DOI:
10.1039/C5RA15990C
(Paper)
RSC Adv., 2015,
5, 96504-96517
Ni/SBA-15 catalysts for CO methanation: effects of V, Ce, and Zr promoters
Received
9th August 2015
, Accepted 6th November 2015
First published on 6th November 2015
Abstract
Ni/SBA-15 catalysts with various promoters (V, Ce, and Zr) were prepared by an ultrasonic coimpregnation method and used in CO methanation. The addition of promoters played a significant role in improving the catalytic activity of the Ni/SBA-15 catalyst. This improvement could be explained by changes in the valences of the V and Ce promoter species through an oxidation–reduction shift cycle process (Mx+ ↔ My+, M = V, Ce), which could trigger electron transfer. This transfer enhanced the electron density of active Ni species and promoted CO dissociation. The Zr promoter could produce oxygen vacancies during calcination and reduction, thereby increasing the ability of CO to adsorb and dissociate. In addition, the formation of a Si–O–M bond (M = Zr, Ce, V) increased the interaction between the active species and support, which facilitated CO methanation. Under 1.0 MPa and a WHSV of 15
000 mL g−1 h−1, 10Ni–5V/SBA-15 exhibited the best catalytic performance (99.9% CO conversion; 95.5% CH4 selectivity).
1. Introduction
Since Sabatier and Senderens first discovered CO methanation in 1902, the reaction has been widely studied by researchers.1–4 It is now used in eliminating trace amounts of CO from feed gas for ammonia synthesis. It is also used in forming methane-rich fuel, which has a high heating value and low CO content, as well as in processes related to Fischer–Tropsch synthesis.1 In the 21st century, the shortage of energy resources and environmental pollution caused by rapid consumption of fossil fuels has prompted many investigators to make great efforts to address these problems.5 Selectively catalytic methanation of CO is used to purify hydrogen for proton-exchange membrane fuel cells, which prevents their anodes from severe poisoning by CO in hydrogen-rich gas. The CO concentration in the gas is <50 ppm.6 Another effective method is production of synthetic or substitute natural gas (SNG) via syngas (CO and H2) from coal and biomass.7 SNG is natural gas that has good combustibility and heating value. Therefore, development of viable and efficient catalysts for methanation is necessary.
Among the methanation catalysts, those based one noble metals such as Ru,8 Rh,9 Pd10 show better performance. But the scarcity and expense of noble metals restrict their large-scale applications. Because of their good activity and cost effectiveness, Ni-based catalysts have been extensively investigated and widely applied. However, Ni-based catalysts are vulnerable to sintering and coking, which may lead to their deactivation. Hence, many efforts have been made to enhance catalytic activity and stability, including selection of appropriate supports, addition of catalytic promoters, and search for advanced preparation methods. There have been extensive studies on catalyst supports such as Al2O3,11 SiO2,12 TiO2,13 ZrO2 (ref. 14) and CeO2.15 Takenaka et al.16 demonstrated that CO methanation performance of Ni-based catalysts on various supports are closely related to the type of catalytic supports. CO conversion at 250 °C increases in the order Ni/MgO < Ni/Al2O3 < Ni/SiO2 < Ni/TiO2 < Ni/ZrO2. Addition of promoters can effectively enhance Ni dispersion, as well as increase the activity and stability of Ni-based catalysts. Wang et al.17 reported that use of ZrO2 as promoter for Ni/SiO2 catalyst significantly improves the catalytic activity, enabling complete conversion of CO into CH4 at 240 °C in hydrogen-rich gas stream through methanation. Hayek et al.18 found that V and Ce are good electronic promoters for Rh- and Pd-based catalysts in CO hydrogenation for hydrocarbon production. Another study revealed that addition of V promoter could markedly enhance the performance of Ni/Al2O3 catalyst in CO methanation.19 Ren et al.20 prepared Ni–Ce/Al2O3 catalysts by microwave heating method for the catalytic methanation of coke oven gas, the temperature gradient between the NiO and the alumina support caused by the difference in microwave energy absorption promoted the formation of a large amount of highly dispersed amorphous NiO, and then enhanced the catalytic activity and stability of this reaction. Mesoporous silica materials have been widely utilized as novel catalytic supports because of their large specific surface area, regular pore structure, and good hydrothermal stability. Zhang et al.21 reported that Ni-based MCM-41 catalysts, which they used in CO methanation, exhibited good activity and CH4 selectivity. However, the Ni species was easily sintered by calcination and high-temperature reduction. The mesoporous silica material SBA-15 has thicker wall of hole, better hydrothermal stability, and stronger mechanical strength compared with MCM-41. Lu et al.22 found that NiO/SBA-15 catalysts prepared by heat treatment showed good activity and thermal stability in CO/CO2 methanation; however, the CH4 selectivity of the catalyst in methanation as compared with other existing catalysts needs further improvement.
In the present work, highly efficient Ni/SBA-15 catalysts were prepared by ultrasonic coimpregnation method, and the effect of Ni loading on CO methanation was studied. More importantly, we investigated the effects of the addition of the promoters V, Ce, and Zr on the catalytic activity and physicochemical properties of the Ni-based catalysts. The catalysts were characterized through X-ray diffraction (XRD) measurements, X-ray photoelectron spectroscopy (XPS), Fourier transform infrared (FT-IR) spectroscopy, N2 physisorption, Raman spectroscopy, H2 temperature-programmed reduction (H2-TPR), and H2 temperature-programmed desorption (H2-TPD) experiments. We aimed to explain the high performance of the catalysts. We found that addition of promoters could improve the low-temperature activity and high-temperature stability of the catalysts in CO methanation. 10Ni–5V/SBA-15 is a promising catalyst because of its high catalytic activity and high stability.
2. Experimental
2.1 Catalyst preparation
2.1.1 SBA-15 synthesis. The SBA-15 support was prepared according to previously reported methods.22 Triblock copolymer (Pluronic P123, 5800 g mol−1 average, Sigma-Aldrich), a structure-directing agent, was dissolved in aqueous HCl solution (2 mol L−1). Tetraethyl orthosilicate (98% purity, Sigma-Aldrich), the silicon source, was subsequently added. The resulting mixture was stirred for 24 h at 40 °C and then hydrothermally treated at mixture systems for 24 h. The solid product was filtered, washed, oven-dried overnight at 80 °C, and finally calcined at 550 °C for 6 h in air to remove the structure-directing agent.
2.1.2 Preparation of the catalysts Ni–(V, Ce, Zr)/SBA-15. Ni–(V, Ce, Zr)/SBA-15 were prepared by ultrasonic coimpregnation method. To prepare Ni/SBA-15 catalysts with various Ni loadings, a predetermined amount of nickel nitrate hexahydrate (Ni(NO3)2·6H2O) was dissolved in ethanol/water at 65 °C for 20 min under vigorous stirring. After addition of a specified mass of SBA-15, the mixtures were ultrasonicated for 30 min at 65 °C to well dispersed in a ultrasonic container and then stirred continuously at constant temperature until the solvent was almost removed. Next, the sample was dried at 90 °C overnight and finally calcined in a muffle furnace at 500 °C for 3 h. To prepare the Ni/SBA-15 catalysts with various V, Ce, and Zr promoters, a known amount of SBA-15 was added to solutions of Ni(NO3)2·6H2O and Zr(NO3)4·5H2O (NH4VO3 or Ce(NO3)3·6H2O), respectively. Subsequent steps were the same as above. The catalysts are designated as xNi–mY/SBA-15 (Y = V, Ce, Zr), where x and m respectively represent the weight content of Ni and promoter species in the catalysts. All reagents were of AR grade (Sinopharm Chemical Reagent Co.).
