301
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Effect of SiO2 Addition to Highly Loaded Ni on CeO2 for CO2 Methanation

, &
Article: 2303309 | Received 20 Oct 2023, Accepted 04 Jan 2024, Published online: 21 Feb 2024

Abstract

In this study, 60 wt% Ni on xSiO2–(1 – x)CeO2 catalysts were prepared by flame spray pyrolysis, and the effects of SiO2 fraction (x = 0–0.2) on the material properties and catalytic activity of CO2 methanation were investigated. After reducing the catalysts in 5%H2−Ar at 500 °C, the catalytic activity was evaluated; the highest catalytic activity was exhibited at x = 0.05. Furthermore, SiO2 addition was found to improve the catalytic activity after 110 hours of the reaction. Powder X-ray diffraction revealed that SiO2 addition did not alter the crystallinity of the catalysts before and after the reduction. Conversely, the appropriate amount of SiO2 doping increased the Ni surface area and pore volume of the reduced catalysts, as confirmed by H2 chemisorption and N2 adsorption–desorption measurements. H2-TPR measurements showed an increasing trend in the reduction temperature of particulate NiO as the SiO2 fraction increased. The shift of the reduction temperatures indicates the better contact between NiO and SiO2, which hindered the sintering of Ni particles during the reduction at 500 °C. Therefore, an appropriate amount of SiO2 addition prevented the growth of Ni particles, resulting in the increase of the Ni surface area. The increase of the Ni surface area contributed to the improvement of the catalytic activity.

1. Introduction

Hydrogen has garnered increased interest as an energy carrier (Graetz Citation2009) because it can be produced using renewable energy (e.g., biomass and solar energy). However, the inherent flammability of H2 presents a formidable challenge (Dadashzadeh et al. Citation2018). The reaction of hydrogen with CO2 to form methane (Centi et al. Citation2013) is promising because CH4 is less flammable than H2 (Dadashzadeh et al. Citation2018). Furthermore, the utility of methane is as a fuel in daily activities, and thermal electric power plants are versatile. Consequently, the methane produced from the reaction of hydrogen with CO2 can be integrated into existing infrastructure.

Numerous catalysts have been proposed for this reaction, with a particular focus on cost-effective Ni-based catalysts, such as Ni/CeO2 (Tada et al. Citation2012; Fukuhara et al. Citation2017), Ni/CeO2–Al2O3 (Wang and Lu Citation1998), Ni/ZrO2 (Jia et al. Citation2019), and Ni/Y2O3 (Muroyama et al. Citation2016). Owing to its high reducibility (Tada et al. Citation2012) and ability to absorb CO2 (Fukuhara et al. Citation2017), CeO2 is a promising metal-oxide support. The aforementioned studies suggest that the active sites for CO2 methanation are at/near the Ni–CeO2 interfaces. One straightforward strategy for increasing the number of active sites is increasing the Ni loading while keeping the Ni size small, thereby effectively increasing the Ni–CeO2 interfaces. However, Ni loading has an upper limit because excessive loading induces the formation of huge Ni particles (Zyryanova et al. Citation2014; Beierlein et al. Citation2019). Some studies have suggested that the optimal Ni loading for CO2 methanation is approximately 15–30 wt% (Liu et al. Citation2013; Nie et al. Citation2017). Recently, highly loaded Ni catalysts (50–92.5 wt% of Ni content) such as Ni–Al hydrotalcite (He et al. Citation2014), Ni–La2O3 (Tang et al. Citation2019), Ni–CeO2 (Fujiwara et al. Citation2021), and a sponge nickel (Tada et al. Citation2017), were demonstrated as active catalysts for CO2 methanation. Despite their promising activity, their high Ni contents could induce the sintering of Ni particles, thereby degrading the activity for long-term use. To prevent the sintering of Ni particles in catalysts, the addition of SiO2 to the catalysts has been examined frequently (e.g., Ni/ZrO2–SiO2; Charisiou et al. Citation2019). Therefore, SiO2 addition to highly loaded Ni catalysts is expected to improve long-term performance.

Flame spray pyrolysis (FSP) was used to synthesize highly loaded NiO particles (60 wt% as Ni) on SiO2–CeO2. FSP facilitates the one-step production of multicomponent metal oxides (Teoh et al. Citation2010; Li et al. Citation2016) and scaling up of the production rate; kg h−1 (Wegner et al. Citation2011). We previously demonstrated that FSP can deposit small particles at high metal loading on metal-oxide supports; for example, 50 wt% Ag on SiO2 (Fujiwara et al. Citation2015) and 60 wt% Cu on ZrO2 (Fujiwara et al. Citation2019; Tada et al. Citation2020). Moreover, particles containing SiO2 (e.g. ZnO (Tani et al. Citation2002), CeZrO2 (Schulz et al. Citation2003), WO3 (Righettoni et al. Citation2010) and MoO3 (Güntner et al. Citation2016)) prepared by FSP exhibited high thermal stability. Thus, FSP allows doping SiO2 to various metal oxides. However, the effect of SiO2 doping to highly loaded metal particles and its catalytic activity have not been examined. It is expected that overloading SiO2 to Ni/CeO2 degrades the catalytic activity for CO2 methanation because Ni/SiO2 is less active than Ni/CeO2 (Liu et al. Citation2022). Therefore, in this study, the effects of the SiO2 fraction in FSP-made 60 wt%Ni/CeO2 were investigated and the optimal SiO2 content for improving the catalytic activity and thermal stability was explored.

