to (100) reflection with low intensity and shifts toward a larger angle. The remaining two reflections almost disappeared. The same phenomenon was observed on Fe/SBA-15, TiO2/SBA-15, NiO/SBA-15 catalysts2931) and boron-doped Ni/SBA-15 catalysts.32) This was explained by the presence of the nitrate salt (Ni(NO3)2) in an amorphous state leading to good dispersion of metal on the surface of supports.30)

Fig. 1

The low angles XRD patterns of SBA-15 and Ni/SBA-15.

The characteristic peaks of NiO crystals appear sharply at 2θ = 37.3, 43.3 and 62.8° (JCPDS card No. 47-1049) while diffraction peaks of CeO2 were not observed on XRD patterns of the CeO2-doped catalysts (Fig. 2). This fact indicated the high crystallinity of NiO and amorphous or well-dispersed state of CeO2 on the surface. Aghamiri et al.33) also believed that Ce affected the structure of the catalyst leading to an increase in the Ni dispersion on SiO2 surface. XRD patterns of the catalysts reduced at 450°C for 4 hours in the H2 flow of 2 L·h−1 showed a new phase assigned to metallic Ni crystallites detected at 2θ = 44.9, 51.8, and 76.4° (JCPDS card No. 87-0712). Furthermore, it was clearly seen that the diffraction peaks of the NiO phase almost were not observed, indicating that Ni2+ ions in the NiO phase were fully reduced to metallic Ni at 450°C for 4 hours in hydrogen. The size of the NiO and Ni crystallites was calculated at the strongest intensity peak corresponding to 2θ = 43.3° and 44.9° respectively by Scherrer formula34) and was presented in Table 1. It was shown that the adding 4–5 mass% CeO2 to Ni/SBA catalyst led to a decrease in crystallite sizes of NiO as well as Ni supported on SBA-15. The results of N2 sorption isotherms (seen in Fig. 3) showed that all materials exhibited Type IV adsorption isotherms with Type-H1 hysteresis loops.35) The loop size decreased with adding nickel and/or ceria macropore character, indicating that nickel and ceria have been attached to the channels of SBA-15.

Fig. 2

XRD patterns of as-synthesized catalysts ((a) Ni/SBA and (c) Ni4Ce/SBA) and catalysts reduced in H2 at 450°C for 4 hours ((b) Ni/SBA and (d) Ni4Ce/SBA).

Table 1 Textural properties of the prepared materials.
Fig. 3

Nitrogen adsorption-desorption isotherms of the samples ((a) SBA-15; (b) Ni/SBA; and (c) Ni4Ce/SBA).

SEM images (Fig. 4) showed that catalyst morphology consisted of cocoon-shaped particles of SBA-15 support with 200–500 nm in diameter and 600–800 nm in length. Their surface was highly covered by spherical NiO particles of dozen nanometres in size. There was no noticeable change in surface morphology and macroscopic structure of the Ni/SBA-15 catalyst by adding CeO2. The SEM micrograph of the CeO2-doped catalyst (Fig. 4(c)) showed that the cocoon-shaped particles were stacked together. As a result, the surface area of the CeO2-doped sample was lower than that of undoped one (176.1 vs. 214.7 m2·g−1).

Fig. 4

SEM images of SBA-15 support (a) and as-synthesized catalysts Ni/SBA (b) and Ni4Ce/SBA (c).

Additionally, from the TEM images of the prepared catalysts (Fig. 5), well-ordered hexagonal channels of a few nm in diameter could be observed. NiO particles of several and dozen nanometres were found to be located both inside and outside of the nanochannels of the support. Similar results were observed in the previous studies.36)

Fig. 5

TEM images of as-synthesized catalysts Ni/SBA (a), Ni4Ce/SBA (b), and Ni5Ce/SBA (c).