2.2 Catalyst evaluation
The performance of the catalysts was studied by using a fixed-bed microreactor (40 cm length, 10 mm inner diameter). In each experiment, 0.4 g of catalyst sample (40–60 mesh) was used. Conditions consisted of reaction gases at a V(H2)/V(CO) ratio of 3
:
1 at 1.0 MPa and total gas flow rate of 100 mL min−1 (corresponding to a weight hourly space velocity (WHSV) of 15
000 mL g−1 h−1). Prior to the reaction, the catalysts were reduced in situ at 500 °C for 2 h under a stream of pure hydrogen (30 mL min−1). The activity in CO methanation at 180–400 °C was examined, and the catalytic stability within 60 h at 500 °C was evaluated. Reaction products were analyzed on a gas chromatograph (GC900SD) equipped with a TDX-01 column and a thermal conductivity detector (TCD). Formulas for CO conversion (XCO), CH4 selectivity (SCH4), and turnover frequency of CO methanation (TOFCO) were calculated according to eqn (1)–(3): |
XCO (%) = (moles of CO reacted) × 100/(moles of CO supplied)
| (1) |
|
SCH4 (%) = (moles of CH4 formed) × 100/(moles of CO reacted)
| (2) |
|
TOFCO (S−1) = (moles of CO converted per second)/(moles of surface Ni atoms)
| (3) |
2.3 Catalyst characterization
XRD measurements were performed on a Rigaku D/Max 2500 system equipped with a Cu Kα source (λ = 1.54056 Å) operated at 40 kV and 100 mA. The crystallite size was calculated by using the Scherrer formula.
N2 adsorption–desorption isotherms at −196 °C were obtained on a Micromeritics ASAP 2020 specific surface area and porosity analyzer. Before measurement, the samples were degassed under vacuum at 300 °C for 4 h. The microscopic feature of the samples was observed by transmission electron microscopy (TEM) (JEM-2010F, JEOL, Japan).
Raman spectroscopic data were obtained using a Renishaw inVia microlaser Raman spectrometer, and the Raman spectra were excited by an argon ion laser (514.5 nm wavelength) with experimental power of 5 mW at a step length of 1 cm−1 in the range between 200 and 1400 cm−1.
XPS measurements were performed with a V.G. Scientific ESCALAB250 using Al Kα radiation (hν = 1486.6 eV). The instrument was calibrated internally by using the carbon deposit C(1s) (Eb = 284.7 eV).
FT-IR spectra of the samples were obtained on a Bruker TENSOR 27 spectrometer. Each 1 mg powdered sample was diluted with vacuum-dried IR-grade KBr.
H2-TPR experiments were carried out with a Micromeritics AutoChem II 2920 analyzer. In a typical measurement, 50 mg of sample (40–60 mesh) was loaded into a U type quartz tube. A stream of pure Ar at 300 °C was introduced for 30 min to remove adsorbed H2O and other volatile gases on the sample. The sample was then cooled to room temperature, and the gas was switched to a gaseous Ar mixture containing 10 vol% H2 at a flow rate of 20 mL min−1. Finally, the temperature was elevated to 900 °C at a rate of 10 °C min−1. H2 consumption was detected by using a TCD.
H2-TPD experiments were carried out in the same system that was used in the H2-TPR measurements. After the catalysts were reduced, H2 desorption patterns were obtained from room temperature to 860 °C at a heating rate of 10 °C min−1. The desorbed H2 was detected by using a TCD. The number of surface Ni sites per unit mass of catalyst was calculated from the H2-TPD results by assuming a H/Ni adsorption stoichiometry of 1
:
1. The peak area of the H2-TPD profile was normalized according to that of the H2-TPR profile of a standard CuO sample. The Ni dispersion (D) was calculated from eqn (4):
|
D (%) = (2 × Vad × M × SF) × 100/(m × P × Vm × dr)
| (4) |
where
Vad (mL) represents the volume of chemisorbed H
2 at standard temperature and pressure (STP) from TPD measurements.
M is the molecular weight of Ni (58 × 69 g mol
−1), and SF is the stoichiometric factor (the Ni/H molar ratio for chemisorption) which is taken as equal to 1.
m is the sample weight (g);
P is the weight fraction of Ni in the sample;
Vm is the molar volume of H
2 (22
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif)
414 mL mol
−1) at STP; and
dr represents the degree of reduction of nickel species calculated from the H
2-TPR profile.
3. Results and discussion
3.1 Characterization of catalysts
3.1.1 XRD measurements. XRD patterns of the samples are shown in Fig. 1. The reduced catalysts 10Ni/SBA-15 and 10Ni/SBA-15 with promoters produced XRD peaks corresponding to metallic Ni crystals (JCPDS no. 65-2865) at 2θ values of 44.370°, 51.890°, and 76.410°. These peaks are respectively attributed to the (111), (200), and (220) planes of metallic Ni particles. There is a broad amorphous silica peak at around 23° for these samples. In addition, the intensities of peaks for metallic Ni phase in the catalysts with promoters are weaker than those of the 10Ni/SBA-15 catalyst. The full width at half maximum of the peak for Ni0 in the promoter-containing catalysts increased, suggesting that the promoters could effectively enhance Ni dispersion.
 |
| Fig. 1 XRD profiles of the reduced 10Ni/SBA-15, 5M/SBA-15 and 10Ni–5M/SBA-15 (M = V, Ce, Zr) samples. | |
The average crystalline sizes of metallic Ni in the catalysts (Table 1) were calculated from the XRD patterns through the Scherrer equation. Generally, highly dispersed Ni species lead to more numerous active sites on the catalysts. No new diffraction peaks on the promoter species appeared for all samples including the 5V/SBA-15, 5Ce/SBA-15 and 5Zr/SBA-15 reduced samples, indicating that the promoter species were highly dispersed on the surface of the SBA-15 support.
Table 1 Physicochemical properties of catalysts
Sample |
Surface areaa (m2 g−1) |
Pore volumeb (cm3 g−1) |
Pore sizec (nm) |
Ni average crystallite sized (nm) |
Ni dispersione (%) |
TOFCOf, 220 °C (×10−3 S−1) |
Calculated by the BET equation. Pore volume obtained from the volume of nitrogen adsorbed at the relative pressure of 0.997. BJH desorption average pore size. Calculated by the XRD diffraction peak using the Scherrer equation. Ni dispersion calculated based on H2-TPR and H2-TPD. Calculated based on the Ni dispersion and the CO conversion at 220 °C. |
SBA-15 |
594.6 |
0.99 |
5.7 |
— |
— |
— |
10Ni/SBA-15 |
513.3 |
0.81 |
5.1 |
13.1 |
15.2 |
2.71 |
10Ni–5V/SBA-15 |
459.0 |
0.66 |
5.2 |
8.6 |
21.3 |
5.50 |
10Ni–5Ce/SBA-15 |
453.3 |
0.71 |
5.3 |
10.3 |
19.9 |
3.62 |
10Ni–5Zr/SBA-15 |
462.3 |
0.74 |
5.0 |
10.1 |
20.3 |
4.35 |
3.1.2 FT-IR spectroscopy. Fig. 2 shows the FT-IR spectra of the samples. Asymmetrical stretching vibration of the Si–O–Si bond at about 1230 and 1091 cm−1 and symmetrical stretching vibration at 803 cm−1 may be discerned from the SBA-15 spectrum. The band at 961 cm−1 is assigned to the asymmetrical stretching vibration of Si–OH (silanol group). The band at 463 cm−1 corresponds to the bending vibration of Si–O bonds in a ring structure.23 The asymmetric stretching vibration of the Si–O–Si bond of other samples shifted slightly toward lower wavenumbers relative to that in the SBA-15 spectrum (Fig. 2). These changes in asymmetric stretching vibration of the Si–O–Si bond are caused by addition of Ni and promoter species that are partially incorporated into the porous structures of SBA-15. The slightly lower intensity for the stretching vibration of the Si–OH bond in the Ni/SBA-15 sample at 961 cm−1 compared with that for the stretching vibration in the SBA-15 sample suggests that the presence of Ni species in SBA-15 causes a structural change in the surface Si–OH group.21
 |
| Fig. 2 FT-IR spectra of the samples SBA-15, 10Ni/SBA-15, 20Zr/SBA-15 and 10Ni–5M/SBA-15 (M = V, Ce, Zr). | |
Moreover, bands at 961 cm−1 of other samples with promoters correspond to vibrations of the Si–OH bond and it can be proved by the peak intensity change of 20Zr/SBA-15 at 961 cm−1. From the FT-IR spectra, it was observed that the peak intensity of 20Zr/SBA-15 at 961 cm−1 almost disappeared compared to SBA-15. The behavior indicates that the Zr–OH bond for the sample with the high Zr species loading almost consumes the Si–OH groups. The samples with promoters have smaller shoulder peaks and lower intensities compared with those in the spectra of SBA-15 and Ni/SBA-15. These differences are caused by the disappearance of Si–OH groups and the appearance of (Si–O)nM
O species, suggesting that partial Si–OH groups were changed or consumed and transformed to the Si–O–M bond (M = Zr,24,25 V,26,27 Ce (ref. 28 and 29)).