2. Experimental

2.1. Preparation of Ni/CeO2-SiO2 by Flame Spray Pyrolysis

As shown in , 60 wt% Ni supported on xSiO2–(1 – x)CeO2 catalysts, namely Si–x, where x is the molar fraction of SiO2 in the support, were prepared using an FSP reactor. The precursor solutions were prepared by dissolving nickel acetate tetrahydrate (Fujifilm Wako Pure Chemical, Osaka, Japan, purity >98.0%), cerium 2-ethylhexanoate in 2-ethylhexanoic acid (Alfa Aesar, Ward Hill, MA), and hexamethyldisiloxane (HMDSO) into a 1:1 mixture of methanol (Sigma-Aldrich, St. Louis, MO, Reagent Grade) and 2-ethylhexanoic acid (Sigma-Aldrich, St. Louis, MO, purity >99%). The SiO2 fraction was controlled by the amount of HMDSO dissolved in the solution, and the total metal concentration (Ni + Ce + Si) was set to 0.2 mol L−1.

Figure 1. Formation of Ni/xSiO2–(1 – x)CeO2 catalysts using an flame spray pyrolysis (FSP) reactor.

Figure 1. Formation of Ni/xSiO2–(1 – x)CeO2 catalysts using an flame spray pyrolysis (FSP) reactor.

The prepared solution was fed to the FSP reactor at 3 mL min−1 and dispersed to fine spray using 5 LSTP min−1 of O2 dispersant (technical grade). The spray was combusted using a plot flame (CH4:O2 = 1.5 LSTP min−1:3.2 LSTP min−1) to generate nanoparticles. The particles were collected using a vacuum pump (Busch Seco SV1040C, Maulburg, Germany) and filtered using a Φ257 mm glass fiber filter (Albet LabScience, GF6, Dassel, Germany) located 65 cm above the FSP reactor. If needed, the as-prepared catalysts were reduced at 500 °C in 5%H2–Ar (100 mL min−1) for 1 h. Subsequently, the reduced powder was cooled to room temperature in Ar (100 mL min−1). The reduction condition was the same as that before the catalytic activity test.

2.2. Evaluation of Catalytic Performance for CO2 Methanation

The catalytic performance was evaluated using a fixed-bed reactor. The catalyst (100 mg) and SiC powder (Fujifilm Wako Pure Chemical, Osaka, Japan, 400 mg) were thoroughly mixed in a mortar and poured into a quartz tube (inner/outer diameters of Φ6/Φ8 mm). The catalyst bed (sample + SiC) was 13–14 mm long for all the catalysts. The filled catalyst was reduced at 500 °C for 1 h under 5% H2–Ar (100 mLSTP min−1) and then cooled to room temperature (c.a. 25 °C). Afterward, a reactant gas (CO2/H2/N2 = 1/4/1) was fed to the catalyst at 60 mLSTP min−1. The temperature of the catalyst bed was measured using a K-type thermocouple (Φ1 mm) placed at the center of the catalyst bed. The temperature-ramping rates below and above 250 °C were 1 and 2 °C min−1, respectively, and the temperature was maintained for at least 30 min before the products were analyzed. The composition of the product gas was measured using an online gas chromatograph (Shimadzu, GC-2014, Kyoto, Japan) equipped with a thermal conductivity detector. CO was not detected by the detector using He as a career gas in all conditions examined in this study. The conversion and reaction rate of CO2 were evaluated using the following equations: (1) CO2conversion (%)=(FCO2_inFCO2_out)/FCO2_in×100(1) (2) Reaction rate,rCO2(mol s1gcat1)=(FCO2_inFCO2_out)/Mcat(2) where FCO2_in and FCO2_out were the flow rates of CO2 (mol s−1) at the inlet and outlet, respectively. Mcat was the mass of the catalyst in the reactor. The Arrhenius plot was obtained by plotting the natural log of the reaction rate against the inverse temperature. To obtain the reaction rate for the plots, the catalytic activity was evaluated using 50 mg of the catalysts (with 400 mg of SiC). From the Arrhenius plots, the apparent activation energy (Ea) was obtained using the Arrhenius equation (EquationEq. (3)). (3) k=Aexp(EaRT)(3) where k, A, R, and T are the kinetic constant, frequency factor, gas constant (8.314 J mol−1 K−1), and temperature of the catalyst bed, respectively. In this study, we assumed that the kinetic constant is proportional to the reaction rate.

2.3. Particle Characterization

The powder X-ray diffraction (PXRD) patterns of catalysts were obtained using a diffractometer (Rigaku Ultima IV, Tokyo, Japan, Cu Kα, 40 kV, 40 mA). The crystallite sizes of CeO2 (1 1 1), NiO (2 0 0), and Ni (1 1 1) were calculated from the peaks at 28°, 43°, and 44°, respectively, using Scherrer’s equation (EquationEq. (4)). (4) Crystallite size, dXRD (nm)=λβ cosθ(4) where K (=0.89) is the shape factor, λ (=0.154 nm) the X-ray wavelength, β the line broadening at half the maximum intensity in radians, and θ the Bragg angle.

