Next Article in Journal
Thermal and Rheological Characterization of Aqueous Nanofluids Based on Reduced Graphene Oxide (rGO) with Manganese Dioxide Nanocomposites (MnO2)
Previous Article in Journal
Manufacturing of Zinc Oxide Nanoparticle (ZnO NP)-Loaded Polyvinyl Alcohol (PVA) Nanostructured Mats Using Ginger Extract for Tissue Engineering Applications
Previous Article in Special Issue
CO2 Activation and Hydrogenation on Cu-ZnO/Al2O3 Nanorod Catalysts: An In Situ FTIR Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Silica-Decorated NiAl-Layered Double Oxide for Enhanced CO/CO2 Methanation Performance

1
Key Laboratory for Green Processing of Chemical Engineering of Xinjiang Bingtuan, School of Chemistry and Chemical Engineering, Shihezi University, Shihezi 832003, China
2
School of Chemical and Biomedical Engineering, Nanyang Technological University, Singapore 637459, Singapore
3
Carbon Neutralization and Environmental Catalytic Technology Laboratory, Bingtuan Industrial Technology Research Institute, Shihezi University, Shihezi 832003, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(17), 3041; https://doi.org/10.3390/nano12173041
Submission received: 5 July 2022 / Revised: 26 August 2022 / Accepted: 30 August 2022 / Published: 1 September 2022

Abstract

:
CO/CO2 hydrogenation has attracted much attention as a pathway to achieve carbon neutrality and production of synthetic natural gas (SNG). In this work, two-dimensional NiAl layered double oxide (2D NiAl-LDO) has been successfully decorated by SiO2 nanoparticles derived from SiCl4 and used as CO/CO2 methanation catalysts. The as-obtained H-SiO2-NiAl-LDO exhibited a large specific surface area of 201 m2/g as well as high ratio of metallic Ni0 species and surface adsorption oxygen that were beneficial for low-temperature methanation of CO/CO2. The conversion of CO methanation was 99% at 400 °C, and that of CO2 was 90% at 350 °C. At 250 °C, the CO methanation reached 85% whereas that of CO2 reached 23% at 200 °C. We believe that this provides a simple method to improve the methanation performance of CO and CO2 and a strategy for the modification of other similar catalysts.

Graphical Abstract

1. Introduction

The CO/CO2 methanation reaction is the study of the synthesis of CH4 from the hydrogenation of CO/CO2 with H2 [1,2]. The methanation reaction can capture a large amount of CO2 and CO emitted from fossil fuel combustion, such as coke-oven gas, which can promote efficient conversion and utilization of CO2. In addition, it also facilitates carbon neutrality, and when combined with renewable H2, it can produce value-added chemicals and energy substitutes [3]. In particular, methanation reactions have widely received a lot of attention [4,5] as the generated synthetic natural gas can effectively relieve pressure on natural gas supplies [6,7]. Methanation catalysts are often based on Group VIII metals (e.g., Ru [8,9,10], Rh [11], Co [12,13], Fe [14,15], Ni [16,17]) that are supported on various oxide supports. Among the active metals used in methanation reactions, precious metals such as Ru and Rh are the most reactive and selective, but their relatively high cost makes them at the disadvantage economically. The use of Co-based catalysts for methanation reactions is also commonly studied, but their conversion and selectivity are slightly inferior to those of Ni-based catalysts. However, despite having better conversion rates than Ni, Fe has been little studied as an active center, probably due to its poor selectivity. As a consequence, nickel-based catalysts have become among the most widely studied materials, as their low cost, high activity and natural abundance make them more attractive for industrial-scale applications. [18]. Ni reacts with CO at a low temperature to form nickel carbonyl, and the high exothermic property of methanation renders Ni prone to carbon deposition and sintering at high temperatures. Therefore, the main problem to be solved by the methanation reaction is the poor low-temperature performance and particles sintering at high temperatures [19,20].
Hydrotalcite is a layered double hydroxide (LDH) material, where its layered structure can offer better exposure of catalytic active sites and may accommodate various anions that are homogeneously distributed in the interlayer gallery. It also exhibits highly tunable basic sites and has a good catalytic effect on CO/CO2 methanation reactions [21,22,23]. There are many methods to prepare hydrotalcite, such as coprecipitation [24,25], hydrothermal method [26], etc. Different preparation methods have their own advantages, which can affect the catalytic performance of the catalyst. For example, Zhang et al. [27] prepared highly stable Ni-based hydrotalcite using a self-sacrificial template method for CO2 methanation reaction. The catalyst formed a structure where Ni particles were embedded into the AlOx matrix, which could minimize the agglomeration and sintering of Ni particles and improved the dispersion and stability of the catalyst. Ren et al. [20] exfoliated NiAl-LDH with an ethanol aqueous solution by ultrasonic treatment and embedded well-dispersed Ru nanoparticles onto the delaminated LDH. It was observed that the targeted activation of CO2 by the exfoliated LDH fraction was important in promoting the CO2 methanation reaction. Yao et al. [28] synthesized highly dispersed oxygen-enriched defective Ni catalysts with different NiAl molar ratios by flash nanoprecipitation (FNP) followed with delamination process in acetone. The low-temperature catalytic performance of the catalyst was found to be significantly improved, which may have been due to the large specific surface area and the defective surface of the catalyst obtained by stripping and FNP technology, respectively. The preparation method of the catalyst has an extremely important influence on dispersion and oxygen vacancy and ultimately affects the performance of the catalyst [29].
Herewith, we employed SiO2 nanoparticles derived from SiCl4 to decorate two-dimensional NiAl layered double oxide (2D NiAl-LDO) and improve its low-temperature CO methanation performance. First, the addition of SiO2 to the methanation catalyst improved and increased the specific surface area and dispersion of the catalyst [30,31]. Moreover, SiCl4 which is a by-product of the polysilicon industry is known to cause serious environmental pollution, and its treatment may pose a major ecological problem. Therefore, the re-use of silicon tetrachloride may help solve a major industrial problem [32]. In addition, we believe that it provides an additional strategy to easily improve low-temperature CO/CO2 methanation and shows potential for the application of similar catalysts.