3.1.3 Texture features. Fig. 3 presents the nitrogen adsorption–desorption isotherms and pore size distribution curves of the samples, and Table 1 lists details of the physical properties of the samples. These samples produced typical type IV isotherms with hysteresis loops, indicating that regular mesoporous structures existed in the samples.30 The samples have high surface areas and pore volumes. SBA-15 loaded with metal species produced hysteresis loops that are markedly smaller than those of SBA-15, corresponding to a decrease in the specific surface area and pore volume. The specific surface area and pore volume of the SBA-15 support were 594.6 m2 g−1 and 0.99 cm3 g−1, respectively. In addition, the 10Ni/SBA-15 catalyst had a pore size of 5.1 nm smaller than that of SBA-15, implying that some Ni species may enter and block the pores. The hysteresis loops for the catalysts with V, Ce, and Zr promoters are similar to an H1-type loop.31,32 The catalysts with V, Ce, and Zr had smaller specific surface area and pore volume compared with those of Ni/SBA-15. These features indicate that the promoters occupy the support of the catalyst. Comparing to the SBA-15, not only the 10Ni/SBA-15 has double pores, but also the SBA-15 with V, Zr, Ce promoter show double pores, perhaps it can be revealed that addition of the species to SBA-15 contribute to the formation of the second pore. A possible explanation: the pore size at ∼6 nm is consistent with that of pure SBA-15 support, the second pore at ∼4 nm is caused by the partial Ni and promoter species located in the channel of SBA-15. Furthermore, the presence of the second pore demonstrates that the Ni species successfully embedded in the pores of mesoporous silica, which can improve the catalytic activity and stability.21 From the TEM images in Fig. 4, it is observed that the ordered mesostructure of the samples isn't destroyed by addition of active metallic and promoters. The pore size of the samples is ca. 5–6 nm, and there are smaller Ni species crystals located inside the pores of the SBA-15. Moreover, the crystalline size of Ni species for 10Ni–5V/SBA-15 sample is smaller than the 10Ni/SBA-15. The results is in good agreement with the conclusions of XRD and N2 physisorption.
 |
| Fig. 3 The nitrogen adsorption–desorption isotherms (a) and pore size distribution curves (b) of the samples. | |
 |
| Fig. 4 TEM images of the 10Ni/SBA-15 (a) and (b) and 10Ni–5V/SBA-15 (c) and (d) catalysts. | |
3.1.4 Raman spectroscopy. The Raman spectra of the samples are presented in Fig. 5, the Ni/SBA-15 sample exhibits the Raman bands at 485, 807, and 970 cm−1, which are assigned to tri-cyclosiloxane rings, siloxane bridges, and surface silanol groups, respectively.33 For the Ni–V/SBA-15 sample, the 773, 846 and 956 cm−1 bands may correspond to new VOx species formed on the surface of the catalyst.34 The band at 1025 cm−1 is ascribed to the V
O bond stretching vibration of the isolated tetrahedral VO2 species anchored on SBA-15. The shoulder band at 1065 cm−1 is attributed to the perturbed silica vibrations for Si(–O−)x functionalities, suggesting the formation of V–O–Si bond.35 The Raman band at 952 cm−1 for the Ni–Zr/SBA-15 sample is assigned to the Si–O–Zr linkages,36 indicating that some Zr species interact with Si–OH to form Si–O–Zr bond. In addition, according to the literature,37 the bands at 760, and 1147 cm−1 may correspond to the intermediate phase of zirconia, and no Raman bands about tetragonal zirconia, monoclinic zirconia and zircon are observed. The results demonstrate that some Zr species on SBA-15 are amorphous zirconia or intermediate phase of zirconia, and are probably present as highly dispersed surface species. For the Raman spectrum of Ni–Ce/SBA-15 sample, the bands at 573 cm−1, 1098 cm−1 are assigned to the defect-induced mode and second-order longitudinal optical mode of CeO2,38 respectively. The band at ∼462 cm−1 attributed to the F2g mode of the cerium oxide overlaps with the Raman band of tri-cyclosiloxane rings. Based on the Raman features of Si–OH groups and Si–O–M (M = V, Zr) linkages, it can be concluded that the 983 cm−1 band may attribute to the Si–O–Ce linkages. In a word, addition of the promoter species can lead to partially M–OH species combine with Si–OH groups to form Si–O–M bond, which is in agreement with the results of FT-IR spectroscopy. Interestingly, it is the formation of Si–O–M bond that increases the interaction between active species and support (see Section 3.1.6), which facilitates CO methanation.
 |
| Fig. 5 Raman spectra of the samples Ni/SBA-15, and Ni–M/SBA-15 (M = V, Ce, Zr). | |
3.1.5 XPS. Fig. 6(a) shows Ni 2p3/2 XPS spectra of the reduced and used catalysts. Two Ni states in all catalysts can be inferred according to their binding energies (BE) in the Ni 2p3/2 XPS spectra. The Ni 2p3/2 peak of reduced 10Ni/SBA-15 at 852.8 eV (BE) is attributed to Ni0 species, and the Ni 2p3/2 main peak at 855.8 eV with a shake-up satellite peak at 816.3 eV is assigned to the Ni2+ species.19,39,40 For the 10Ni–5V/SBA-15 and the 10Ni–5Ce/SBA-15 reduced catalysts, the Ni 2p3/2 binging energies show slight chemical shifts relative to that of Ni/SBA-15 reduced catalyst (Table 2). These shifts suggest that addition of V, Ce promoters could reduce the binding energy of Ni 2p3/2 and increase the electronic density of nickel on the catalyst surface. The relative amounts of the various Ni oxidation states on the surface of the reduced catalysts is shown in Table 3. Compared with those of reduced 10Ni/SBA-15, the relative amounts of metallic Ni in the promoter-containing catalysts were higher. In particular, the relative amount of metallic Ni in reduced 10Ni–5V/SBA-15 was as high as 30.6%. Therefore, the promoter-containing catalysts had more numerous active Ni sites. In addition, it is observed that the Ni 2p3/2 binging energies and the relative amount of metallic Ni of the used catalysts have almost no changes relative to their reduced counterparts. The behaviors suggest that there are no obvious changes in the Ni status (Ni species valence, Ni dispersion and number of Ni active sites) of the catalysts after the lifetime test.