The particle morphology was investigated by using a high angle annular dark field scanning transmission electron microscope (HAADF-STEM) equipped with a spherical aberration corrector for a STEM probe (JEOL JEM-ARM200F, Tokyo, Japan). The sample was dispersed in ethanol (Wako, Osaka, Japan, purity >99.5%), and the suspension was dropped onto a carbon-coated Cu grid (Ohken Shoji Co., Tokyo, Japan, NP-C15) to deposit the particles on the microgrid. The distributions of Ni, Ce and Si species in the particles were recorded using an energy dispersive X-ray (EDX) detector.

N2 adsorption for the catalysts was performed using a BELSORP-Mini II (MicrotracBEL Corp., Osaka, Japan). Before the measurements, the samples were degassed at 150 °C under vacuum at <1 kPa for 1 h. The specific surface area (SSA) was determined from the amount of adsorbed N2 on the particle surface at 77 K using the Brunauer–Emmett–Teller method.

The reducibility of the samples was evaluated by temperature-programmed reduction by H2 (H2-TPR) using BELCAT II (MicrotracBEL Corp., Osaka, Japan). The sample of c.a. 25 mg was filled into a quartz tube. Before the measurement, the sample was heated at 300 °C for 1 h and cooled down to 45 °C in an Ar flow. The sample was heated from 45 °C to 700 °C with 5 °C min−1 of ramping rate in a 5% H2–Ar flow (100 mL min−1). The product gas was introduced to a molecular sieve trap (Shinwa Chemical Industries Ltd., 3A, Kyoto, Japan) and, subsequently, to the thermal conductivity detector. To calculate the H2 consumption from the sample, a commercial CuO powder (Fujifilm Wako Pure Chemical, Osaka, Japan, purity >99.9%) was used as the reference material.

The surface area of Ni in the catalysts was investigated by H2 pulse titration using BELCAT II (MicrotracBEL Corp., Osaka, Japan). About 50 mg of the sample was placed in a quartz tubular reactor and reduced by 5% H2–Ar at 500 °C for 1 h. Then, the sample was purged with Ar at 500 °C for 15 min to remove all adsorbed H2 on the surface and cooled to 40 °C. After stabilizing the temperature at 40 °C, 0.966 mLSTP of 10% H2–Ar was injected into 50 mL min−1 of Ar carrier every 2 min until the equilibrium of H2 adsorption on the catalyst surface was reached. The surface area of metallic Ni was calculated using EquationEq. (5), assuming that every Ni atom on the catalyst surface adsorbs one hydrogen atom. (5) SSA (mNi2 gcat1)=2 × MH2×NavANi(5) where MH2, Nav, and ANi are the amount of H2 molecules adsorbed on the catalyst (mol gcat1), Avogadro’s number (6.02 × 1023 [atom mol−1]), and the number of Ni surface atoms per unit area (15.4 × 1018 [atom m−2]).

3. Results and Discussion

shows the CO2 methanation activity of 60 wt% Ni on xSiO2–(1 – x)CeO2 catalysts (Si–x). For all the catalysts, the conversion of CO2 was initiated at 175–200 °C and reached the equilibrium value above 300–350 °C. The concentration of CO in the product was below the detectable limit (<0.01%), implying selective CH4 formation. When the SiO2 fraction increased to x = 0.05, CO2 conversion increased. Further increase in the SiO2 fraction degraded the catalytic activity. Therefore, the optimal SiO2 addition is beneficial for catalytic performance. The reaction rate of the optimal catalyst (Si–0.05) was 8.6 × 10−2 mol h−1 gcat−1 at 225 °C. Notably, the reaction rate of the optimal catalyst was higher than those of the Ru catalysts (e.g., 3 wt% Ru/CeO2; 8.1 × 10−3 mol h−1 gcat−1 at 225 °C (Wang et al. Citation2015) and 4 wt% Ru/TiO2; 5.6 × 10−2 mol h−1 gcat−1 at 240 °C (Zhou et al. Citation2022). Therefore, FSP-made catalysts performed excellently.

Figure 2. (a) CO2 methanation activity of Si–x. (b) Catalytic activity of Si–0 and Si–0.05 for 110 h.

Figure 2. (a) CO2 methanation activity of Si–x. (b) Catalytic activity of Si–0 and Si–0.05 for 110 h.