Figure 6 showed the reduction behavior of both CeO2-doped and undoped catalysts and 4 mass% CeO2/SBA-15. The H2-TPR profile of 4 mass% CeO2/SBA sample multiplied by 30 times showed that CeO2/SBA sample was reduced unsignificantly. The ceria subsurface oxygen was reduced at high temperatures around 800°C.37) For Ni-based catalysts, the reduction peaks of free NiO species, appearing at a low temperature of 350–370°C,38) were observed in all catalysts. Comparing to Ni/SBA-15 catalyst, the maximum reduction temperature in H2-TPR patterns shifted to a lower value (340°C versus 366°C) and bigger reduction peaks were observed in Ce-modified catalysts. These facts indicated that NiO species are more easily and better reduced by introducing CeO2. Besides, two reduction peaks in the range of 480–580°C and 600–740°C observed in H2-TPR profiles of Ce-modified catalysts mighty be attributed to the strong interaction of NiO with CeO2 and SBA-15.22,39) These phases of nickel should be crucial for obtaining sintering-resistant nickel particles during the calcination and reduction process. The good dispersion of NiO particles in the CeO2-doped catalyst is explained by the strong interaction of Ni2+ with CeO2 forming Ce3+ ions and oxygen vacancies that dispersed NiO particles better on the catalyst surface.40)

Fig. 6

H2-TPR spectra of catalysts.

The raman spectra of Ni/SBA and Ni4Ce/SBA were also illustrated in Fig. 7. Two bands at around 490 cm−1 (α) and 1070 cm−1 (β), a strong band at around 465 cm−1 and a stretch band at 900–1200 cm−1 were detected on raman spectra of Ni/SBA and NiCe/SBA catalysts, respectively. According to the authors,32) for SBA-15 support and NiO/SBA-15 catalyst, two bands detected in raman spectra at about 487.3 cm−1 and 1068.7 cm−1 were attributed to the cyclic tetrasiloxane rings and Si-OH stretching mode, respectively. So, two bands detected on raman spectrum of Ni/SBA at about 490 cm−1 (α) and 1070 cm−1 (β) were attributed to the corresponding cyclic tetrasiloxane rings and Si-OH stretching mode, typical for SBA-15 support.36,41) However, compared to Thielemann’s report,41) the blue shift of β band (1070 vs. 970 cm−1) in this case indicated the formation of Si–O–Ni bond.38) For the raman spectrum of Ni4Ce/SBA sample, the band at around 465 cm−1 was attributed to the F2g mode of the cerium oxide. This overlapped with the raman band of tri-cyclosiloxane rings of SBA-1538) or pure cubic CeO242) or the symmetric “breathing” vibrations of the oxygen anions around the cerium cation of crystalline CeO2 of fluorite structure. The band stretching between 900–1200 cm−1 in raman spectrum of Ni4Ce/SBA catalyst might include the band at 983 cm−1, attributed to the Si–O–Ce linkages38) and the bands at 1098 cm−1, assigned to the second-order longitudinal optical mode of CeO2.43) The formation of Si–O–Ce bond improved the interaction between active species and support, which facilitated CO methanation.38)

Fig. 7

Raman spectra of catalysts.

CO2-TPD patterns (Fig. 8) showed that all samples exhibited the weak (α), moderate (β), and strong basic sites (γ), corresponding to three CO2 desorption zones in the range of 115–150°C; 200–300°C and 500–600°C, respectively. In particular, the number of weak basic sites was dominant. This changed noticeably while the amount of average and strong basic sites decreased when CeO2 was added. The higher basicity of the NiCe/SBA sample, as seen in Table 1, was explained by the adsorption of CO2 on surface Ce4+ and Ce3+ ions to form carbonate CO32− species.44) The good migration ability of oxygen in the crystal and the electron delocalization formed by the oxygen vacancy could increase the electron density in the nano-CeO2 structure45,46) that increased the strength and the amount of the basic sites.47) Another cause was the high dispersion of CeO2 in catalysts as demonstrated in XRD analysis. The decrease in particle size of CeO2 was proved to be favorable for the increase of basicity of catalysts.48)

Fig. 8

CO2-TPD patterns of catalysts.

3.2 Activity of catalysts

The results in Fig. 9 showed that the CO2 conversions increased when catalyst was modified with 4 mass% CeO2 at all the temperature values in range of 250–350°C. Besides, the reaction temperature was lowered at the same CO2 conversion (e.g. 50%) when Ni/SBA was replaced by Ni4Ce/SBA, from 325 down to 312°C. The CH4 selectivity of CeO2-doped catalyst reached over 99% at all temperatures. It could be explained that the modification of CeO2 led to increasing of CO2 adsorption capacity and catalytic activity as a consequence.49) According to authors,45,46) in the CeO2 crystal, oxygen has a good migration ability and the electron delocalization formed by the oxygen vacancy that can increase the electron density in the nano-CeO2 structure and the strength and the amount of the basic sites. The obtained results (Fig. 10) are well proven that at a typical reaction temperature of 350°C, the CO2 adsorption capacity of Ni4Ce/SBA catalyst was slightly higher than that of Ni/SBA (10.9 compared to 10.4 $\textit{mg}_{\textit{CO}_{2}}.g_{\textit{cat}}^{ - 1}$, respectively).