 |
| Fig. 6 XPS spectra for the samples: (a) Ni 2p3/2; (b) V 2p3/2 and Ce 3d; (c) Zr 3d and O 1s. | |
Table 2 Binding energy values of the samples for XPS spectra
Samples |
Binding energy/eV |
Ni 2p3/2 |
Zr 3d5/2 |
Ce 3d5/2 |
V 2p3/2 |
Ni2+ |
ZrI species (Zr4+) |
Ce4+ |
Ce3+ |
V5+ |
V4+ |
V3+ |
The error measurements is ±0.1 eV. |
10Ni/SBA-15-reduced |
855.8a |
— |
— |
— |
— |
— |
— |
10Ni/SBA-15-used |
855.8 |
— |
— |
— |
— |
— |
— |
5Ce/SBA-15-reduced |
— |
— |
882.4 |
885.5 |
— |
— |
— |
10Ni–5Ce/SBA-15-reduced |
855.6 |
— |
882.5 |
885.8 |
— |
— |
— |
10Ni–5Ce/SBA-15-used |
855.6 |
— |
882.5 |
885.9 |
— |
— |
— |
5Zr/SBA-15-reduced |
— |
182.2 |
— |
— |
— |
— |
— |
10Ni–5Zr/SBA-15-reduced |
855.7 |
182.2 |
— |
— |
— |
— |
— |
10Ni–5Zr/SBA-15-used |
855.7 |
182.2 |
— |
— |
— |
— |
— |
5V/SBA-15-reduced |
— |
— |
— |
— |
517.4 |
516.0 |
515.1 |
10Ni–5V/SBA-15-reduced |
855.5 |
— |
— |
— |
517.6 |
516.4 |
515.5 |
10Ni–5V/SBA-15-used |
855.5 |
— |
— |
— |
517.6 |
516.4 |
515.5 |
Table 3 Surface Ni, Zr, O compositions of the samples for the XPS spectra
Catalysts |
Ni 2p3/2 |
Zr 3d5/2 |
O 1s |
Ni0 |
Ni2+ |
ZrI species |
ZrII species |
OI species |
OII species |
Percentage of Ni, Zr and O species. |
10Ni/SBA-15-reduced |
16.9a |
83.1a |
— |
— |
— |
— |
10Ni/SBA-15-used |
16.4 |
83.6 |
— |
— |
— |
— |
5Ce/SBA-15-reduced |
— |
— |
— |
— |
— |
— |
10Ni–5Ce/SBA-15-reduced |
25.0 |
75.0 |
— |
— |
— |
— |
10Ni–5Ce/SBA-15-used |
24.2 |
75.8 |
— |
— |
— |
— |
5Zr/SBA-15-reduced |
— |
— |
79.6a |
20.4a |
81.2a |
18.8a |
10Ni–5Zr/SBA-15-reduced |
25.8 |
74.2 |
74.5 |
25.5 |
75.7 |
24.3 |
10Ni–5Zr/SBA-15-used |
26.5 |
73.5 |
74.2 |
25.8 |
75.5 |
24.5 |
5V/SBA-15-reduced |
— |
— |
— |
— |
— |
— |
10Ni–5V/SBA-15-reduced |
30.6 |
69.4 |
— |
— |
— |
— |
10Ni–5V/SBA-15-used |
30.1 |
69.9 |
— |
— |
— |
— |
The V 2p3/2 XPS spectrum of reduced 10Ni–5V/SBA-15 catalyst was fitted with peaks at 517.6, 516.4, and 515.5 eV (BE) in Fig. 6(b). These peaks correspond to V5+, V4+, and V3+ (ref. 41–43), respectively. It is observed that the 10Ni–5V/SBA-15 reduced catalyst has higher V 2p3/2 binding energy (Table 2 and Fig. 6(b)) relative to the 5V/SBA-15 sample, implying the occurrence of electron transfer from V species to Ni,44 and which further enhances the Ni electron cloud density. As the V4+ peaks are broader than those of V5+ and V3+ (Fig. 6(b)), high-valence V species can undergo reduction to V4+ and V3+ by reaction gas during CO methanation.42 Mori et al.45 observed that the promoter V plays an important role in CO dissociation by interacting with charged V3+ centers and with oxygen atoms in chemisorbed CO. It was observed that the portion of V4+ in 10Ni–5V/SBA-15 used catalyst was increased compared with that in the 10Ni–5V/SBA-15 reduced catalyst, which should be caused by the oxidation of V3+ or the reduction of V5+ in CO methanation, and it has been proved by previous literature data.46 In addition, based on the XPS, XRD, Raman spectroscopy, and H2-TPR (see Section 3.1.6), there are the metallic Ni, Ni2+ species (nickel oxide), and the VOx species (V2O5, VO2, V2O3) on the surface of the 10Ni–5V/SBA-15 reduced catalyst. Combined with the results of Ni 2p3/2 XPS spectra, it can be found that the addition of V promoter reduces the binding energy of Ni 2p relative to that of Ni/SBA-15 catalyst (Ni 2p from 855.8 for the 10Ni/SBA-15 catalyst to 855.5 for the 10Ni–5V/SBA-15), which may be caused by the coexistence of the V3+, V4+, and V5+ that increase the electron cloud density of metallic Ni atoms (Ni0 → Niδ−). Thus, V3+, V4+, and V5+ that coexist in reduced 10Ni/5V–SBA-15 could trigger electron transfer and promote CO dissociation through the oxidation–reduction shift cycle (V3+ ↔ V4+ ↔ V5+). These behaviors confirm that the catalyst 10Ni/5V–SBA-15 has high activity and high CH4 selectivity.
XPS spectra of Ce 3d of the promoter-containing samples are shown in Fig. 6(b). For the high-resolution spectra of Ce 3d5/2 and Ce 3d3/2 of the samples with Ce promoter, ionization features were fitted with eight Gaussian distributions representing different states in the Ce 3d core-level XPS spectrum. Some investigations based on Ce 3d XPS analysis have revealed that fit for the Ce 3d Gaussian peak corresponds to Ce3+ and Ce4+ states.47–49 Peaks in the Ce 3d XPS spectrum of the 10Ni–5Ce/SBA-15 reduced catalyst at 907.8 and 916.7 eV (BE) are ascribed to Ce 3d3/2 ionization of Ce4+. The peak at 903.7 eV (BE) is assigned to Ce3+ 3d3/2. Bands at 882.5, 884.3, 898.6, and 900.9 eV (BE) correspond to Ce4+ 3d5/2, and the peak at 885.8 eV (BE) is attributable to Ce3+ 3d5/2. So based on the XPS, XRD, Raman spectroscopy, and H2-TPR (see Section 3.1.6), the surface of the 10Ni–5Ce/SBA-15 reduced catalyst shows the presence of metallic Ni, Ni2+ species (nickel oxide), highly dispersed CeO2, and the reduced Ce3+ species. Moreover, it can be found that Ce species with various oxidation states (Ce3+ and Ce4+) co-exist in the reduced catalysts, which is attributed to the partial reduction of Ce4+ to Ce3+ during catalyst reduction. Therefore, in the activation process, it can be concluded that the CeO2 species migrating to metallic Ni surface will be partially reduced into Ce3+ due to the catalytic action of Ni and H spillover, which can trigger electron migration and increase the density of the electron cloud of Ni species. These behaviors ultimately improve the ability of CO to dissociate and the catalytic performance.
Fig. 6(c) shows that Zr 3d and O 1s XPS spectra of the samples with Zr species. All the samples exhibit a spin–orbit doublet of the Zr 3d core level into 3d5/2 and 3d3/2 levels. It can be observed that there is an energy gap of 2.4 eV between Zr 3d5/2 and Zr 3d3/2, implying the existence of ZrO2-like species.50 Moreover, XPS peaks of Zr 3d core level correspond to two kinds of zirconium species, which are referred as species I (ZrI) with low BE in the range 181.9–182.4 eV and species II (ZrII) with higher BE in the range 183.2–183.9 eV, respectively. Among two kinds of zirconium species, the binding energy of ZrI species in the reduced samples is similar to Zr4+ in pure ZrO2. However, the binding energy of Zr4+ (182.2) in the reduced samples has slightly lower value relative to that of stoichiometric ZrO2 (182.6 eV) and a shift of 0.4 eV between them indicates the existence of some oxygen deficiency,51 which are conducive to CO dissociation. The position of ZrI species shifts toward the lower binding energy may be associated with the holes created by oxygen vacancies in the zirconia lattice.52 The ZrII species (183.2–183.9 eV) is attributed to the formation of partially reduced Zrδ+ sites. Meanwhile, the integrated area ratio of ZrII/(ZrII + ZrI) for the 10Ni–5Zr/SBA-15 reduced sample (25.5%) in Table 3 is higher than that for the 5Zr/SBA-15 reduced sample (20.4%), revealing the increase of oxygen vacancies.51 The peak of the O 1s can be fitted into two peaks in Fig. 6(c): the peak at 531.1 eV corresponds to the OI species (lattice oxygen), the peak at 532.5 eV belongs to the OII species (adsorbed oxygen), respectively; and the 10Ni–5Zr/SBA-15 reduced catalyst has higher integrated area ratio of OII/(OII + OI) than that of the 5Zr/SBA-15 (from 18.8% to 24.3%), the behavior also implies an increase in the number of oxygen vacancies, which is consistent with the conclusion of the ZrII/(ZrII + ZrI). Comparing to the 10Ni–5Zr/SBA-15 reduced catalyst, there are no obvious changes for Zr 3d and O 1s spectra of the 10Ni–5Zr/SBA-15 used catalysts, suggesting that the Zr and O status of 10Ni–5Zr/SBA-15 catalyst surface has little changes after the lifetime test. In addition, combined with the results of XRD, Raman spectroscopy, and H2-TPR (see Section 3.1.6), it can be concluded that metallic Ni, Ni2+ species (nickel oxide), amorphous zirconia or intermediate phase of zirconia containing partially reduced Zrδ+ species are present on the surface of the 10Ni–5Zr/SBA-15 reduced catalyst. In a word, Zr species contributed to the generation of the oxygen vacancies during calcination and reduction, thereby enhancing the ability of CO to adsorb and dissociate. CO adsorbed on an oxygen vacancy and subsequently reacted with hydrogen, producing CH4.6,47,53 These observations show that Zr could improve the catalytic activity and CH4 selectivity in CO methanation.