Despite the high catalytic activity of the FSP-made catalysts, high Ni loading (60 wt%) could suffer from the sintering of Ni particles during the reaction. SiO2 doping is expected to improve thermal stability by blocking the contact of Ni particles. Herein, the effect of SiO2 on catalytic activity for 110 h was investigated. shows the CO2 conversion of Si–0 and Si–0.05 within 110 h. During the test, the temperature initially increased to 250 °C and subsequently to 300 °C. The initial CO2 conversions at 250 °C by the catalysts with x = 0 and 0.05 were 61 and 70%, respectively. The conversion at 300 °C by both catalysts did not change (88–91%) for 100 h. Afterward, the temperature was reduced to 250 °C and then maintained for 10 h. The conversions after temperature reduction were 55% and 64% for x = 0 and 0.05, respectively. Therefore, the catalytic activity slightly degraded after the reaction at 300 °C for 100 h, regardless of SiO2 addition. Nevertheless, the conversion of the SiO2-doped catalyst (x = 0.05) after the long-time reaction (at 110 h) was higher than the initial conversion of the non-doped catalyst (x = 0), indicating that the optimal SiO2 addition improved the conversion after the long-time reaction.

As shown in , the optimal amount of SiO2 addition improved the catalytic activity. To elucidate the effect of SiO2 on the FSP-made Ni/CeO2, the material properties were investigated. First, the crystallinity of the catalysts was analyzed by PXRD. shows the PXRD patterns of Si–x. Despite the SiO2 fraction, the peaks of bunsenite NiO (37° and 43°) and fluorite CeO2 (28°, 33°, and 47°) were observed, which is consistent with other FSP-made NiO (Azurdia et al. Citation2008) and CeO2 (Mädler et al. Citation2002). Si-related compounds were not detected by PXRD as they were present as amorphous SiO2.

Figure 3. PXRD patterns of Si–x (a) before and (b) after reduction in 5%H2–Ar at 500 °C for 1 h.

Figure 3. PXRD patterns of Si–x (a) before and (b) after reduction in 5%H2–Ar at 500 °C for 1 h.

After reducing Si–x in 5%H2−Ar at 500 °C for 1 h, the peaks of metallic Ni (44° and 52°) and fluorite CeO2 (28°, 33°, and 47°) were observed, as shown in . The peaks of NiO disappeared owing to the reduction of NiO into metallic Ni. In addition, the reduction increased the intensity of the CeO2 peaks, indicating the particle growth of CeO2. The increase in the peak intensity of Si–0.2 was lower than that of the others because the presence of SiO2 prevents the growth of CeO2 (Reddy et al. Citation2005).

To quantitatively evaluate the particle growth, the crystallite sizes of Si–x were obtained. shows that the crystallite sizes of CeO2 and NiO in Si–x before the reduction were 3–4 nm and 4–5 nm, respectively. After the reduction, the crystallite size of CeO2 slightly increased to 4–5 nm, but the size did not vary according to the SiO2 fraction. Similarly, SiO2 addition did not alter the crystallite size of Ni. Therefore, SiO2 doping did not affect the crystallinity of the FSP-made Ni/CeO2.

Table 1. Crystallite sizes of Si–x before and after reduction in 5%H2–Ar at 500 °C for 1 h.

During the PXRD measurements, the reduced catalysts were exposed to the atmosphere. Consequently, the surface of Ni particles was subjected to oxidation, resulting in a small crystallite size compared to the actual Ni size. To assess the Ni size while mitigating the effects of surface oxidation, the Ni surface area was obtained by H2 pulse titration. exhibits the Ni surface area of Si–x reduced in 5%H2–Ar at 500 °C for different durations (1–15 h). As the SiO2 fraction increased from 0 to 0.05, the Ni surface area of Si–x reduced for 1 h increased from 16 to 20 mNi2 gcat−1. As x increased further, the surface area decreased. As a result, the optimal SiO2 fraction for maximum Ni surface area was x = 0.05. Interestingly, at the optimal x, the catalytic activity reached the highest among the tested catalysts. The Ni surface area at x = 0.1 was comparable with that at x = 0, although the CO2 conversion of the former was lower than that of the latter. This is due to the increase of inactive Ni/SiO2 and the decrease of active Ni/CeO2 (Liu et al. Citation2022). Therefore, the increase of the Ni surface area by the appropriate SiO2 addition promoted the catalytic activity.

Figure 4. Ni surface area of Si–x reduced in 5%H2–Ar at 500 °C for different durations (1–15 h).

Figure 4. Ni surface area of Si–x reduced in 5%H2–Ar at 500 °C for different durations (1–15 h).

The effect of the SiO2 addition on the sintering of Ni particles was investigated by monitoring the evolution of the Ni surface area at the different reduction times (). Over time, the surface area of Si–0 gradually decreased; the surface areas were 16, 15, and 14 mNi2 gcat−1 at 1, 5, and 15 h of the reduction, respectively. Conversely, the surface area of Si–0.05 remained comparable (20, 19, and 19 mNi2 gcat−1 at 1, 5, and 15 h). Similarly, the surface area of Si–0.1 did not change as reduction time increased. Therefore, the SiO2 doping to FSP-made 60 wt% Ni/CeO2 prevented the sintering of Ni particles.