2. Materials and Methods

2.1. Catalysts Preparation

Nickel acetate tetrahydrate (Ni(CH3COO)2·4H2O), aluminum nitrate nonahydrate (Al(NO3)3·9H2O), Na2CO3, NaOH and silicon tetrachloride (SiCl4) were purchased from McLean; all reagents are analytical grade.
For the methanation reaction, three samples were prepared, namely NiAl-LDH, SiO2-NiAl-LDH and SiCl4-NiAl-LDH, which indicate pristine NiAl-LDH, SiO2 and SiCl4 supported NiAl-LDH, respectively. In brief, 9.525 g Ni(CH3COO)2·4H2O and 4.785 g Al(NO3)3·9H2O were dissolved in 100 mL deionized water and denoted as solution A. Solution B was prepared by dissolving 3.24 g Na2CO3 in 50 mL deionized water. Lastly, 6 g NaOH was dissolved in 50 mL of deionized water and denoted as solution C. For the synthesis of NiAl-LDH, solutions A and C were added into solution B in dropwise manner while keeping constant pH of 8 by adjusting the addition rate of solution C. The final addition of solution C was 25 mL, which indicates that 3 g of NaOH was ultimately used, and after co-precipitation, the total NiAl-LDH suspension was 180 mL.
To prepare SiO2-NiAl-LDH, one third of the resulting aqueous suspension i.e., 60 mL was withdrawn and added with 300 uL of SiCl4 (98%), after which the pH of the solution dropped to ca. 6.5. Thereafter, both suspensions were then stirred at 400 rpm for 6 h at room temperature and subsequently placed in an oven at 80 °C and aged for 12 h. After aging, both suspensions were filtered with anhydrous ethanol and deionized water until the pH of the supernatant was neutral. After drying at 80 °C, the samples were ground to powder.
The dried powder sample without prior addition of SiCl4 (pure NiAl-LDH) was then equally divided into two parts, one of which was added with 25.6 mL of cyclohexane and 300 uL SiCl4 and stirred on a magnetic stirrer until the solvent evaporated, and the resulting powder was denoted as SiCl4-NiAl -LDH.
All three samples were then calcined in a muffle furnace with a heating rate of 2 °C/min up to 500 °C for 2 h, and then reduced in H2 atmosphere (99.999%) at 500 °C for 2 h to obtain oxide catalysts, namely H-NiAl-LDO, H-SiO2-NiAl-LDO and H-SiCl4-NiAl-LDO.

2.2. Catalytic Testing

A dry stainless-steel reaction tube was initially filled with quartz wool and sand followed with 0.075 g of catalyst powder. A thermocouple probe was inserted into the reaction tube and its tip was brought into contact with the catalyst bed. Upon loading the reaction tube into the reaction furnace, nitrogen gas was introduced to set the tube pressure to 0.1 MPa and it was heated to 500 °C at 5 °C/min. Upon reaching the desired temperature, hydrogen gas was then introduced to reduce the catalyst at 500 °C for 2 h at 0.1 MPa. Afterwards, hydrogen gas flow was stopped, and nitrogen was re-introduced into the reaction tube while decreasing the temperature to 150 °C. Subsequently, the gas was switched to CO syngas (H2/CO = 3:1) or CO2 syngas (H2/CO = 4/1), both with a flow rate of 65 mL/min and a space speed of 52,000 (mL g−1 h−1) before the reaction began at 150 °C. The gas analysis was performed with gas chromatograph GC9790PLUS at 50 °C intervals. Two samples were taken at each temperature point.
In this experiment, the CO, CO2 conversion and CH4 selectivity were calculated by the following formulas:
CO   Conversion   ( % ) = n CO , i n n CO , o u t n CO , i n × 100 % ,
CO 2   conversion   ( % ) = n CO 2 , i n n CO 2 , o u t n CO 2 , i n × 100 % ,
CH 4   Selectivity   ( % ) = n CH 4 , o u t n CO , i n n C O , o u t × 100 % ,

2.3. Catalyst Characterization

X-ray diffraction (XRD) is mainly used to determine the crystal phase and structure, which was performed on D8 advance from Bruker, Ettingen, Germany, using Cu Kα as the radiation source, voltage 45 KV, and current 40 mA.
Scanning electron microscope (SEM) SU8010 (Hitachi, Tokyo, Japan) was used to analyze the morphology of the catalysts.
Transmission electron microscope (TEM) FEI Tecnai G2 F30 (FEI, Hillsboro, OR, USA) was also used for morphology as well as structure and composition analysis.
Temperature-programmed reduction (H2-TPR) is to analyze the reduction profile of the active species in the catalyst by H2, from which the interactions between the active metal and the support can be determined. The instrument model was Autochem II2920 (Micromeritics, Norcross, GA, USA).
XPS (Thermo ESCALAB 250XI, Waltham, MA, USA) was used to analyze the valence state and binding energy of elements in catalyst samples.
BET analyzer (ASAP 2460), was used to determine the specific surface area, pore volume, pore size and pore distribution of the sample.
The elemental analysis of the sample was determined by ICP (Agilent Co., Ltd., Santa Clara, CA, USA, ICPOES730).

