Haidong Li,
Jun Ren*,
Xiang Qin,
Zhifeng Qin,
Jianying Lin and
Zhong Li
Key Laboratory of Coal Science and Technology (Taiyuan University of Technology), Ministry of Education and Shanxi Province, No. 79 Yingze West Street, Taiyuan 030024, China. E-mail: renjun@tyut.edu.cn; Fax: +86 0351 6018598; Tel: +86 0351 6018598
First published on 6th November 2015
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).
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.
:
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) |
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) |
414 mL mol−1) at STP; and dr represents the degree of reduction of nickel species calculated from the H2-TPR profile.
![]() | ||
| 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.
| 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) |
|---|---|---|---|---|---|---|
| a Calculated by the BET equation.b Pore volume obtained from the volume of nitrogen adsorbed at the relative pressure of 0.997.c BJH desorption average pore size.d Calculated by the XRD diffraction peak using the Scherrer equation.e Ni dispersion calculated based on H2-TPR and H2-TPD.f 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 |
![]() | ||
| 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)).
![]() | ||
| Fig. 3 The nitrogen adsorption–desorption isotherms (a) and pore size distribution curves (b) of the samples. | ||
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.
| 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+ | |
| a 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 |
| Catalysts | Ni 2p3/2 | Zr 3d5/2 | O 1s | |||
|---|---|---|---|---|---|---|
| Ni0 | Ni2+ | ZrI species | ZrII species | OI species | OII species | |
| a 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.
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.
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. 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.
000 mL g−1 h−1 was evaluated. Fig. 11 shows that the CH4 selectivity of 10Ni–5V/SBA-15 slightly decreased from 93.3 to 90.2% during 60 h of the test, whereas the CO conversion did not obviously change (from 99.8 to 99.2%). There are two possible reasons for such behavior: (1) the water gas shift reaction (CO + H2O → CO2 + H2) slightly increases levels of the byproduct CO2. (2) Carbon deposition at high temperature leads to a slight decrease in catalytic activity. 10Ni–5V/SBA-15 exhibited high activity and stability at high temperature, which because the promoter V prevents sintering of active species at high temperature.
![]() | ||
Fig. 11 60 hour stability results of the 10Ni–5V/SBA-15 catalyst (reaction conditions: P = 1 MPa, WHSV = 15 000 mL g−1 h−1, H2/CO = 3). | ||
Furthermore, it is widely accepted that the lower vanadium species can effectively enhance the CO dissociation rate.45 The obtained XPS spectra indicate that during CO methanation, the change in oxidation state of V gave rise to an electronic effect. It subsequently enhanced the density of the electron cloud of active Ni0 species through the oxidation–reduction shift cycle (V3+ ↔ V4+ ↔ V5+), which promotes CO dissociation and improves CH4 selectivity. The shift cycle is due to the oxidation characteristics of V species and the reduction properties of the reactant gas. Thus, the results demonstrate that the catalyst 10Ni–5V/SBA-15 is a promising candidate for use in methanation.
![]() | ||
| 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.
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.
A Combined XPS and Raman Study, J. Phys. Chem. C, 2007, 111, 9471–9479 CAS.
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 |