The addition of V, and Ce promoters reduce the binding energy of Ni 2p3/2 (Fig. 6 and Table 2), and it is found that the catalytic performances of the Ni/SBA-15 catalysts with promoters are significantly improved at 240 °C with the decrease of the binding energy of Ni 2p3/2 relative to the Ni/SBA-15. As we all know, there are 9.4 electrons existing in the 3d electron orbit of nickel atoms and 0.6 electron in the 4s electron orbit, thus, the 0.6 electron hole of the 3d electron orbit of nickel atoms can accept the electrons of other atoms or ions. On one hand, during the activation process, the promoter species MxOy (M = V, Ce) can transfer to the Ni0 surface, which will be partially reduced into low valence (V<5+, Ce<4+) on account of catalytic action of Ni and H spillover. The behaviors can reduce the binding energy of Ni 2p3/2 and increase the electron density of Ni0 species. On the other hand, in the reaction process, the increase of the electron density of Ni0 species (Ni0 → Niδ−) can enhance the Ni–C bond and weaken the C–O bond of the Ni–C–O bond,54 which improve the ability of CO to dissociate. In a word, a plausible explanation for the enhancement of CO methanation activity over promoter-containing catalysts is summarized as follows: changes in valences of the promoter species V and Ce (Mx+ ↔ My+, M = V, Ce) can trigger electron transfer and can enhance the electron density of Ni0 species (Ni0 → Niδ−), which weaken the C–O bond of the Ni–C–O bond. The promoter Zr species can induce the formation of oxygen vacancies during calcination and reduction. Such behaviors improve the ability of CO to dissociate and thus enhance the activity and selectivity of the catalysts with V, Ce, and Zr promoters in CO methanation.
3.1.6 H2-TPR measurements. The interaction between NiO species and supports was investigated through H2-TPR measurements. Fig. 7 presents H2-TPR profiles of the samples, for the 5V/SBA-15 sample, the reduction peak at 400–600 °C corresponds to the reduction of VOx species, and the amount of H2 consumption during the reduction process indicates that the reduction of V2O5 to V2O3 occurs in this temperature range. The 5Ce/SBA-15 has reduction peak at 495 °C, which is attributed to the reduction of ceria species. It is observed that the 5Zr/SBA-15 has only a weak reduction peak at high temperature, which belongs to the partial reduction of Zr oxide species. For the catalysts, according to the peak positions, reduction of NiO species of the catalysts can be divided into three stages, namely, α-, β-, and γ-stages. The α-stage is assigned to the reduction of free NiO species interacting weakly with the support in the low-temperature range (300–360 °C). The β-stage is attributed to NiO species that moderately interact with the support. These species can be reduced between 370–510 °C. The γ-stage corresponds to NiO species that strongly interact with the support and undergo reduction at high temperature (520–630 °C). We found that Tmax of reduction peaks of the three stages over the promoter-containing catalysts is slightly shifted to higher temperature relative to that of the 10Ni/SBA-15 catalyst. This shift is consistent with a previous finding55 and implies stronger interaction between the NiO species and support. The result may be associated with the formed Si–O–M bond (confirmed by the FT-IR and Raman) via improving the acid strength and amounts of acid sites of the support can affect and increase the interaction between active species and the support,17,26,29 thereby enhancing the activity and stability of the catalysts. Moreover, the dispersion of active metal on the support is strongly influenced by metal-support interaction. Areas of the reduction peaks of the catalysts with promoters are markedly larger than those of 10Ni/SBA-15. This difference suggests that the promoters could effectively improve Ni dispersion. Therefore, catalysts with the promoters V, Ce, and Zr possessed higher Ni dispersion, and more numerous active Ni species after reduction. Among the catalysts, 10Ni–5V/SBA-15 showed the highest Ni dispersion.
 |
| Fig. 7 H2-TPR profiles of the 10Ni/SBA-15, 5M/SBA-15 and 10Ni–5M/SBA-15 (M = V, Ce, Zr) samples. | |
3.1.7 H2-TPD measurements. Fig. 8 shows H2-TPD profiles of all samples. 10Ni/SBA-15 produced only two peaks in the low-temperature range (<400 °C). In contrast, catalysts with V, Ce, and Zr promoters produced several peaks at lower (<400 °C) and higher (>400 °C) temperatures. The lower-temperature peaks can be assigned to H2 that is weakly chemisorbed on a surface with highly dispersed Ni and a high density of defects. Such defects can act as traps during surface hydrogen diffusion and thus reduce the activation energy of H2 dissociation.56 The higher-temperature peaks can be due to H2 that is strongly chemisorbed to the catalyst surface. It may also be due to desorption of hydrogen adsorbed on promoters with low valence or spillovered hydrogen, which enhance the storage capacity for H2.57–59 However, for the SBA-15 samples with promoters, there is no H2 desorption peak for 5Zr/SBA-15 samples, but the 5V/SBA-15 and 5Ce/SBA-15 samples show a weak H2 desorption peak at high temperature, which may be associated with the spillover of H2.
 |
| Fig. 8 H2-TPD profiles of the 10Ni/SBA-15, 5M/SBA-15 and 10Ni–5M/SBA-15 (M = V, Ce, Zr) samples. | |
Integrated areas for the catalysts with various promoters are markedly larger than that of the 10Ni/SBA-15 catalyst, demonstrating that catalysts with promoters had higher Ni dispersion and that more surface-dissociated hydrogen formed on the catalysts. Table 1 lists the dispersion of Ni calculated from the H2-TPD and H2-TPR results. Since 10Ni–5V/SBA-15 had the strongest ability for H2 adsorption and highest Ni dispersion (21.3%), the promoter V had the best promoting effect. Luo et al.57,60 found that H2 chemisorbed on the Rh surface at high temperature over Rh–V/SiO2 catalyst could reach and remain in the lower-valence V. During desorption, the stored hydrogen could flow back to Rh and then desorb from Rh. A similar observation in Ni–V/Al2O3 catalyst was reported by Liu et al.19 On the basis of these studies and combined with the results of H2-TPD, it is hypothesized that the promoter Ce plays a similar role as V in causing the spread of H2 from Ni to low-valence promoter species and in enhancing hydrogen storage and migration ability, which eventually improve the catalytic activity and selectivity (see Section 3.2.3). However, this phenomenon is not evident in the case of Zr promoter.
3.2 CO methanation over Ni–M/SBA-15
3.2.1 Catalytic activities. Fig. 9 shows the catalytic performance of Ni/SBA-15 catalysts in CO methanation at 1.0 MPa at a WHSV of 15
000 mL g−1 h−1 at 180–400 °C. As described in Fig. 9, 5Ni/SBA-15 had low CH4 selectivity at 180–400 °C, and the amount of Ni species supported on the SBA-15 support was >5 wt%. The CO conversion and CH4 selectivity was markedly greater than that of 5Ni/SBA-15. Thus, the catalytic activity and CH4 selectivity increased with the increase in Ni loading; but when the amount of Ni species reached 20 wt%, the catalytic performance has almost no change. Furthermore, SBA-15 had larger specific surface area, which increased Ni dispersion and increased the number of active sites of catalysts with low Ni loading and thus improved their catalytic performance. The CO conversion and CH4 selectivity on 10Ni/SBA-15 reached 99.8% and 89.2%, respectively. Notably, in terms of the stability test, the 10Ni/SBA-15 catalyst with low loading exhibits superior stability than 25Ni/SBA-15 (Fig. 9(c)), which is caused by that active Ni species with high loading is more likely to sintering. So, 10% of Ni on SBA-15 is chosen as the optimum loading for the further study of the role of promoters.