The SiO2 doping of the catalysts can hinder the sintering of Ni particles. Therefore, the porosities of Si–x are expected to change in the presence of SiO2. shows the N2 adsorption–desorption isotherms of Si–x reduced in 5%H2-Ar at 500 °C for 1 h. For all the SiO2 fractions, the isotherms were classified as type IV with a hysteresis above p/po = 0.85, indicating the presence of a porous structure. From the isotherms, the pore size distributions of the catalysts were obtained. shows that all the catalysts consisted of mesopores and macropores (10–100 nm). The pore distribution of x = 0.05 was broader than those of the others (x = 0 and 0.1). The broad distribution of the former was attributed to its large pore volume compared to that of the latter. FSP-made powders, including NiO and CeO2 (Fujiwara et al. Citation2021), typically have large pore volumes because of their fractal structure (Pratsinis Citation1998). Therefore, the large pore volume of Si–0.05 contributed to its excellent catalytic activity.

Figure 5. (a) N2 adsorption–desorption isotherms and (b) pore size distribution of Si–0, Si–0.05, and Si–0.1 reduced in 5%H2–Ar at 500 °C for 1 h.

Figure 5. (a) N2 adsorption–desorption isotherms and (b) pore size distribution of Si–0, Si–0.05, and Si–0.1 reduced in 5%H2–Ar at 500 °C for 1 h.

As proven by the H2 adsorption measurements () and N2 adsorption (), an appropriate amount of SiO2 doping prevents the Ni sintering. It is expected that SiO2 particles were present between Ni and/or CeO2 particles (Tani et al. Citation2002). Such SiO2 particles hindered the sintering of Ni particles resulting in an increase in the Ni surface area (Charisiou et al. Citation2019). The reduction in the Ni surface area at x ≥ 0.1 may be due to excess SiO2 doping. Overloaded SiO2 particles were deposited on the Ni particles covering the Ni surface (Schulz et al. Citation2003). HAADF-STEM and EDX mapping measurements were performed to confirm the location of SiO2 in the catalysts. According to HAADF-STEM images in , the primary particle sizes of the reduced Si–0.05 and Si–0.2 were about 10–20 nm with high uniformity. For both catalysts, EDX mappings of Ni species () exhibited 10–20 nm of special spots. In contrast, Si (c, g) and Ce (d, h) species were uniformly dispersed through the particles indicating that the sizes of CeO2 and SiO2 particles were smaller than those of Ni particles. In addition, the distributions of Si and Ce species were close to Ni particles. Therefore, CeO2 and SiO2 particles were located around Ni particles. At a high SiO2 fraction (x = 0.2), the location of SiO2 and CeO2 was further investigated as illustrated in . Because of the characteristic lattice fringes of CeO2 (0.32 nm), CeO2 particles were distinguished. The size of CeO2 was below 5 nm, which was in agreement with the crystallite size. Also, amorphous compounds (dotted circles) that correspond to SiO2 were found. Both CeO2 and SiO2 were present on relatively large particles (∼10 nm), which were Ni particles as confirmed by the EDX mappings (). Some of the SiO2 particles on Ni particles hindered the contact between adjacent Ni particles, preventing the sintering of Ni particles. However, it is difficult to quantify the amount of SiO2 located between Ni particles.

Figure 6. (a, e) HAADF-STEM and (b–d, f–h) EDX mapping images of Ni, Si and Ce species of Si–0.05 and Si–0.2 reduced under 5% H2–Ar at 500 °C for 1 h.

Figure 6. (a, e) HAADF-STEM and (b–d, f–h) EDX mapping images of Ni, Si and Ce species of Si–0.05 and Si–0.2 reduced under 5% H2–Ar at 500 °C for 1 h.

Figure 7. High-resolution HAADF-STEM image of Si–0.2 reduced under 5% H2–Ar at 500 °C for 1 h.

Figure 7. High-resolution HAADF-STEM image of Si–0.2 reduced under 5% H2–Ar at 500 °C for 1 h.

In this study, the activity test and material properties were investigated after reducing catalysts in 5%H2–Ar at 500 °C. The reducibility of Si–x was explored to clarify the state of the Ni species in the catalysts after the reduction. shows the H2-TPR profiles of Si–x catalysts. At x ≤ 0.2, two small peaks at approximately 180 and 240 °C were observed. In contrast, these peaks were not observed in Si–1 (60 wt% Ni/SiO2). Large peaks above 250 °C were observed for all the catalysts. The peak locations shifted to higher temperatures at higher SiO2 fractions. The peaks at relatively low temperatures (<250 °C) corresponded to the reduction of Ni–O–Ce sites at the Ni–CeO2 interface and/or that of Ni2+ substituted into the CeO2 lattice (Shan et al. Citation2003; Tada et al. Citation2021). Notably, the low-temperature peaks were not observed in Si–1, which did not contain CeO2. The peaks above 250 °C corresponded to the reduction of NiO particles to metallic Ni, and the peak locations changed based on the interaction between NiO and the support (Shan et al. Citation2003). According to literature (He et al. Citation2009), interactions between NiO and SiO2 particles form nickel silicate, for which the reduction temperature is higher than that of the particulate NiO. Therefore, the shift of the peaks (at 300–700 °C) to higher temperatures indicates better contact between NiO and SiO2. This interaction prevents the sintering of Ni particles during the reduction at 500 °C, resulting in the increase of the Ni surface area, as shown in .

Figure 8. H2-TPR profiles of Si–x catalysts. As a reference, the H2-TPR profile of Si–1 (60 wt% Ni/SiO2) was obtained.