3. Results and Discussion

Figure 1a shows the XRD (X-ray diffraction) patterns of NiAl-LDH and the two catalyst precursors SiO2-NiAl-LDH and SiCl4-NiAl-LDH, where SiCl4 was added to NiAl-LDH in water and cyclohexane as solvent, respectively. The NiAl-LDH sample was found to exhibit good hydrotalcite structure with distinct (003), (006), (009), (015), (110) and (113) crystal planes, typical characteristic peaks of hydrotalcite, indicating that the NiAl-LDH catalyst precursor was successfully prepared. SiCl4-NiAl-LDH displays all of the characteristic peaks of hydrotalcite albeit having slightly lower intensity, indicating that the deposition of SiCl4 on the NiAl-LDH powder does not significantly affect the crystallinity of the sample. In addition, the characteristic peaks of SiO2 were absent, indicating highly dispersed SiO2 particles [33]. On the contrary, SiO2-NiAl-LDH exhibits lower crystallinity as shown by the broadening of the characteristic peaks.
Figure 1b shows the XRD spectra of the oxide catalysts after the H2-programmed reduction treatment (TPR). The reduced catalysts exhibited characteristic peaks at 2θ of 44.51, 51.84, and 76.37°, corresponding to the (111), (200), and (220) crystallographic planes of Ni (JADPS #04-0850), respectively, and at 37.25 and 62.88°, corresponding to the (111) and (200) crystallographic planes of NiO (JADPS #47-1049). Weak diffraction peaks for NiO were present in all of the catalysts, suggesting that a small proportion of NiO was not completely reduced, possibly due to exposure to air when the sample was extracted after reduction. No diffraction peaks of Al2O3 or other AlOx species were observed in all of the catalysts, which may have been due to the amorphous phase of Al2O3.
From the SEM (Scanning electron microscopy) images in Figure 2, H-NiAl-LDH catalyst exhibited irregular nanoparticles with considerable aggregation, which may greatly reduce the utilization of the catalyst as only a fraction of the active metals on the catalyst surface can be in contact with the reaction gas. Similarly, extensive agglomeration was also observed for H-SiCl4-NiAl-LDO sample, which hardly shows the typical plate-like hydrotalcite morphology. In contrast, H-SiO2-NiAl-LDO sample demonstrated well dispersed nanospheres which constituted assemblies of small nanosized particles. This morphology would promote better exposure of the active sites as it may increase the probability of reduction of NiO to the active metal Ni.
Figure 2d–f shows the transmission electron microscopy (TEM) images of the reduced catalysts. Similar to the SEM images, particles agglomeration was observed for H-NiAl-LDO and H-SiCl4-NiAl-LDO catalysts, whereas H-SiO2-NiAl-LDO showed better dispersion with only a few small agglomerates. The low degree of agglomeration in the latter catalyst suggested that the addition of SiCl4 to the aqueous solution during the material’s synthesis was effective in inhibiting the particles aggregation, which led to a better dispersion of the catalyst and impeded the deactivation of the catalyst at high temperatures, thereby improving the catalytic activity.
The high-resolution (HR) TEM images of the reduced catalysts (Figure 2g–i) show the lattice arrangement of the (200) crystal plane of NiO and (111) crystal plane of Ni, with lattice spacings of approximately 0.2090 and 0.2035 nm, respectively, which correlate well with the XRD patterns. The TEM energy dispersive spectroscopy (EDS) results in Figure 3 showed uniform dispersion of Ni, Al and Si elements in all of the reduced catalysts [34].
To determine the structural properties of the catalysts, we performed N2 adsorption–desorption and pore size distribution tests on the reduced catalysts. As seen in Figure 4a, all of the catalysts exhibited type-IV isotherms, which indicated the existence of uniform cylindrical pores. In particular, H-NiAl-LDO and H-SiO2-NiAl-LDO catalysts exhibited H3-type hysteresis loops, which were generally associated to the layered structural aggregates and mesoporous or macroporous materials. On the other hand, H-SiCl4-NiAl-LDO catalyst exhibited H2-type hysteresis loops, indicating the presence of mesoporous structure. As the catalytic reaction occurred on the catalyst surface, a higher specific surface area is generally more favorable as it renders more active sites. Neither the isotherm nor the hysteresis loop type of the used catalyst changed. However, after high temperature reaction, the catalyst skeleton collapsed and the smaller pores disappeared to form larger pores, so the used catalysts had larger pore sizes than fresh catalysts.
According to Table 1, the specific surface area of sample H-NiAl-LDO, H-SiO2-NiAl-LDO and H-SiCl4-NiAl-LDO are157, 210, and 147 m2/g, respectively. The results show that the addition of SiCl4 into the aqueous suspension during the synthesis of NiAl-LDH greatly increases the specific surface area of the catalyst, owing to less particle aggregation as shown by SEM and TEM images. By observing the specific surface area, pore diameter, and pore volume of the catalysts prior to and after the catalytic test, it was observed that the specific surface area of the spent catalysts decreased, which may have been due to the sintering of the catalysts [35].
As shown in Figure 4b, the pore sizes of the reduced catalysts are in the range of 2–15 nm, among which H-NiAl-LDO and H-SiCl4-NiAl-LDO have larger pore sizes, 6.72 nm and 7.21 nm, respectively, while that of H-SiO2-NiAl-LDO catalyst is 4.98 nm, suggesting the effect of the addition of SiCl4 on the pore size of the catalyst. In particular, it can be observed that there are two maxima in the pore size distribution of the H-SiCl4-NiAl-LDO catalyst (Figure 4b). The SEM image of this sample (Figure 2b) indicates that the particles in general exhibit two morphologies: small granular material at the bottom of the picture and a large oval-shaped spheres, which may have different pore sizes, resulting in two maxima in the pore size distribution. This could also imply that the addition of SiCl4 using cyclohexane as a solvent does not yield a homogeneous mixture, as can be seen in the SEM image where there are larger particles in some areas of the catalyst, which could be due to the smaller interaction between the hydrotalcite and SiCl4.
To explore the surface composition and chemical valence state of each element in the catalyst and the effect of SiCl4 on the oxygen vacancies, we conducted X-ray photoelectron spectroscopy (XPS). As shown in Figure 5a, the XPS spectrum of Ni was divided into two spin orbitals, 2p1/2 and 2p3/2. The center position of the peak at 852.8 eV corresponded to Ni0, which is the active metal Ni monomer after H2 reduction, the amount of which has an important influence on the catalytic activity. The center position at 855.4 and 873.7 eV corresponded to Ni2+, which may be due to the fact that the H2 reduction at 500 °C may not be sufficient to oxidize the difficult-to-reduce Ni-Al oxides to Ni0 and that it is difficult to avoid some Ni being oxidized during the non-in situ XPS characterization [36]. The center position at 857 and 877.7 eV corresponded to Ni3+, possibly originating from Ni2O3 [37]. A satellite peak with relatively higher binding energy can also be observed next to Ni3+ [38]. The binding energy of Ni 2p may change due to the different interaction forces of the components between the different catalysts. The proportion of Ni0 and oxygen species for different catalysts are summarized in Table 2. It can be seen that H-SiO2-NiAl-LDO catalyst had the highest proportion of Ni0, which indicate facile reduction of the NiO species under H2 atmosphere to produce active metal for the methanation reaction. In contrast, pure H-NiAl-LDH is relatively stable, where Ni interacts more strongly with the support, thus rendering reduction difficult [39]. In summary, the Ni 2p orbital spectra of the reduced catalysts showed that the binding energy between Ni and the support was decreased by the addition of SiCl4 into the aqueous synthesis suspension to produce SiO2, thereby allowing NiO species to be reduced at a lower reduction temperature [40], as demonstrated in the subsequent H2-TPR characterization.