 |
| Fig. 9 CO conversion (a) and CH4 selectivity (b) of the Ni/SBA-15 catalysts with different Ni loadings (reaction conditions: P = 1 MPa, WHSV = 15 000 mL g−1 h−1, H2/CO = 3), (c) 60 hour stability results of the 10Ni/SBA-15 and 25Ni/SBA-15 catalysts (reaction conditions: P = 1 MPa, WHSV = 15 000 mL g−1 h−1, H2/CO = 3). | |
3.2.2 Effects of V, Ce, and Zr promoters. Test results for catalysts with V, Ce, and Zr promoters are shown in Fig. 10. All catalysts exhibited higher activity, and the temperature for maximum CO conversion shifted to a temperature lower than that for maximum CO conversion over 10Ni/SBA-15. These changes indicate that the promoters improved the catalytic activity of Ni/SBA-15 catalysts. Based on Zr promoter, catalysts with various Zr loadings (3, 5, and 7 wt%) were used in CO methanation. When the Zr loading was 5 wt%, the 10Ni–5Zr/SBA-15 catalyst displayed the highest activity, and CH4 selectivity reached 93.6%; we could achieve higher Ni dispersion and greater number of active sites than those of catalysts with other Zr loadings (Fig. 10). It is worth noting that the decrease in activity and CH4 selectivity with increasing Zr loading may be ascribed to partial coverage of Ni active sites. Furthermore, we investigated the activity and CH4 selectivity of the catalysts with V and Ce promoters (5 wt%). CO conversion over 10Ni–5Ce/SBA-15 reached 99.8%, and the CH4 selectivity reached 92.1%. 10Ni–5V/SBA-15 exhibited the highest activity and CH4 selectivity (99.9% CO conversion, 95.5% CH4 selectivity). Compared with that of 10Ni/SBA-15, the performance of the catalysts with V, Ce, and Zr promoters significantly increased at 240 °C, indicating that catalysts with promoters exhibited high catalytic performance at low temperature.
 |
| Fig. 10 CO conversion (a) and CH4 selectivity (b) of the Ni/SBA-15 catalysts with V, Ce, Zr promoters (reaction conditions: P = 1 MPa, WHSV = 15 000 mL g−1 h−1, H2/CO = 3). | |
On the basis of detailed characterization, the improved catalytic performance of Ni/SBA-15 catalysts with V, Ce, and Zr promoters can be explained by several main factors. On one hand, formation of the Si–O–M bond (M = V, Ce, Zr) via addition of promoters can increase the interaction between Ni species and support. On the other hand, promoter species undergoing the oxidation–reduction shift cycle (Mx+ ↔ My+, M = V, Ce) can trigger electron transfer and significantly increase the electron cloud density of nickel species (Ni0 → Niδ−), thereby weakening the C–O bond of the Ni–C–O bond and improving the ability of CO to dissociate. Compared with promoter Ce species, the variation among different valences of V can produce stronger electronic effect during the oxidation–reduction shift cycle process (confirmed by XPS) since V metal oxide species are more easily reduced from high valence to low valence,18 consequently making the Ni/SBA-15 catalyst exhibits better catalytic performance. The promoter Zr species could cause the formation of oxygen vacancies during calcination and reduction, thus enhancing the ability of CO to adsorb and dissociate. Furthermore, all promoters could improve Ni dispersion, thereby increasing the number of active Ni species and the ability of H2 to adsorb. In particular, V contributed the greatest improvement. In summary, Ni/SBA-15 catalysts with V, Ce, and Zr promoters exhibited higher catalytic performances than Ni/SBA-15. Among the catalysts, 10Ni–5V/SBA-15 showed the highest activity and CH4 selectivity.
3.3 Apparent activation energies
Sehested et al.61 reported that CO dissociation at the nickel active sites is the rate-determining step of CO methanation. During methanation, CO adsorbed on the metallic Ni atom surface dissociated and formed adsorbed carbon, which then combined with activated H and formed CH4. Thus, the apparent activation energy of CO dissociation acted as the reaction activation energy. The relation between the logarithm of the activity of CO methanation and the reciprocal of temperature (Fig. 12) is described by the Arrhenius formula. The apparent activation energies of CO dissociation over the Ni-based catalysts are obtained from the slope of the trend lines. As given in Fig. 12, the apparent activation energies over various catalysts decrease in the order 10Ni/SBA-15 (131.72 kJ mol−1) > 10Ni–5Ce/SBA-15 (120.36 kJ mol−1) > 10Ni–5Zr/SBA-15 (115.61 kJ mol−1) > 10Ni–5V/SBA-15 (103.53 kJ mol−1). Therefore, addition of promoters enhanced the ability of CO to dissociate and reduced reaction energy barriers. XRD and XPS results together suggest that addition of V, Ce, and Zr promoters could improve the Ni dispersion and increase the Ni0 content of the catalyst surface, thus enhancing the active sites of metallic Ni. Changes in the valences of V and Ce promoter species could initiate electron transfer and could increase the electron cloud density of Ni0 species (Ni0 → Niδ−), thereby weakening the C–O bond of the Ni–C–O bond, increasing the ability of CO to dissociation, and reducing the reaction energy barriers.
 |
| Fig. 12 Arrhenius plots for CO methanation on the 10Ni/SBA-15 and 10Ni–5M/SBA-15 (M = V, Ce, Zr) catalysts. | |
Table 1 lists the calculated TOF values of the catalysts in CO methanation at 220 °C at low level of CO conversion and at high WHSV (15
000 mL g−1 h−1). TOF values of CO methanation over the catalysts increase follow the order 10Ni/SBA-15 (2.71 × 10−3 s−1) < 10Ni–5Ce/SBA-15 (3.62 × 10−3 s−1) < 10Ni–5Zr/SBA-15 (4.35 × 10−3 s−1) < 10Ni–5V/SBA-15 (5.50 × 10−3 s−1), further confirming that addition of promoters enhanced the activity and CH4 selectivity of the catalysts in CO methanation.
4. Conclusions
Highly efficient Ni–(V, Ce, Zr)/SBA-15 catalysts for CO methanation were prepared by ultrasonic coimpregnation method. We examined the performance of the catalysts at 1.0 MPa at a WHSV of 15
000 mL g−1 h−1 at 180–400 °C. Notably, 10Ni/SBA-15 catalysts with V, Zr, and Ce promoters showed catalytic performances higher than that of 10Ni/SBA-15. Among the catalysts, 10Ni–5V/SBA-15 exhibited the highest catalytic activity and CH4 selectivity (99.9% CO conversion, 95.5% CH4 selectivity) and displayed high stability in a 60 h life test at 550 °C.
Characterization using H2-TPR, H2-TPD, XPS, and XRD measurements revealed that promoters enhanced the Ni dispersion. This enhancement led to enhanced exposure of Ni active sites, which improved the ability of H2 to adsorb and dissociate, accounting for the excellent catalytic performances of the catalysts in CO methanation. Addition of V and Ce promoters to Ni/SBA-15 catalysts could effectively improve the ability of CO to dissociate via electronic effect in methanation. Zr promoter species led to the formation of oxygen vacancies during calcination and reduction, thereby increasing the ability of CO to adsorb and dissociate in CO methanation. These behaviors could lower the reaction energy barriers and could increase the catalytic activity and CH4 selectivity. Results of FT-IR spectroscopy, Raman spectroscopy and H2-TPR measurements, revealed that addition of promoters increased the interaction between the active species and support via the Si–O–M bond. All of the promoters could prevent sintering of active species, thereby contributing to enhancement of the activity and stability of the Ni–(V, Zr, Ce)/SBA-15 catalysts.
Because of its smallest Ni particle size, highest Ni dispersion, greatest number of Ni active sites, strongest H2 adsorption ability, and strongest electronic effect, 10Ni–5V/SBA-15 had higher catalytic activity, as well as CH4 selectivity and stability in CO methanation. Therefore, 10Ni–5V/SBA-15 may be a promising catalyst for CO methanation.
Acknowledgements
This work has been supported by a grant from the National Natural Science Foundation of China (21376159 and 21276169).
References
- G. A. Mills and F. W. Steffgen, Catalytic Methanation, Catal. Rev., 2006, 8, 159–210 Search PubMed.
- H. Takano, K. Izumiya, N. Kumagai and K. Hashimoto, 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 CrossRef CAS.
- L. P. Teh, S. Triwahyono, A. A. Jalil, C. R. Mamat, S. M. Sidik, N. A. A. Fatah, R. R. Mukti and T. Shishido, Nickel-promoted mesoporous ZSM5 for carbon monoxide methanation, RSC Adv., 2015, 5, 64651–64660 RSC.