Figure 8. H2-TPR profiles of Si–x catalysts. As a reference, the H2-TPR profile of Si–1 (60 wt% Ni/SiO2) was obtained.

Using the profiles, the H2 consumptions by Si–x were calculated, as shown in . For all the catalysts, the amounts of H2 consumed by Si–x were comparable (8.5–8.8 mmol gcat−1). As discussed in , the H2 consumption predominantly originated from the reduction of nickel oxides. As confirmed by PXRD (), the major nickel species in Si–x was NiO. Therefore, assuming that all Ni species in Si–x were NiO, Si–x contains 8.8 mmol gcat−1 of NiO. The NiO content is consistent with the amounts of H2 consumption (8.5–8.8 mmol gcat−1). Therefore, NiO in the catalysts was completely reduced to metallic Ni by the reduction at 500 °C because, at x ≤ 0.2, all reduction peaks appeared below 500 °C ().

Table 2. H2 consumption of Si–x calculated using the peak area of the H2-TPR profiles.

shows the Arrhenius plots of Si–0 and Si–0.05 for the reaction rate of CO2. The Arrhenius plots were obtained using the reaction conditions (200–230 °C), where the CO2 conversion was 4–24%. The reaction rate of Si–0.05 was higher than that of Si–0. Conversely, for both Si–0 and Si–0.05, the slopes of the plots were similar. The apparent activation energy values obtained by the slopes were 100 and 105 kJ mol−1 for Si–0 and Si–0.05, respectively. The activation energies were consistent with the values obtained by Ni/CeO2 catalysts (95–112 kJ mol−1) in the literature (Bian et al. Citation2020; Lin et al. Citation2021), indicating that SiO2 addition did not modify the reactivity of the active sites in the catalysts. According to literature (Liu et al. Citation2022), Ni/SiO2 shows lower activity than Ni/CeO2 for CO2 methanation. Thus, the Ni/SiO2 compounds did not work as an active sites for the reaction, even though the optimal SiO2 doping (x = 0.05) led to an increase in the Ni surface area. In the case of excess SiO2 doping (x > 0.05), the Ni surface area decreased, and the fraction of inactive Ni/SiO2 increased. Therefore, there is an optimal SiO2 fraction that can be determined by maximizing the Ni surface area while minimizing the loss of active Ni/CeO2 compounds.

Figure 9. Arrhenius plots of Si–0 and Si–0.05 for the reaction rate of CO2.

Figure 9. Arrhenius plots of Si–0 and Si–0.05 for the reaction rate of CO2.

4. Conclusions

In this study, 60 wt% Ni on xSiO2–(1 – x)CeO2 catalysts were prepared by FSP, and the effects of SiO2 fraction (x = 0–0.2) on the material properties and CO2 methanation activity were investigated. The SiO2 addition did not alter the crystallinity of Ni/CeO2 catalysts. Conversely, the appropriate amount of SiO2 doping (x ≤ 0.05) increased the Ni surface area of the catalysts reduced in 5%H2−Ar at 500 °C because SiO2 particles located between Ni particles prevented the sintering of Ni particles as confirmed by HAADF-STEM and EDX mapping. The resistance of Ni sintering by SiO2 was attributed to the increase of the pore volume at x = 0.05. However, excess SiO2 addition (x ≥ 0.1) decreased both the Ni surface area and pore volume of the catalysts. At the optimal SiO2 fraction (x = 0.05), where the Ni surface area was maximal, the highest catalytic activity for CO2 methanation was exhibited. Furthermore, SiO2 addition improved the long-term catalytic activity. Therefore, the optimal amount of SiO2 addition to 60 wt% Ni/CeO2 catalyst through FSP enhances the thermal stability and catalytic activity.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

This study was supported by the tryout program (JPMJTM20BJ) of the Japan Science and Technology Agency and the research grant by the Information Center of Particle Technology, Japan. Electron microscopy measurements were supported by “ARIM Project of the Ministry of Education, Culture, Sports, Science and Technology, Japan (MEXT), Grant No. JPMXP1223OS0036” at the Research Center for Ultra-High Voltage Electron Microscopy (Nanotechnology Open Facilities) in Osaka University.