Figure 5b compares the XPS spectra of O1s on the surface of the reduced catalysts. The main peak can be divided into three subsidiary peaks, where the peaks near 530.1, 531.2, and 532.6 eV corresponded to the lattice oxygen (Olatt), defect oxygen (Odef), and surface adsorption oxygen (Osurf), respectively. It is apparent from the graphs that the peaks for both H-SiO2-NiAl-LDO and H-SiCl4-NiAl-LDO catalysts are clearly shifted towards the ground binding energy, indicating a weakening of the interaction forces between their active centers and the carriers, allowing NiO to be reduced to Ni0 at lower temperatures. As summarized in Table 2, H-SiO2-NiAl-LDO catalyst had the highest ratio of surface adsorbed oxygen at 37% and the lattice oxygen ratio of 15%. The latter was attributed to a metal–oxygen bond with a relatively stable structure. The former was due to the presence of water on the catalyst surface [41], which was more active than the latter, thereby facilitating the reduction process and further influencing the generation of active metals [42], which eventually promotes higher reactivity. This also suggested that the H-SiO2-NiAl-LDO catalyst had a greater variety of surface hydroxyl groups, which facilitated better adsorption of CO gas during catalysis and the facile formation of intermediate bicarbonate salts and formate substances during the reaction [41]. On the other hand, the lattice oxygen ratio of H-SiCl4-NiAl-LDO catalyst was found to be significantly higher, therefore, making it harder to be reduced.
By analyzing the characteristic peaks of Si in the XPS spectra of the two samples, we found that Si was still present as SiO2 in the H-SiCl4-NiAl-LDO Catalyst with SiCl4 added with cyclohexane as the solvent. This suggests that the catalyst may contain Si–O–Al bonds, which may account for its improved performance [43]. While Si spectra was divided into two peaks in the H-SiO2-NiAl-LDO Catalyst with SiCl4 added with water as the solvent, corresponding to two compounds, SiO2(Al2O3)0.22 and SiO0.92, as shown in Figure 5c [44,45]. This suggests that the catalyst may contain Si-O-Al bonds, which may account for its improved performance.
We employed H2-TPR characterization to assess the reduction properties of the catalyst and the stability between the active component and support. According to the reduction temperature of each catalyst, it was inferred that the addition of SiCl4 in different solvents affected the interaction between Ni particles and the support. From Figure 5d, the reduction peak can generally be divided into three stages. The temperature below 450 °C is α-NiO, which belongs to the free NiO on the catalyst surface. The presence of this peak indicates weak interaction between the active metal and the support. The reduction temperature between 450 and 700 °C (middle-temperature region) can be divided into β1-NiO (450–550 °C), which belongs to the Ni-rich phase that is easily reduced, and β2-NiO (550–700 °C), which belongs to the Al-rich phase that was difficult to be reduced [46].
Figure 5d shows that the hydrogen consumption peaks of the NiAl-LDO catalyst were mainly located in the β2-NiO region (566 °C and 672 °C), thus can be hardly reduced. Both reduction peak temperatures are attributed to the Al-rich orthoclinic region, and as NiO is better reduced than Al2O3, the material reduced by the higher temperature reduction peak contains more Al. It can be clearly seen that in both reduction peaks, the substance containing more Ni consumes more H2, which means that the NiAl-LDO catalyst also contains more of N-rich orthoclinic. The SiCl4-NiAl-LDO catalyst exhibited a small reduction peak at 360 °C and 424 °C, which can be ascribed to free NiO particles located on the oxide surface [47]. Higher reduction temperature was observed at 540 °C and 594 °C, indicating the presence of both β1-NiO and β2-NiO. On the other hand, the reduction temperature peaks of the SiO2-NiAl-LDO catalyst were located at 522 °C and 549 °C, with higher H2 consumption at the former temperature, indicating that most of the NiO in the SiO2-NiAl-LDO catalyst was highly reducible to Ni0 as they belonged to the β1-NiO region with more Ni-rich phase. Compared with the NiAl-LDO catalyst, the reduction temperature of the SiO2-NiAl-LDO catalysts was significantly lower. The H2 was consumed at temperatures below 550 °C, indicating that the catalyst was easily reduced, whereby more metallic Ni was generated, which plays an important role as an active component for the methanation reaction [48,49].
Table 3 presents the quantitative analysis of the catalysts as given by inductively coupled plasma (ICP), showing the mass percentage of each element. Using the same amount of SiCl4, it was found that more Si was attached to the H-SiCl4-NiAl-LDO catalyst, this may be due to the addition of deionized water to the H-SiO2-NiAl-LDO catalyst during stirring and ageing, resulting in the loss of some of the SiCl4. However, the active metal content is lower than that of H-SiO2-NiAl-LDO, indicating that adding SiCl4 to water can retain more active components than cyclohexane. By comparing the Ni contents of the H-NiAl-LDO and H-SiO2-NiAl-LDO catalysts, it was observed that the H-SiO2-NiAl-LDO catalysts contained less active Ni metal but achieved better performance, which may have been because the H-SiO2-NiAl-LDO catalyst had a relatively high proportion of Ni0. The proportion of active metals in the catalytic process was also relatively high; therefore, a better catalytic effect can be obtained with a relatively small content of Ni.
Finally, we tested the catalytic performances of the reduced catalysts for the methanation of CO/CO2 and observed that the addition of SiCl4 in different solvents had different effects on catalyst performance. As shown in Figure 6, in general, the conversion of each catalyst increased with an increase in temperature. The optimum temperature was found to be around 300 to 400 °C, beyond which (c.a. 450 °C), the conversion of the catalyst decreased, which could be due to sintering of the catalysts at higher temperature. With the space velocity of 52,000 mL·g−1·h−1, the CO conversion of the H-SiO2-NiAl-LDO catalyst reached 85% at 250 °C and 95% at 300 °C. As a comparison, the activity of the H-NiAl-LDO and H-SiCl4-NiAl-LDO catalysts only began at 300 °C, with lower CO conversion of 84 and 89%, respectively. For CO2 methanation, H-SiO2-NiAl-LDO catalyst showed activity at lower temperature of 200 °C, with CO2 conversion of 23%. The maximum conversion of 90% was achieved at 350 °C. As a comparison, H-NiAl-LDO and H-SiCl4-NiAl-LDO catalysts only started to react after 250 °C. While H-NiAl-LDO could achieve similar CO2 conversion of 85% at 350 °C, H-SiCl4-NiAl-LDO exhibited much lower conversion of 40% at the same temperature.
The conversion of each catalyst increases with temperature, and when the temperature reaches a certain point, performance decreases significantly. There are two possible reasons for this phenomenon, which could be the effect of the time of the catalyst reaction on the catalyst performance, or it could be the collapse of the catalyst pores due to high temperatures, resulting in carbon build-up and sintering. In order to provide evidence for this phenomenon, we found that various Ni-based catalysts for CO and CO2 methanation have been shown in previous studies to suffer from reduced performance at high temperatures [50,51,52,53]. We then tested the stability of the three catalysts in this study at 350 °C and 500 °C for 10 h (Figure 7) to demonstrate that the performance degradation after 450 °C was due to high temperature catalyst sintering rather than the timing of the catalyst reaction. For CO2 methanation reaction we tested all three catalysts and for CO we chose a typical H-NiAl-LDO catalyst. It was observed that there was no decrease in stability for 10 h at the reaction temperature used for either catalyst, and the hydrotalcite catalysts generally had excellent stability, while the conversion at 500 °C was lower than at 350 °C. Therefore, the drop in performance after 450 °C is not related to the stability of the catalysts.
Overall, the results suggest there is different interactions between SiO2 and the NiAl-LDH support, depending on the preparation method, whereby the introduction of SiCl4 in the aqueous synthesis suspension significantly improved the low-temperature activity of the catalyst for CO and CO2 methanation by minimizing the particles agglomeration, increasing specific surface area, tuning the reducibility of the NiO species, and increasing the surface adsorption oxygen. Lastly, Table 4 shows the comparison of the catalytic activity of the prepared catalyst (H-SiO2-NiAl-LDO) with those of reported Ni-based catalysts. Given that higher space velocity and lower reaction temperature were used in this study, it can be concluded that the addition of SiCl4 into NiAl-LDH to form SiO2 has exhibited good low-temperature catalytic performance.