Fig. 9

CO2 conversion in solo-methanation of 20 mol% CO2 gas on Ni/SBA catalysts modified with varying in CeO2 contents.

Fig. 10

CO2 adsorption by time at 350°C on the reduced catalysts (T = 350°C; mcat = 0.05 g; 6 mol% CO2 + 94 mol% N2; GHSV = 15,000 h−1).

On the other hand, Boreskov et al.50) showed that smaller crystal size of Ni led to a higher of CO2 conversion, and the optimal values was in the range of 10–20 nm, corresponding to the highest catalytic activity. Meanwhile, the results in Table 1 showed that the increase in CeO2 content from 3 to 4 mass% led to a reduction in the crystallite size of reduced Nio (from 23.3 to 20.5 nm) and an increase of CO2 conversion consequently. However, if CeO2 content was continually increased to 5 mass%, the catalyst activity was dropped. This could be explained that when the CeO2 content was high, the Ni sites were occupied by the larger amount of produced O species, then the reaction conversion decreased.51)

It could be observed from Fig. 11 that adding of 1 mol% of CO into the CO2 feedstock stream led to an enhancement of CO2 conversion as well as a drop in reaction temperature of methanation. Moreover, the reaction temperature of CO2 hydrogenation without CO (case II) was 313°C while it was only 290°C in case I (with 1% CO in the feedstock stream) at the conversion of 50% on Ni4Ce/SBA catalyst. The CH4 selectivity still reached 98–99%. This result was also confirmed by Tada et al.52) It could be explained by thermal dynamic data of following reactions:53)   

\begin{align} &\text{CO$_{2}$(g)} + \text{H$_{2}$(g)} \rightleftharpoons \text{CO(g)} + \text{H$_{2}$O (g)}, \\ &\Delta \text{H} = +41\,\text{kJ${\cdot}$mol$^{-1}$} \end{align} (2)
  
\begin{align} &\text{CO(g)} + \text{3H$_{2}$(g)} \rightleftharpoons \text{CH$_{4}$(g)} + \text{H$_{2}$O(g)},\\ &\Delta \text{H} = -206\,\text{kJ${\cdot}$mol$^{-1}$} \end{align} (3)
  
\begin{align} &\text{CO$_{2}$(g)} + \text{4H$_{2}$(g)} \rightleftharpoons \text{CH$_{4}$(g)} + \text{2H$_{2}$O(g)},\\ &\Delta \text{H} = -165\,\text{kJ${\cdot}$mol$^{-1}$} \end{align} (4)

Fig. 11

CO2 conversion in methanation of mixture of 19 mol% CO2 with 1 mol% CO (case I - dashed lines) and with 1 mol% N2 (case II - solid lines) on Ni4Ce/SBA catalyst.

The reaction (2) is endothermic so that it takes place at high temperature. The reaction (3), which is exothermic, easily occur by presence of a small amount of CO in feedstock stream. Furthermore, heat released from this reaction causes the temperature in the reaction (2) decreased and the reaction (2) shifts forward. As a result, both CO2 conversion and CH4 selectivity increase. It is also observed that the conversion of CO in the mixture reaches 100% at 250°C on Ni4Ce/SBA catalyst.

4. Conclusion

The catalytic activity of NiO/SBA-15 in methanation of carbon dioxide was enhanced when modified by CeO2. Adding CeO2 led to an increase in CO2 conversion and decrease in reaction temperature due to the enhancement of the dispersion of metallic Nio, improvement of the NiO reducibility and CO2 adsorption. However, the conversion decreased when CeO2 content was higher than 4 mass%. In addition, adding a small amount of CO lowered the temperature of CO2 methanation, showing a promotional effect of co-reaction/CO in the CO2 hydrogenation.

Acknowledgments

This research was supported by Vietnam Academy of Science and Technology under the grant No. ĐLTE00.10/18-19.

REFERENCES
 
© 2020 The Japan Institute of Metals and Materials
feedback
Top