- Q. Liu, J. Gao, M. Zhang, H. Li, F. Gu, G. Xu, Z. Zhong and F. Su, Highly active and stable Ni/[gamma]-Al2O3 catalysts selectively deposited with CeO2 for CO methanation, RSC Adv., 2014, 4, 16094–16103 RSC.
- Z. Yuan and B. Chen, Process synthesis for addressing the sustainable energy systems and environmental issues, AIChE J., 2012, 58, 3370–3389 CrossRef CAS.
- D. C. D. da Silva, S. Letichevsky, L. E. P. Borges and L. G. Appel, The Ni/ZrO2 catalyst and the methanation of CO and CO2, Int. J. Hydrogen Energy, 2012, 37, 8923–8928 CrossRef CAS.
- J. Kopyscinski, T. J. Schildhauer and S. M. A. Biollaz, Production of synthetic natural gas (SNG) from coal and dry biomass – A technology review from 1950 to 2009, Fuel, 2010, 89, 1763–1783 CrossRef CAS.
- P. Panagiotopoulou, D. I. Kondarides and X. E. Verykios, Selective methanation of CO over supported Ru catalysts, Appl. Catal., B, 2009, 88, 470–478 CrossRef CAS.
- A. Boffa, C. Lin, A. T. Bell and G. A. Somorjai, Promotion of CO and CO2 Hydrogenation over Rh by Metal Oxides: The Influence of Oxide Lewis Acidity and Reducibility, J. Catal., 1994, 149, 149–158 CrossRef CAS.
- S. Y. Wang, S. H. Moon and M. A. Vannice, The effect of SMSI (strong metal-support interaction) behavior on CO adsorption and hydrogenation on Pd catalysts: II. Kinetic behavior in the methanation reaction, J. Catal., 1981, 71, 167–174 CrossRef CAS.
- J. Li, L. Zhou, Q. Zhu and H. Li, CO methanation over a macro–mesoporous Al2O3 supported Ni catalyst in a fluidized bed reactor, RSC Adv., 2015, 5, 64486–64494 RSC.
- M. A. A. Aziz, A. A. Jalil, S. Triwahyono, R. R. Mukti, Y. H. Taufiq-Yap and M. R. Sazegar, Highly active Ni-promoted mesostructured silica nanoparticles for CO2 methanation, Appl. Catal., B, 2014, 147, 359–368 CrossRef CAS.
- S. Tada, R. Kikuchi, K. Wada, K. Osada, K. Akiyama, S. Satokawa and Y. Kawashima, Long-term durability of Ni/TiO2 and Ru–Ni/TiO2 catalysts for selective CO methanation, J. Power Sources, 2014, 264, 59–66 CrossRef CAS.
- H. Lu, X. Yang, G. Gao, K. Wang, Q. Shi, J. Wang, C. Han, J. Liu, M. Tong, X. Liang and C. Li, Mesoporous zirconia-modified clays supported nickel catalysts for CO and CO2 methanation, Int. J. Hydrogen Energy, 2014, 39, 18894–18907 CrossRef CAS.
- B. Jenewein, M. Fuchs and K. Hayek, The CO methanation on Rh/CeO2 and CeO2/Rh model catalysts: a comparative study, Surf. Sci., 2003, 532–535, 364–369 CrossRef CAS.
- S. Takenaka, Complete removal of carbon monoxide in hydrogen-rich gas stream through methanation over supported metal catalysts, Int. J. Hydrogen Energy, 2004, 29, 1065–1073 CrossRef CAS.
- Y. Wang, R. Wu and Y. Zhao, Effect of ZrO2 promoter on structure and catalytic activity of the Ni/SiO2 catalyst for CO methanation in hydrogen-rich gases, Catal. Today, 2010, 158, 470–474 CrossRef CAS.
- K. Hayek, B. Jenewein, B. Klötzer and W. Reichl, Surface reactions on inverse model catalysts: CO adsorption and CO hydrogenation on vanadia- and ceria-modified surfaces of rhodium and palladium, Top. Catal., 2000, 14, 25–33 CrossRef CAS.
- Q. Liu, F. Gu, X. Lu, Y. Liu, H. Li, Z. Zhong, G. Xu and F. Su, Enhanced catalytic performances of Ni/Al2O3 catalyst via addition of V2O3 for CO methanation, Appl. Catal., A, 2014, 488, 37–47 CrossRef CAS.
- Z. Qin, J. Ren, M. Miao, Z. Li, J. Lin and K. Xie, The catalytic methanation of coke oven gas over Ni–Ce/Al2O3 catalysts prepared by microwave heating: Effect of amorphous NiO formation, Appl. Catal., B, 2015, 164, 18–30 CrossRef CAS.
- J. Zhang, Z. Xin, X. Meng and M. Tao, Synthesis, characterization and properties of anti-sintering nickel incorporated MCM-41 methanation catalysts, Fuel, 2013, 109, 693–701 CrossRef CAS.
- B. Lu and K. Kawamoto, Preparation of the highly loaded and well-dispersed NiO/SBA-15 for methanation of producer gas, Fuel, 2013, 103, 699–704 CrossRef CAS.
- K. Kamiya, T. Yoko, K. Tanaka and M. Takeuchi, Thermal evolution of gels derived from CH3Si(OC2H5)3 by the sol–gel method, J. Non-Cryst. Solids, 1990, 121, 182–187 CrossRef CAS.
- S. Kongwudthiti, P. Praserthdam, W. Tanakulrungsank and M. Inoue, The influence of Si–O–Zr bonds on the crystal-growth inhibition of zirconia prepared by the glycothermal method, J. Mater. Process. Technol., 2003, 136, 186–189 CrossRef CAS.
- J. Anderson, C. Fergusson, I. Rodriguezramos and A. Guerreroruiz, Influence of Si/Zr ratio on the formation of surface acidity in silica–zirconia aerogels, J. Catal., 2000, 192, 344–354 CrossRef CAS.
- M. Piumetti, B. Bonelli, P. Massiani, S. Dzwigaj, I. Rossetti, S. Casale, M. Armandi, C. Thomas and E. Garrone, Effect of vanadium dispersion and of support properties on the catalytic activity of V-containing silicas, Catal. Today, 2012, 179, 140–148 CrossRef CAS.
- F. Gao, Y. Zhang, H. Wan, Y. Kong, X. Wu, L. Dong, B. Li and Y. Chen, The states of vanadium species in V–SBA-15 synthesized under different pH values, Microporous Mesoporous Mater., 2008, 110, 508–516 CrossRef CAS.
- Q. Dai, X. Wang, G. Chen, Y. Zheng and G. Lu, Direct synthesis of Cerium(III)-incorporated SBA-15 mesoporous molecular sieves by two-step synthesis method, Microporous Mesoporous Mater., 2007, 100, 268–275 CrossRef CAS.
- S. C. Laha, P. Mukherjee, S. R. Sainkar and R. Kumar, Cerium Containing MCM-41-Type Mesoporous Materials and their Acidic and Redox Catalytic Properties, J. Catal., 2002, 207, 213–223 CrossRef CAS.
- S. S. Chang, B. Clair, J. Ruelle, J. Beauchene, F. Di Renzo, F. Quignard, G. J. Zhao, H. Yamamoto and J. Gril, Mesoporosity as a new parameter for understanding tension stress generation in trees, J. Exp. Bot., 2009, 60, 3023–3030 CrossRef CAS PubMed.
- X. Yang, Wendurima, G. Gao, Q. Shi, X. Wang, J. Zhang, C. Han, J. Wang, H. Lu, J. Liu and M. Tong, Impact of mesoporous structure of acid-treated clay on nickel dispersion and carbon deposition for CO methanation, Int. J. Hydrogen Energy, 2014, 39, 3231–3242 CrossRef CAS.
- X. Yang, X. Wang, G. Gao, Wendurima, E. Liu, Q. Shi, J. Zhang, C. Han, J. Wang, H. Lu, J. Liu and M. Tong, Nickel on a macro–mesoporous Al2O3@ZrO2 core/shell nanocomposite as a novel catalyst for CO methanation, Int. J. Hydrogen Energy, 2013, 38, 13926–13937 CrossRef CAS.
- G. Du, S. Lim, M. Pinault, C. Wang, F. Fang, L. Pfefferle and G. Haller, Synthesis, characterization, and catalytic performance of highly dispersed vanadium grafted SBA-15 catalyst, J. Catal., 2008, 253, 74–90 CrossRef CAS.