4. Conclusions

The effect of treating nickel-aluminum hydrotalcite-derived oxides with SiCl4 in water and cyclohexane on the catalyst formation was studied and the effect of SiCl4 addition was explored for the first time in the application of Ni-based hydrotalcite for methanation reaction. The performance rankings of the prepared catalysts in this study are: H-SiO2-NiAl-LDO > H-NiAl-LDO > H-SiCl4-NiAl-LDO, whereby H-SiO2-NiAl-LDO demonstrates greater performance in low-temperature methanation reaction. This could be attributed to several factors, such as better particle dispersion and low agglomeration, which leads to larger specific surface area. In addition, high reducibility of the NiO species also increased the catalytic activity as shown by lower reduction temperature under H2 environment, suggesting more metallic Ni was generated as the active species. In summary, the addition of SiCl4 to form SiO2 has a promoting effect on the low temperature performance of CO/CO2 methanation. This method of introducing SiCl4 into water to form SiO2 to modify NiAl-LDO to improve CO/CO2 methanation performance provides a strategy for the modification of other metal catalysts. In addition, further fine tuning, such as varying the amount of SiCl4 as well as adjusting the Ni/Al molar ratio may further change the particles size, surface area, and aggregation state, which will eventually affect the catalytic activity of the catalysts.