References

  • Azurdia JA, McCrum A, Laine RM. 2008. Systematic synthesis of mixed-metal oxides in NiO–Co3O4, NiO–MoO3, and NiO–CuO systems via liquid-deed flame spray pyrolysis. J Mater Chem. 18:3249–3258. doi: 10.1039/b801745j.
  • Beierlein D, Häussermann D, Pfeifer M, Schwarz T, Stöwe K, Traa Y, Klemm E. 2019. Is the CO2 methanation on highly loaded Ni-Al2O3 catalysts really structure-sensitive? Appl Catal B Environ. 247:200–219. doi: 10.1016/j.apcatb.2018.12.064.
  • Bian Z, Chan YM, Yu Y, Kawi S. 2020. Morphology dependence of catalytic properties of Ni/CeO2 for CO2 methanation: a kinetic and mechanism study. Catal Today. 347:31–38. doi: 10.1016/j.cattod.2018.04.067.
  • Centi G, Quadrelli EA, Perathoner S. 2013. Catalysis for CO2 conversion: a key technology for rapid introduction of renewable energy in the value chain of chemical industries. Energy Environ Sci. 6:1711–1731. doi: 10.1039/c3ee00056g.
  • Charisiou ND, Papageridis KN, Siakavelas G, Sebastian V, Hinder SJ, Baker MA, Polychronopoulou K, Goula MA. 2019. The influence of SiO2 doping on the Ni/ZrO2 supported catalyst for hydrogen production through the glycerol steam reforming reaction. Catal Today. 319:206–219. doi: 10.1016/j.cattod.2018.04.052.
  • Dadashzadeh M, Kashkarov S, Makarov D, Molkov V. 2018. Risk assessment methodology for onboard hydrogen storage. Int J Hydrogen Energy. 43:6462–6475. doi: 10.1016/j.ijhydene.2018.01.195.
  • Fujiwara K, Kayano S, Nishijima M, Kobayashi K, Nanba T, Tsujimura T. 2021. Porous NiO prepared by flame spray pyrolysis for 80 wt% Ni–CeO2 catalyst and its activity for CO2 methanation. J Jpn Petrol Inst. 64:261–270. doi: 10.1627/jpi.64.261.
  • Fujiwara K, Sotiriou GA, Pratsinis SE. 2015. Enhanced Ag+ ion release from aqueous nanosilver suspensions by absorption of ambient CO2. Langmuir. 31:5284–5290. doi: 10.1021/la504946g.
  • Fujiwara K, Tada S, Honma T, Sasaki H, Nishijima M, Kikuchi R. 2019. Influences of particle size and crystallinity of highly loaded CuO/ZrO2 on CO2 hydrogenation to methanol. AIChE J. 65:e16717.
  • Fukuhara C, Hayakawa K, Suzuki Y, Kawasaki W, Watanabe R. 2017. A novel nickel-based structured catalyst for CO2 methanation: a honeycomb-type Ni/CeO2 catalyst to transform greenhouse gas into useful resources. Appl Catal A Gen. 532:12–18. doi: 10.1016/j.apcata.2016.11.036.
  • Graetz J. 2009. New approaches to hydrogen storage. Chem Soc Rev. 38:73–82. doi: 10.1039/b718842k.
  • Güntner AT, Righettoni M, Pratsinis SE. 2016. Selective sensing of NH3 by Si-doped α-MoO3 for breath analysis. Sens Actuators B Chem. 223:266–273. doi: 10.1016/j.snb.2015.09.094.
  • He L, Lin Q, Liu Y, Huang Y. 2014. Unique catalysis of Ni–Al hydrotalcite derived catalyst in CO2 methanation: cooperative effect between Ni nanoparticles and a basic support. J Energy Chem. 23:587–592. doi: 10.1016/S2095-4956(14)60144-3.
  • He S, Jing Q, Yu W, Mo L, Lou H, Zheng X. 2009. Combination of CO2 reforming and partial oxidation of methane to produce syngas over Ni/SiO2 prepared with nickel citrate precursor. Catal Today. 148:130–133. doi: 10.1016/j.cattod.2009.03.009.
  • Jia X, Zhang X, Rui N, Hu X, Liu C-j. 2019. Structural effect of Ni/ZrO2 catalyst on CO2 methanation with enhanced activity. Appl Catal B Environ. 244:159–169. doi: 10.1016/j.apcatb.2018.11.024.
  • Li S, Ren Y, Biswas P, Stephen DT. 2016. Flame aerosol synthesis of nanostructured materials and functional devices: processing, modeling, and diagnostics. Prog Energy Combust Sci. 55:1–59. doi: 10.1016/j.pecs.2016.04.002.
  • Lin S, Hao Z, Shen J, Chang X, Huang S, Li M, Ma X. 2021. Enhancing the CO2 methanation activity of Ni/CeO2 via activation treatment-determined metal-support interaction. J Energy Chem. 59:334–342. doi: 10.1016/j.jechem.2020.11.011.
  • Liu J, Li C, Wang F, He S, Chen H, Zhao Y, Wei M, Evans DG, Duan X. 2013. Enhanced low-temperature activity of CO2 methanation over highly-dispersed Ni/TiO2 catalyst. Catal Sci Technol. 3:2627–2633. doi: 10.1039/c3cy00355h.
  • Liu J, Wu X, Chen Y, Zhang Y, Zhang T, Ai H, Liu Q. 2022. Why Ni/CeO2 is more active than Ni/SiO2 for CO2 methanation? Identifying effect of Ni particle size and oxygen vacancy. Int J Hydrogen Energy. 47:6089–6096. doi: 10.1016/j.ijhydene.2021.11.214.