- X. Gao, S. R. Bare, B. M. Weckhuysen and I. E. Wachs, In situ Spectroscopic Investigation of Molecular Structures of Highly Dispersed Vanadium Oxide on Silica under Various Conditions, J. Phys. Chem. B, 1998, 102, 10842–10852 CrossRef CAS.
- P. Gruene, T. Wolfram, K. Pelzer, R. Schlögl and A. Trunschke, Role of dispersion of vanadia on SBA-15 in the oxidative dehydrogenation of propane, Catal. Today, 2010, 157, 137–142 CrossRef CAS.
- X. Gao, J. L. G. Fierro and I. E. Wachs, Structural Characteristics and Catalytic Properties of Highly Dispersed ZrO2/SiO2 and V2O5/ZrO2/SiO2 Catalysts, Langmuir, 1999, 15, 3169–3178 CrossRef CAS.
- M. Zhang, E. K. Salje, G. C. Capitani, H. Leroux, A. M. Clark, J. Schlüter and R. C. Ewing, Annealing of alpha-decay damage in zircon: a Raman spectroscopic study, J. Phys.: Condens. Matter, 2000, 12, 3131 CrossRef CAS.
- Z. Wu, M. Li, J. Howe, H. M. Meyer 3rd and S. H. Overbury, Probing defect sites on CeO2 nanocrystals with well-defined surface planes by Raman spectroscopy and O2 adsorption, Langmuir, 2010, 26, 16595–16606 CrossRef CAS PubMed.
- J. Gao, C. Jia, J. Li, F. Gu, G. Xu, Z. Zhong and F. Su, Nickel Catalysts Supported on Barium Hexaaluminate for Enhanced CO Methanation, Ind. Eng. Chem. Res., 2012, 51, 10345–10353 CrossRef CAS.
- Y. Zhang, Z. Qin, G. Wang, H. Zhu, M. Dong, S. Li, Z. Wu, Z. Li, Z. Wu, J. Zhang, T. Hu, W. Fan and J. Wang, Catalytic performance of MnOx–NiO composite oxide in lean methane combustion at low temperature, Appl. Catal., B, 2013, 129, 172–181 CrossRef CAS.
- C. Hess, G. Tzolova-Müller and R. Herbert, The Influence of Water on the Dispersion of Vanadia Supported on Silica SBA-15:
A Combined XPS and Raman Study, J. Phys. Chem. C, 2007, 111, 9471–9479 CAS. - B. P. Barbero, L. E. Cadús and L. Hilaire, XPS studies for vanadium pentoxide along the catalytic bed: oxidative dehydrogenation of propane, Appl. Catal., A, 2003, 246, 237–242 CrossRef CAS.
- A. B. Boffa, A. T. Bell and G. A. Somorjai, Vanadium Oxide Deposited on an Rh Foil: CO and CO2 Hydrogenation Reactivity, J. Catal., 1993, 139, 602–610 CrossRef CAS.
- R. Zanoni, F. Decker, C. Coluzza, F. Artuso, N. Cimino, G. Di Santo and E. Masetti, Surface evolution of Ni–V transparent oxide films upon Li insertion reactions, Surf. Interface Anal., 2002, 33, 815–824 CrossRef CAS.
- T. Mori, A. Miyamoto, N. Takahashi, M. Fukagaya, T. Hattori and Y. Murakami, Promotion effects of vanadium, niobium, molybdenum, tungsten, and rhenium oxides on surface reactions in the carbon monoxide hydrogenation over ruthenium/aluminum oxide catalyst, J. Phys. Chem., 1986, 90, 5197–5201 CrossRef CAS.
- X. Lu, F. Gu, Q. Liu, J. Gao, Y. Liu, H. Li, L. Jia, G. Xu, Z. Zhong and F. Su, VOx promoted Ni catalysts supported on the modified bentonite for CO and CO2 methanation, Fuel Process. Technol., 2015, 135, 34–46 CrossRef CAS.
- R. Razzaq, H. Zhu, L. Jiang, U. Muhammad, C. Li and S. Zhang, Catalytic Methanation of CO and CO2 in Coke Oven Gas over Ni–Co/ZrO2–CeO2, Ind. Eng. Chem. Res., 2013, 52, 2247–2256 CrossRef CAS.
- L. Qiu, F. Liu, L. Zhao, Y. Ma and J. Yao, Comparative XPS study of surface reduction for nanocrystalline and microcrystalline ceria powder, Appl. Surf. Sci., 2006, 252, 4931–4935 CrossRef CAS.
- A. E. Nelson and K. H. Schulz, Surface chemistry and microstructural analysis of CexZr1−xO2−y model catalyst surfaces, Appl. Surf. Sci., 2003, 210, 206–221 CrossRef CAS.
- S. Ardizzone and C. L. Bianchi, XPS characterization of sulphated zirconia catalysts: the role of iron, Surf. Interface Anal., 2000, 30, 77–80 CrossRef CAS.
- A. G. Sato, D. P. Volanti, D. M. Meira, S. Damyanova, E. Longo and J. M. C. Bueno, Effect of the ZrO2 phase on the structure and behavior of supported Cu catalysts for ethanol conversion, J. Catal., 2013, 307, 1–17 CrossRef CAS.
- G.-S. Wu, L.-C. Wang, Y.-M. Liu, Y. Cao, W.-L. Dai, H.-Y. He and K.-N. Fan, Implication of the role of oxygen anions and oxygen vacancies for methanol decomposition over zirconia supported copper catalysts, Appl. Surf. Sci., 2006, 253, 974–982 CrossRef CAS.
- S. Hwang, U. G. Hong, J. Lee, J. H. Baik, D. J. Koh, H. Lim and I. K. Song, Methanation of Carbon Dioxide Over Mesoporous Nickel–M–Alumina (M = Fe, Zr, Ni, Y, and Mg) Xerogel Catalysts: Effect of Second Metal, Catal. Lett., 2012, 142, 860–868 CrossRef CAS.
- J. Zhang, Z. Xin, X. Meng, Y. Lv and M. Tao, Effect of MoO3 on the heat resistant performances of nickel based MCM-41 methanation catalysts, Fuel, 2014, 116, 25–33 CrossRef CAS.
- R. Takahashi, S. Sato, T. Sodesawa, M. Yoshida and S. Tomiyama, Addition of zirconia in Ni/SiO2 catalyst for improvement of steam resistance, Appl. Catal., A, 2004, 273, 211–215 CrossRef CAS.
- M. Kong, Y. Li, X. Chen, T. Tian, P. Fang, F. Zheng and X. Zhao, Tuning the Relative Concentration Ratio of Bulk Defects to Surface Defects in TiO2 Nanocrystals Leads to High Photocatalytic Efficiency, J. Am. Chem. Soc., 2011, 133, 16414–16417 CrossRef CAS PubMed.
- H. Y. Luo, H. W. Zhou, L. W. Lin, D. B. Liang, C. Li, D. Fu and Q. Xin, Role of Vanadium Promoter in Rh–V/SiO2 Catalysts for the Synthesis of C2-Oxygenates from Syngas, J. Catal., 1994, 145, 232–234 CrossRef CAS.
- A. G. Boudjahem, S. Monteverdi, M. Mercy and M. M. Bettahar, Study of nickel catalysts supported on silica of low surface area and prepared by reduction of nickel acetate in aqueous hydrazine, J. Catal., 2004, 221, 325–334 CrossRef CAS.
- F. Benseradj, F. Sadi and M. Chater, Hydrogen spillover studies on diluted Rh/Al2O3 catalyst, Appl. Catal., A, 2002, 228, 135–144 CrossRef CAS.
- H. Y. Luo, W. Zhang, H. W. Zhou, S. Y. Huang, P. Z. Lin, Y. J. Ding and L. W. Lin, A study of Rh–Sm–V/SiO2 catalysts for the preparation of C2-oxygenates from syngas, Appl. Catal., A, 2001, 214, 161–166 CrossRef CAS.
- J. Sehested, S. Dahl, J. Jacobsen and J. R. Rostrup-Nielsen, Methanation of CO over Nickel:
Mechanism and Kinetics at High H2/CO Ratios, J. Phys. Chem. B, 2004, 109, 2432–2438 CrossRef PubMed.
|
This journal is © The Royal Society of Chemistry 2015 |
Click here to see how this site uses Cookies. View our privacy policy here.