Author Contributions

Conceptualization, W.Y. and F.Y.; methodology, W.Y., J.Z. and H.Z.; data curation, W.Y., W.B. and Y.L.; writing—original draft preparation, W.Y. and J.L.; writing—review and editing, W.Y., P.G. and F.Y.; supervision, F.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by National Natural Science Foundation of China (No.22068034), Program of Science and Technology Innovation Team in Bingtuan (No.2020CB006).

Data Availability Statement

The data is available on reasonable request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rui, N.; Zhang, X.; Zhang, F.; Liu, Z.; Cao, X.; Xie, Z.; Zou, R.; Senanayake, S.D.; Yang, Y.; Rodriguez, J.A.; et al. Highly active Ni/CeO2 catalyst for CO2 methanation: Preparation and characterization. Appl. Catal. B Environ. 2021, 282, 119581. [Google Scholar] [CrossRef]
  2. Wolf, M.; Wong, L.H.; Schüler, C.; Hinrichsen, O. CO2 methanation on transition-metal-promoted Ni-Al catalysts: Sulfur poisoning and the role of CO2 adsorption capacity for catalyst activity. J. CO2 Util. 2020, 36, 276–287. [Google Scholar] [CrossRef]
  3. Wang, L.; Liu, H.; Ye, H.; Hu, R.; Yang, S.; Tang, G.; Li, K.; Yang, Y. Vacuum Thermal Treated Ni-CeO2/SBA-15 Catalyst for CO2 Methanation. Nanomaterials 2018, 8, 759. [Google Scholar] [CrossRef]
  4. Wang, X.; Zhu, L.; Liu, Y.; Wang, S. CO2 methanation on the catalyst of Ni/MCM-41 promoted with CeO2. Sci. Total Environ. 2018, 625, 686–695. [Google Scholar] [CrossRef] [PubMed]
  5. Zhu, M.; Tian, P.; Cao, X.; Chen, J.; Pu, T.; Shi, B.; Xu, J.; Moon, J.; Wu, Z.; Han, Y.-F. Vacancy engineering of the nickel-based catalysts for enhanced CO2 methanation. Appl. Catal. B Environ. 2021, 282, 119561. [Google Scholar] [CrossRef]
  6. Dewangan, N.; Hui, W.M.; Jayaprakash, S.; Bawah, A.-R.; Poerjoto, A.J.; Jie, T.; Jangam, A.; Hidajat, K.; Kawi, S. Recent progress on layered double hydroxide (LDH) derived metal-based catalysts for CO2 conversion to valuable chemicals. Catal. Today 2020, 356, 490–513. [Google Scholar] [CrossRef]
  7. Ulmer, U.; Dingle, T.; Duchesne, P.N.; Morris, R.H.; Tavasoli, A.; Wood, T.; Ozin, G.A. Fundamentals and applications of photocatalytic CO2 methanation. Nat. Commun. 2019, 10, 3169. [Google Scholar] [CrossRef] [PubMed]
  8. Lee, W.-T.; van Muyden, A.P.; Bobbink, F.D.; Huang, Z.; Dyson, P.J. Indirect CO2 Methanation: Hydrogenolysis of Cyclic Carbonates Catalyzed by Ru-Modified Zeolite Produces Methane and Diols. Angew. Chem. Int. Ed. 2019, 58, 557–560. [Google Scholar] [CrossRef]
  9. Li, X.; Han, Y.; Huang, Y.; Lin, J.; Pan, X.; Zhao, Z.; Zhou, Y.; Wang, H.; Yang, X.; Wang, A.; et al. Hydrogenated TiO2 supported Ru for selective methanation of CO in practical conditions. Appl. Catal. B Environ. 2021, 298, 120597. [Google Scholar] [CrossRef]
  10. Sakpal, T.; Lefferts, L. Structure-dependent activity of CeO2 supported Ru catalysts for CO2 methanation. J. Catal. 2018, 367, 171–180. [Google Scholar] [CrossRef]
  11. Yang, Y.; Li, Y.K.; Zhao, Y.X.; Wei, G.P.; Ren, Y.; Asmis, K.R.; He, S.G. Catalytic Co-Conversion of CH4 and CO2 Mediated by Rhodium-Titanium Oxide Anions RhTiO2. Angew. Chem. Int. Ed. 2021, 60, 13788–13792. [Google Scholar] [CrossRef]
  12. Su, H.-Y.; Yu, C.; Liu, J.-X.; Zhao, Y.; Ma, X.; Luo, J.; Sun, C.; Li, W.-X.; Sun, K. CO activation and methanation mechanism on hexagonal close-packed Co catalysts: Effect of functionals, carbon deposition and surface structure. Catal. Sci. Technol. 2020, 10, 3387–3398. [Google Scholar] [CrossRef]
  13. Tsai, Y.-T.; Mo, X.; Campos, A.; Goodwin, J.G.; Spivey, J.J. Hydrotalcite supported Co catalysts for CO hydrogenation. Appl. Catal. A Gen. 2011, 396, 91–100. [Google Scholar] [CrossRef]
  14. Luo, L.; Wang, M.; Cui, Y.; Chen, Z.; Wu, J.; Cao, Y.; Luo, J.; Dai, Y.; Li, W.X.; Bao, J.; et al. Surface Iron Species in Palladium-Iron Intermetallic Nanocrystals that Promote and Stabilize CO2 Methanation. Angew. Chem. Int. Ed. 2020, 59, 14434–14442. [Google Scholar] [CrossRef]
  15. Williamson, D.L.; Jones, M.D.; Mattia, D. Highly Selective, Iron-Driven CO2 Methanation. Energy Technol. 2019, 7, 294–306. [Google Scholar] [CrossRef]
  16. Yan, X.; Sun, W.; Fan, L.; Duchesne, P.N.; Wang, W.; Kubel, C.; Wang, D.; Kumar, S.G.H.; Li, Y.F.; Tavasoli, A.; et al. Nickel@ Siloxene catalytic nanosheets for high-performance CO2 methanation. Nat. Commun. 2019, 10, 2608. [Google Scholar] [CrossRef]
  17. Yan, X.; Yuan, C.; Bao, J.; Li, S.; Qi, D.; Wang, Q.; Zhao, B.; Hu, T.; Fan, L.; Fan, B.; et al. A Ni-based catalyst with enhanced Ni–support interaction for highly efficient CO methanation. Catal. Sci. Technol. 2018, 8, 3474–3483. [Google Scholar] [CrossRef]
  18. Tsiotsias, A.I.; Charisiou, N.D.; Yentekakis, I.V.; Goula, M.A. Bimetallic Ni-Based Catalysts for CO2 Methanation: A Review. Nanomaterials 2021, 11, 28. [Google Scholar] [CrossRef]
  19. Chen, Y.; Qiu, B.; Liu, Y.; Zhang, Y. An active and stable nickel-based catalyst with embedment structure for CO2 methanation. Appl. Catal. B Environ. 2020, 269, 118801. [Google Scholar] [CrossRef]
  20. Ren, J.; Ouyang, S.; Xu, H.; Meng, X.; Wang, T.; Wang, D.; Ye, J. Targeting Activation of CO2 and H2 over Ru-Loaded Ultrathin Layered Double Hydroxides to Achieve Efficient Photothermal CO2 Methanation in Flow-Type System. Adv. Energy Mater. 2017, 7, 1601657. [Google Scholar] [CrossRef] [Green Version]
  21. Hatta, A.H.; Jalil, A.A.; Hassan, N.S.; Hamid, M.Y.; Rahman, A.F.; Teh, L.P.; Prasetyoko, D. A review on recent bimetallic catalyst development for synthetic natural gas production via CO methanation. Int. J. Hydrogen Energy 2021, in press. [Google Scholar] [CrossRef]
  22. Huynh, H.L.; Yu, Z.X. CO2 Methanation on Hydrotalcite-Derived Catalysts and Structured Reactors: A Review. Energy Technol. 2020, 8, 1901475. [Google Scholar] [CrossRef]
  23. Sikander, U.; Sufian, S.; Salam, M.A. A review of hydrotalcite based catalysts for hydrogen production systems. Int. J. Hydrogen Energy 2017, 42, 19851–19868. [Google Scholar] [CrossRef]
  24. Mebrahtu, C.; Krebs, F.; Perathoner, S.; Abate, S.; Centi, G.; Palkovits, R. Hydrotalcite based Ni–Fe/(Mg, Al)Ox catalysts for CO2 methanation—tailoring Fe content for improved CO dissociation, basicity, and particle size. Catal. Sci. Technol. 2018, 8, 1016–1027. [Google Scholar] [CrossRef]
  25. Ren, J.; Mebrahtu, C.; Palkovits, R. Ni-based catalysts supported on Mg–Al hydrotalcites with different morphologies for CO2 methanation: Exploring the effect of metal–support interaction. Catal. Sci. Technol. 2020, 10, 1902–1913. [Google Scholar] [CrossRef]