  • Mädler L, Stark W, Pratsinis S. 2002. Flame-made ceria nanoparticles. J Mater Res. 17:1356–1362. doi: 10.1557/JMR.2002.0202.
  • Muroyama H, Tsuda Y, Asakoshi T, Masitah H, Okanishi T, Matsui T, Eguchi K. 2016. Carbon dioxide methanation over Ni catalysts supported on various metal oxides. J Catal. 343:178–184. doi: 10.1016/j.jcat.2016.07.018.
  • Nie W, Zou X, Chen C, Wang X, Ding W, Lu X. 2017. Methanation of carbon dioxide over Ni–Ce–Zr oxides prepared by one-pot hydrolysis of metal nitrates with ammonium carbonate. Catalysts. 7:104. doi: 10.3390/catal7040104.
  • Pratsinis SE. 1998. Flame aerosol synthesis of ceramic powders. Prog Energy Combust Sci. 24:197–219. doi: 10.1016/S0360-1285(97)00028-2.
  • Reddy BM, Khan A, Lakshmanan P, Aouine M, Loridant S, Volta J-C. 2005. Structural characterization of nanosized CeO2−SiO2, CeO2−TiO2, and CeO2−ZrO2 catalysts by XRD, Raman, and HREM techniques. J Phys Chem B. 109:3355–3363. doi: 10.1021/jp045193h.
  • Righettoni M, Tricoli A, Pratsinis SE. 2010. Thermally stable, silica-doped ε-WO3 for sensing of acetone in the human breath. Chem Mater. 22:3152–3157. doi: 10.1021/cm1001576.
  • Schulz H, Stark WJ, Maciejewski M, Pratsinis SE, Baiker A. 2003. Flame-made nanocrystalline ceria/zirconia doped with alumina or silica: structural properties and enhanced oxygen exchange capacity. J Mater Chem. 13:2979–2984. doi: 10.1039/b307754c.
  • Shan W, Luo M, Ying P, Shen W, Li C. 2003. Reduction property and catalytic activity of Ce1−XNiXO2 mixed oxide catalysts for CH4 oxidation. Appl Catal A Gen. 246:1–9. doi: 10.1016/S0926-860X(02)00659-2.
  • Tada S, Fujiwara K, Yamamura T, Nishijima M, Uchida S, Kikuchi R. 2020. Flame spray pyrolysis makes highly loaded Cu nanoparticles on ZrO2 for CO2-to-methanol hydrogenation. Chem Eng J. 381:122750. doi: 10.1016/j.cej.2019.122750.
  • Tada S, Ikeda S, Shimoda N, Honma T, Takahashi M, Nariyuki A, Satokawa S. 2017. Sponge Ni catalyst with high activity in CO2 methanation. Int J Hydrogen Energy. 42:30126–30134. doi: 10.1016/j.ijhydene.2017.10.138.
  • Tada S, Nagase H, Fujiwara N, Kikuchi R. 2021. What are the best active sites for CO2 methanation over Ni/CeO2? Energy Fuels. 35:5241–5251. doi: 10.1021/acs.energyfuels.0c04238.
  • Tada S, Shimizu T, Kameyama H, Haneda T, Kikuchi R. 2012. Ni/CeO2 catalysts with high CO2 methanation activity and high CH4 selectivity at low temperatures. Int J Hydrogen Energy. 37:5527–5531. doi: 10.1016/j.ijhydene.2011.12.122.
  • Tang G, Gong D, Liu H, Wang L. 2019. Highly loaded mesoporous Ni–La2O3 catalyst prepared by colloidal solution combustion method for CO2 methanation. Catalysts. 9:442. doi: 10.3390/catal9050442.
  • Tani T, Mädler L, Pratsinis SE. 2002. Synthesis of zinc oxide/silica composite nanoparticles by flame spray pyrolysis. J Mater Sci. 37:4627–4632. doi: 10.1023/A:1020660702207.
  • Teoh WY, Amal R, Mädler L. 2010. Flame spray pyrolysis: an enabling technology for nanoparticles design and fabrication. Nanoscale. 2:1324–1347. doi: 10.1039/c0nr00017e.
  • Wang F, Li C, Zhang X, Wei M, Evans DG, Duan X. 2015. Catalytic behavior of supported Ru nanoparticles on the {1 0 0},{1 1 0}, and {1 1 1} facet of CeO2. J Catal. 329:177–186. doi: 10.1016/j.jcat.2015.05.014.
  • Wang S, Lu GM. 1998. Role of CeO2 in Ni/CeO2–Al2O3 catalysts for carbon dioxide reforming of methane. Appl Catal B Environ. 19:267–277. doi: 10.1016/S0926-3373(98)00081-2.
  • Wegner K, Schimmöller B, Thiebaut B, Fernandez C, Rao TN. 2011. Pilot plants for industrial nanoparticle production by flame spray pyrolysis. KONA. 29:251–265. doi: 10.14356/kona.2011025.
  • Zhou J, Gao Z, Xiang G, Zhai T, Liu Z, Zhao W, Liang X, Wang L. 2022. Interfacial compatibility critically controls Ru/TiO2 metal-support interaction modes in CO2 hydrogenation. Nat Commun. 13:327. doi: 10.1038/s41467-021-27910-4.
  • Zyryanova М, Snytnikov P, Gulyaev R, Amosov YI, Boronin A, Sobyanin V. 2014. Performance of Ni/CeO2 catalysts for selective CO methanation in hydrogen-rich gas. Chem Eng J. 238:189–197. doi: 10.1016/j.cej.2013.07.034.