  26. Zhang, Q.; Xu, R.; Liu, N.; Dai, C.; Yu, G.; Wang, N.; Chen, B. In situ Ce-doped catalyst derived from NiCeAl-LDHs with enhanced low-temperature performance for CO2 methanation. Appl. Surf. Sci. 2022, 579, 152204–152216. [Google Scholar] [CrossRef]
  27. Tang, S.; Yao, Y.; Chen, T.; Kong, D.; Shen, W.; Lee, H.K. Recent advances in the application of layered double hydroxides in analytical chemistry: A review. Anal. Chim. Acta 2020, 1103, 32–48. [Google Scholar] [CrossRef]
  28. Yao, Y.; Yu, F.; Li, J.; Li, J.; Li, Y.; Wang, Z.; Zhu, M.; Shi, Y.; Dai, B.; Guo, X. Two-dimensional NiAl layered double oxides as non-noble metal catalysts for enhanced CO methanation performance at low temperature. Fuel 2019, 255, 115770. [Google Scholar] [CrossRef]
  29. Vita, A.; Italiano, C.; Pino, L.; Frontera, P.; Ferraro, M.; Antonucci, V. Activity and stability of powder and monolith-coated Ni/GDC catalysts for CO2 methanation. Appl. Catal. B Environ. 2018, 226, 384–395. [Google Scholar] [CrossRef]
  30. Cisneros, S.; Chen, S.; Diemant, T.; Bansmann, J.; Abdel-Mageed, A.M.; Goepel, M.; Olesen, S.E.; Welter, E.S.; Parlinska-Wojtan, M.; Gläser, R.; et al. Effects of SiO2-doping on high-surface-area Ru/TiO2 catalysts for the selective CO methanation. Appl. Catal. B Environ. 2021, 282, 119483. [Google Scholar] [CrossRef]
  31. Li, P.; Zhu, M.; Dan, J.; Kang, L.; Lai, L.; Cai, X.; Zhang, J.; Yu, F.; Tian, Z.; Dai, B. Two-dimensional porous SiO2 nanomesh supported high dispersed Ni nanoparticles for CO methanation. Chem. Eng. J. 2017, 326, 774–780. [Google Scholar] [CrossRef]
  32. Moreno, Y.P.; Cardoso, M.B.; Ferrão, M.F.; Moncada, E.A.; dos Santos, J.H. Effect of SiCl4 on the preparation of functionalized mixed-structure silica from monodisperse sol–gel silica nanoparticles. Chem. Eng. J. 2016, 292, 233–245. [Google Scholar] [CrossRef]
  33. Tada, S.; Shoji, D.; Urasaki, K.; Shimoda, N.; Satokawa, S. Physical mixing of TiO2 with sponge nickel creates new active sites for selective CO methanation. Catal. Sci. Technol. 2016, 6, 3713–3717. [Google Scholar] [CrossRef]
  34. Yan, W.; Bao, W.; Tang, Y.; Li, Y.; Zhang, J.; Zhao, H.; Li, J.; Yu, F. Enhanced CO hydrogenation performance via two-dimensional NiAl-layered double oxide decorated by SiO2 nanoparticles. Int. J. Hydrogen Energy 2022, in press. [Google Scholar] [CrossRef]
  35. Cao, H.-X.; Zhang, J.; Guo, C.-L.; Chen, J.G.; Ren, X.-K. Highly dispersed Ni nanoparticles on 3D-mesoporous KIT-6 for CO methanation: Effect of promoter species on catalytic performance. Chin. J. Catal. 2017, 38, 1127–1137. [Google Scholar] [CrossRef]
  36. Yang, H.; Zhang, Y.; Liu, Q. Highly Efficient Ni-Phyllosilicate Catalyst with Surface and Interface Confinement for CO2 and CO Methanation. Ind. Eng. Chem. Res. 2021, 60, 6981–6992. [Google Scholar] [CrossRef]
  37. Liu, Z.-Q.; Xu, Q.-Z.; Wang, J.-Y.; Li, N.; Guo, S.-H.; Su, Y.-Z.; Wang, H.-J.; Zhang, J.-H.; Chen, S. Facile hydrothermal synthesis of urchin-like NiCo2O4 spheres as efficient electrocatalysts for oxygen reduction reaction. Int. J. Hydrogen Energy 2013, 38, 6657–6662. [Google Scholar] [CrossRef]
  38. Li, Z.; Mo, L.; Kathiraser, Y.; Kawi, S. Yolk–Satellite–Shell Structured Ni–Yolk@Ni@SiO2 Nanocomposite: Superb Catalyst toward Methane CO2 Reforming Reaction. ACS Catal. 2014, 4, 1526–1536. [Google Scholar] [CrossRef]
  39. Chang, Z.; Liu, Z.; Wang, C.; Li, J.; Zeng, J.; Liu, Y.; Zhang, M.; Li, J.; Yu, F. Three-dimensional porous Mn–Ni/Al2O3 micro-spheres for enhanced low temperature CO hydrogenation to produce methane. Int. J. Hydrogen Energy 2021, 46, 7912–7925. [Google Scholar] [CrossRef]
  40. Lv, Y.; Xin, Z.; Meng, X.; Tao, M.; Bian, Z.; Gu, J.; Gao, W. Essential role of organic additives in preparation of efficient Ni/KIT-6 catalysts for CO methanation. Appl. Catal. A Gen. 2018, 558, 99–108. [Google Scholar] [CrossRef]
  41. Zhang, M.J.; Yu, F.; Li, J.B.; Chen, K.; Yao, Y.B.; Li, P.P.; Zhu, M.Y.; Shi, Y.L.; Wang, Q.; Guo, X.H. High CO Methanation Performance of Two-Dimensional Ni/MgAl Layered Double Oxide with Enhanced Oxygen Vacancies via Flash Nanoprecipitation. Catalysts 2018, 8, 363–376. [Google Scholar] [CrossRef]
  42. Zeng, J.; Lu, K.; Zhang, J.; Sun, Y.; Chang, Z.; Li, J.; Dai, B.; Yu, F.; Li, J.; Liu, J. Solution plasma-assisted preparation of highly dispersed NiMnAl-LDO catalyst to enhance low-temperature activity of CO2 methanation. Int. J. Hydrogen Energy 2022, 47, 2234–2244. [Google Scholar] [CrossRef]
  43. Alexander, M.R.; Short, R.D.; Jones, F.R.; Stollenwerk, M.; Zabold, J.; Michaeli, W. An X-ray photoelectron spectroscopic investigation into the chemical structure of deposits formed from hexamethyldisiloxane/oxygen plasmas. J. Mater. Sci. 1996, 31, 1879–1885. [Google Scholar] [CrossRef]
  44. Alfonsetti, R.; Lozzi, L.; Passacantando, M.; Picozzi, P.; Santucci, S. XPS studies on SiOx thin films. Appl. Surf. Sci. 1993, 70–71, 222–225. [Google Scholar] [CrossRef]
  45. Ayame, A.; Kitagawa, T. X-Ray photoelectron spectroscopic measurement and chemical characteristics of silica, alumina and silica-alumina. Bunseki Kagaku 1991, 40, 673–678. [Google Scholar] [CrossRef]
  46. Zhao, A.; Ying, W.; Zhang, H.; Ma, H.; Fang, D. Ni–Al2O3 catalysts prepared by solution combustion method for syngas methanation. Catal. Commun. 2012, 17, 34–38. [Google Scholar] [CrossRef]
  47. Lin, X.; Lin, L.; Huang, K.; Chen, X.; Dai, W.; Fu, X. CO methanation promoted by UV irradiation over Ni/TiO2. Appl. Catal. B Environ. 2015, 168–169, 416–422. [Google Scholar] [CrossRef]
  48. Guo, C.L.; Wu, Y.Y.; Qin, H.Y.; Zhang, J.L. CO methanation over ZrO2/Al2O3 supported Ni catalysts: A comprehensive study. Fuel Process. Technol. 2014, 124, 61–69. [Google Scholar] [CrossRef]
  49. Zhang, X.; Rui, N.; Jia, X.; Hu, X.; Liu, C.-J. Effect of decomposition of catalyst precursor on Ni/CeO2 activity for CO methanation. Chin. J. Catal. 2019, 40, 495–503. [Google Scholar] [CrossRef]
  50. Jomjaree, T.; Sintuya, P.; Srifa, A.; Koo-Amornpattana, W.; Kiatphuengporn, S.; Assabumrungrat, S.; Sudoh, M.; Watanabe, R.; Fukuhara, C.; Ratchahat, S. Catalytic performance of Ni catalysts supported on CeO2 with different morphologies for low-temperature CO2 methanation. Catal. Today 2021, 375, 234–244. [Google Scholar] [CrossRef]
  51. Li, S.; Gong, D.; Tang, H.; Ma, Z.; Liu, Z.-T.; Liu, Y. Preparation of bimetallic Ni@Ru nanoparticles supported on SiO2 and their catalytic performance for CO methanation. Chem. Eng. J. 2018, 334, 2167–2178. [Google Scholar] [CrossRef]
  52. Li, S.; Liu, G.; Zhang, S.; An, K.; Ma, Z.; Wang, L.; Liu, Y. Cerium-modified Ni-La2O3/ZrO2 for CO2 methanation. J. Energy. Chem. 2020, 43, 155–164. [Google Scholar] [CrossRef]
  53. Lv, Y.; Xin, Z.; Meng, X.; Tao, M.; Bian, Z.; Gu, J.; Gao, W. Effect of La, Mg and Mo additives on dispersion and thermostability of Ni species on KIT-6 for CO methanation. Appl. Catal. A Gen. 2017, 543, 125–132. [Google Scholar] [CrossRef]
  54. Ahmad, F.; Lovell, E.C.; Masood, H.; Cullen, P.J.; Ostrikov, K.K.; Scott, J.A.; Amal, R. Low-Temperature CO2 Methanation: Synergistic Effects in Plasma-Ni Hybrid Catalytic System. ACS Sustain. Chem. Eng. 2020, 8, 1888–1898. [Google Scholar] [CrossRef]
  55. Bukhari, S.N.; Chong, C.C.; Teh, L.P.; Vo, D.V.N.; Ainirazali, N.; Triwahyono, S.; Jalil, A.A.; Setiabudi, H.D. Promising hydrothermal technique for efficient CO2 methanation over Ni/SBA-15. Int. J. Hydrogen Energy 2019, 44, 20792–20804. [Google Scholar] [CrossRef]
  56. Cerdá-Moreno, C.; Chica, A.; Keller, S.; Rautenberg, C.; Bentrup, U. Ni-sepiolite and Ni-todorokite as efficient CO2 methanation catalysts: Mechanistic insight by operando DRIFTS. Appl. Catal. B Environ. 2020, 264, 118546. [Google Scholar] [CrossRef]
  57. Dai, Y.; Xu, M.; Wang, Q.; Huang, R.; Jin, Y.; Bian, B.; Tumurbaatar, C.; Ishtsog, B.; Bold, T.; Yang, Y. Enhanced activity and stability of Ni/La2O2CO3 catalyst for CO2 methanation by metal-carbonate interaction. Appl. Catal. B Environ. 2020, 277, 119271. [Google Scholar] [CrossRef]
  58. Wierzbicki, D.; Moreno, M.V.; Ognier, S.; Motak, M.; Grzybek, T.; Da Costa, P.; Gálvez, M.E. Ni-Fe layered double hydroxide derived catalysts for non-plasma and DBD plasma-assisted CO2 methanation. Int. J. Hydrogen Energy 2020, 45, 10423–10432. [Google Scholar] [CrossRef]
  59. Zhang, F.; Lu, B.; Sun, P. Highly stable Ni-based catalysts derived from LDHs supported on zeolite for CO2 methanation. Int. J. Hydrogen Energy 2020, 45, 16183–16192. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction patterns of (a) NiAl-LDH, SiO2-NiAl-LDH, and SiCl4-NiAl-LDH and (b) H-NiAl-LDO, H-SiO2-NiAl-LDO, and H-SiCl4-NiAl-LDO.
Figure 1. X-ray diffraction patterns of (a) NiAl-LDH, SiO2-NiAl-LDH, and SiCl4-NiAl-LDH and (b) H-NiAl-LDO, H-SiO2-NiAl-LDO, and H-SiCl4-NiAl-LDO.
Nanomaterials 12 03041 g001
Figure 2. Scanning electron microscopy images of (a) H-NiAl-LDO, (b) H-SiCl4-NiAl-LDO, and (c) H-SiO2-NiAl-LDO; Transmission electron microscopy (TEM) images of (d) H-NiAl-LDO, (e) H-SiCl4-NiAl-LDO and (f) H-SiO2-NiAl-LDO and High-resolution TEM images of (g) H-NiAl-LDO, (h) H-SiCl4-NiAl-LDO, and (i) H-SiO2-NiAl-LDO.
Figure 2. Scanning electron microscopy images of (a) H-NiAl-LDO, (b) H-SiCl4-NiAl-LDO, and (c) H-SiO2-NiAl-LDO; Transmission electron microscopy (TEM) images of (d) H-NiAl-LDO, (e) H-SiCl4-NiAl-LDO and (f) H-SiO2-NiAl-LDO and High-resolution TEM images of (g) H-NiAl-LDO, (h) H-SiCl4-NiAl-LDO, and (i) H-SiO2-NiAl-LDO.
Nanomaterials 12 03041 g002
Figure 3. TEM-energy dispersive spectroscopy (TEM-EDS) images of (a) H-NiAl-LDO (Reprinted with permission from [34], Copyright 2022 Elsevier), (b) H-SiCl4-NiAl-LDO, (c) H-SiO2-NiAl-LDO.
Figure 3. TEM-energy dispersive spectroscopy (TEM-EDS) images of (a) H-NiAl-LDO (Reprinted with permission from [34], Copyright 2022 Elsevier), (b) H-SiCl4-NiAl-LDO, (c) H-SiO2-NiAl-LDO.
Nanomaterials 12 03041 g003
Figure 4. Nitrogen adsorption-desorption isotherms of (a) reduced samples and (c) used samples; Pore size distribution of (b) reduced samples and (d) used samples.
Figure 4. Nitrogen adsorption-desorption isotherms of (a) reduced samples and (c) used samples; Pore size distribution of (b) reduced samples and (d) used samples.
Nanomaterials 12 03041 g004
Figure 5. (a) Ni 2p XPS spectra, (b) O 1s XPS spectra, (c) Si 2p XPS spectra and (d) H2-TPR profiles of H-NiAl-LDO, H-SiCl4-NiAl-LDO and H-SiO2-NiAl-LDO catalysts.
Figure 5. (a) Ni 2p XPS spectra, (b) O 1s XPS spectra, (c) Si 2p XPS spectra and (d) H2-TPR profiles of H-NiAl-LDO, H-SiCl4-NiAl-LDO and H-SiO2-NiAl-LDO catalysts.
Nanomaterials 12 03041 g005
Figure 6. Catalytic performances of catalysts: (a) CO conversion, (b) CH4 selectivity (CO); (c) CO2 conversion, (d) CH4 selectivity (CO2).
Figure 6. Catalytic performances of catalysts: (a) CO conversion, (b) CH4 selectivity (CO); (c) CO2 conversion, (d) CH4 selectivity (CO2).
Nanomaterials 12 03041 g006
Figure 7. (ad) The 10 h CO/CO2 methanation performance of each catalyst.
Figure 7. (ad) The 10 h CO/CO2 methanation performance of each catalyst.
Nanomaterials 12 03041 g007
Table 1. Textural properties of the prepared catalysts.
Table 1. Textural properties of the prepared catalysts.
SamplesSBET (m2·g−1)DBJH (nm)Vp (cm3·g−1)
NiAl-LDO1576.720.38
SiCl4-NiAl-LDO1477.210.38
SiO2-NiAl-LDO2104.980.34
NiAl-LDO-Used1148.370.35
SiCl4-NiAl-LDO-Used1327.650.37
SiO2-NiAl-LDO-Used1636.300.34
Table 2. Quantification of the metallic Ni and oxygen species of the prepared catalysts.
Table 2. Quantification of the metallic Ni and oxygen species of the prepared catalysts.
CatalystsNi0/(Ni0 + Ni2+ + Ni3+)Osurf/(Osurf + Odef + Olatt)Olatt/(Osurf + Odef + Olatt)
H-NiAl-LDO0.130.120.49
H-SiCl4-NiAl-LDO0.190.220.44
H-SiO2-NiAl-LDO0.210.370.15
Table 3. Elemental analysis of the prepared catalysts.
Table 3. Elemental analysis of the prepared catalysts.
SamplesH-NiAl-LDOH-SiCl4-NiAl-LDOH-SiO2-NiAl-LDO
Ni (wt%)60.6251.3353.31
Al (wt%)10.435.707.29
Si (wt%)04.703.02
Table 4. Comparison of low temperature catalysts prepared via different methods.
Table 4. Comparison of low temperature catalysts prepared via different methods.
CatalystsWHSV (mL·g−1·h−1)T (°C)CO2 Conversion (%)Ref
15% Ni/Al2O3 (Plasma)870025040[54]
Ni/SBA-1515,00025082[55]
Ni-sepiolite36,00025010[56]
Ni/La2O2CO320,00030025[57]
Ni20Fe1.5-LDH12,00025076[58]
Ni/MgAl-LDH501725020[25]
NaY-Ni5-LDHs30,00025013[59]
H-SiO2-NiAl-LDO52,00020023This study
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yan, W.; Li, Y.; Zeng, J.; Bao, W.; Zhao, H.; Li, J.; Gunawan, P.; Yu, F. Silica-Decorated NiAl-Layered Double Oxide for Enhanced CO/CO2 Methanation Performance. Nanomaterials 2022, 12, 3041. https://doi.org/10.3390/nano12173041

AMA Style

Yan W, Li Y, Zeng J, Bao W, Zhao H, Li J, Gunawan P, Yu F. Silica-Decorated NiAl-Layered Double Oxide for Enhanced CO/CO2 Methanation Performance. Nanomaterials. 2022; 12(17):3041. https://doi.org/10.3390/nano12173041

Chicago/Turabian Style

Yan, Wenxia, Yangyang Li, Junming Zeng, Wentao Bao, Huanhuan Zhao, Jiangbing Li, Poernomo Gunawan, and Feng Yu. 2022. "Silica-Decorated NiAl-Layered Double Oxide for Enhanced CO/CO2 Methanation Performance" Nanomaterials 12, no. 17: 3041. https://doi.org/10.3390/nano12173041

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop