Next Article in Journal
An Efficient Strategy for the Fabrication of CuS as a Highly Excellent and Recyclable Photocatalyst for the Degradation of Organic Dyes
Next Article in Special Issue
DeNOx of Nano-Catalyst of Selective Catalytic Reduction Using Active Carbon Loading MnOx-Cu at Low Temperature
Previous Article in Journal
Highly Efficient Biphasic System for the Synthesis of Alkyl Lactates in the Presence of Acidic Ionic Liquids
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Encapsulated Ni@La2O3/SiO2 Catalyst with a One-Pot Method for the Dry Reforming of Methane

1
Department of Chemical Engineering, School of Petrochemical Technology and Energy Engineering, Zhejiang Ocean University, Zhoushan 316022, Zhejiang, China
2
School of Naval Architecture and Mechanical-electrical Engineering, Zhejiang Ocean University, Zhoushan 316022, Zhejiang, China
3
School of Food and Pharmaceutical, Zhejiang Ocean University, Zhoushan 316022, Zhejiang, China
4
Institute of Innovation & Application, Zhejiang Ocean University, Zhoushan 316022, Zhejiang, China
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(1), 38; https://doi.org/10.3390/catal10010038
Submission received: 22 November 2019 / Revised: 17 December 2019 / Accepted: 25 December 2019 / Published: 28 December 2019

Abstract

:
Ni nanoparticles encapsulated within La2O3 porous system (Ni@La2O3), the latter supported on SiO2 (Ni@La2O3)/SiO2), effectively inhibit carbon deposition for the dry reforming of methane. In this study, Ni@La2O3/SiO2 catalyst was prepared using a one-pot colloidal solution combustion method. Catalyst characterization demonstrates that the amorphous La2O3 layer was coated on SiO2, and small Ni nanoparticles were encapsulated within the layer of amorphous La2O3. During 50 h of dry reforming of methane at 700 °C and using a weight hourly space velocity (WHSV) of 120,000 mL gcat−1 h−1, the CH4 conversion obtained was maintained at 80%, which is near the equilibrium value, while that of impregnated Ni–La2O3/SiO2 catalyst decreased from 63% to 49%. The Ni@La2O3/SiO2 catalyst exhibited very good resistance to carbon deposition, and only 1.6 wt% carbon was formed on the Ni@La2O3/SiO2 catalyst after 50 h of reaction, far lower than that of 11.5 wt% deposited on the Ni–La2O3/SiO2 catalyst. This was mainly attributed to the encapsulated Ni nanoparticles in the amorphous La2O3 layer. In addition, after reaction at 700 °C for 80 h with a high WHSV of 600,000 mL gcat−1 h−1, the Ni@La2O3/SiO2 catalyst exhibited high CH4 conversion rate, ca. 10.10 mmol gNi−1 s−1. These findings outline a simple synthesis method to prepare supported encapsulated Ni within a metal oxide porous structure catalyst for the dry reforming of methane reaction.

Graphical Abstract

1. Introduction

Dry reforming of methane (DRM) is a promising process, as it can simultaneously convert CO2 and CH4 present in CO2-rich natural gas reservoirs to produce syngas. The latter serves as the raw material to produce liquid fuels through gas-to-liquid technology (via Fischer–Tropsch synthesis) [1]. Due to the strong endothermic nature of the DRM reaction, most of previous studies were conducted at temperatures higher than 600 °C for high conversions of CO2 and CH4. Therefore, developing a robust catalyst that possesses good stability and excellent resistance to coke plays a crucial role in the DRM reaction [2,3,4].
C O 2 + C H 4 2 C O + 2 H 2 ,   Δ H 298   K = + 247   k J   m o l 1
Many catalysts have been investigated for the DRM reaction. As a result, precious-metals-based catalysts, for instance, Ru, Rh, Pd, Ir, Pt, etc., exhibited good catalytic activity and stability. However, the disadvantage of using noble-metals-based catalysts is their high cost [5,6,7]. Ni-based catalysts, due to good catalytic activity and low cost, have been widely investigated for the DRM reaction [8,9,10]. However, Ni-based catalysts are prone to carbon deposition and metal sintering during DRM [11,12]. Designing a DRM catalyst that resists against carbon deposition and metal sintering could be accomplished by making appropriate choice of support, promoter, structure, and methods of preparation [2,3,4]. Because oxygen species in the CeO2 lattice can effectively lower carbon accumulation by oxidation of accumulated carbon, various Ni/CeO2 compositions were extensively studied to prevent carbon deposition [13,14,15,16]. In addition, small metallic Ni nanoparticles more effectively inhibit the nucleation and growth of coke, thereby restraining coke deposition on the catalyst [17,18]. Ni particles, smaller than 5 nm, can effectively reduce carbon deposition [15,18,19,20,21]. Small Ni particles also have poor thermal stability, especially at high reaction temperatures. Encapsulated Ni-based catalysts [22,23,24,25,26,27], such as core-shell Ni@SiO2 [28,29,30], were used to suppress carbon deposition and enhance the stability of small Ni particles in the DRM reaction. Sandwiched SiO2@Ni@CeO2 [31] and SiO2@Ni@ZrO2 [32] catalysts were applied for the DRM, exhibiting high catalytic activity and excellent coke resistance. However, encapsulated catalysts usually require complex preparation processes. The development of a simple method for preparing encapsulated Ni-based catalyst is thus required.
As La2O3 can promote CO2 adsorption and activation [33]; Ni/La2O3 [34,35] and La2O3 promoted Ni-based catalysts [36,37,38,39] have shown excellent activity in the DRM reaction. However, due to the low specific surface area of La2O3, the main problem with Ni/La2O3 catalysts is the poor dispersion of Ni [40,41], leading to carbon deposition on Ni/La2O3 catalysts [40,41]. Li et al. [42] reported ordered mesoporous Ni/La2O3 catalysts with large specific surface area for DRM, and the experimental result indicated that the increased interface between Ni and La2O3 is beneficial for suppressing carbon deposition.
A colloidal solution combustion method has been reported to prepared mesoporous CeO2 [43]. In this study, a novel Ni@La2O3/SiO2 catalyst was prepared via one-pot synthesis using the colloidal solution combustion method. For the Ni@La2O3/SiO2 catalyst, small Ni particles were encapsulated within an amorphous La2O3 layer and supported on SiO2. The prepared Ni@La2O3/SiO2 catalyst has an abundant interface between Ni and La2O3, which was more active and stable compared to the Ni–La2O3/SiO2 catalyst prepared using the standard impregnation method.

2. Results and Discussion

2.1. Characterization of Fresh and Reduced Catalysts

The N2 adsorption results of the Ni@La2O3/SiO2 and Ni–La2O3/SiO2 catalysts are listed in Table 1. The specific surface area (SBET) of the Ni@La2O3/SiO2 is 19.0 m2 g−1, which is smaller than that of the Ni–La2O3/SiO2 catalyst. The pore volume and average pore size of Ni@La2O3/SiO2 catalyst are 0.21 cm3 g−1 and 43.9 nm, respectively, which are larger than those of Ni–La2O3/SiO2 catalyst.
It is noted that the Ni@La2O3/SiO2 catalyst has a significantly lower specific surface area compared with the recently reported mesoporous Ni–La2O3 (172 m2.g−1) [42] and Ni–La2O3/SiO2 (190 m2 g−1) [38] catalysts for DRM. In our previous report [43], mesoporous Ni–La2O3 (70.4 m2 g−1) had been synthesized by the same colloidal solution combustion method with colloidal SiO2 as a template, and the silica was then removed by NaOH etching to form mesopores. Compared with our previously reported mesoporous Ni–La2O3 catalysts, the Ni@La2O3/SiO2 catalyst may be more suitable for high-temperature reactions due to the use of silica as support of La2O3 and Ni.
Figure 1 shows the powder XRD patterns of the fresh and reduced catalysts. For the fresh Ni–La2O3/SiO2, the peaks at 2θ = 37.2°, 43.3°, and 62.9° are attributed to NiO [44,45]. The reduced Ni–La2O3/SiO2 showed weak Ni peaks at 44.5° and 51.7° [44,45]. As shown in Table 1, the crystallite size of Ni in reduced Ni–La2O3/SiO2 is 16.4 nm, which is about twice that of NiO in fresh Ni–La2O3/SiO2. This indicates that NiO in the fresh Ni–La2O3/SiO2 catalyst is unstable during the reduction process and sintering. In contrast, there are no obviously NiO peaks found in fresh Ni@La2O3/SiO2 catalyst, and a broad Ni peak at 44.5° is found in reduced Ni@La2O3/SiO2 catalyst. This indicates that Ni particle size in the reduced Ni@La2O3/SiO2 catalyst is smaller than that in the reduced Ni–La2O3/SiO2 catalyst. It should be noted here that no La2O3 peak was found in any catalyst. This might be due to the very small La2O3 crystals formed not able to detect by XRD, or that the La2O3 was in the amorphous phase [37]. The results indicate that La2O3 was highly dispersed or amorphous in these catalytic systems. Similar results were reported in the literature [38]. The morphology of La2O3 (small particle or amorphous phase) needs to be further confirmed by TEM analysis.
The TEM images of the fresh and reduced catalysts are shown in Figure 2. For the fresh Ni–La2O3/SiO2 catalyst, dark aggregated NiO particles and SiO2 particles~20 nm with a smooth surface are shown in Figure 2a, indicating that most of NiO was not loaded onto SiO2 supported but aggregated instead. The particle size distribution of NiO is displayed in the inset of Figure 2a. The representative high-resolution TEM images in Figure 2b show lattice fringes corresponding to La2O3 and NiO, thus illustrating the formation of NiO and La2O3 in the fresh Ni–La2O3/SiO2 catalyst.
Figure 2c,d show TEM images of the fresh Ni@La2O3/SiO2 catalyst. SiO2 nanospheres are coated with a layer of amorphous La2O3, and there are no NiO particles observed, indicating that small NiO particles are highly dispersed on the La2O3 coating. A TEM image of the reduced Ni@La2O3/SiO2 catalyst is shown in Figure 2e. As observed, SiO2 is coated with an amorphous La2O3 layer, on which metallic Ni particles are encapsulated within the amorphous La2O3 layer. An average Ni particle size about 3.5 nm was obtained by counting more than 100 Ni particles as shown in Figure 2f. This result is consistent with our previous report of mesoporous Ni–La2O3 prepared via colloidal solution combustion method [44].
Combined with TEM and SBET results, it can be concluded that the lower SBET of Ni@La2O3/SiO2 may be due to the fact that the surface of SiO2 nanoparticles was covered with La2O3 and NiO. On the contrary, the SiO2 nanoparticles in the Ni@La2O3/SiO2 catalyst structure may not be completely covered by La2O3 and NiO, and the exposed SiO2 surface resulted in a slightly larger SBET.
The BF-STEM images and the element distribution profiles of Si, La, and Ni are shown in Figure 3. SiO2 is surrounded by La2O3 and Ni nanoparticles. The signal of Ni is accompanied by the existence of La, but the signal of La is not necessarily accompanied by Ni, indicating that the nickel is encapsulated by La2O3. The signal of La2O3 is distributed around the signal of silica, indicating that SiO2 is encapsulated by La2O3. The Ni particles are encapsulated by amorphous La2O3 and wrapped on silica.
The H2-TPR profiles of the catalysts are shown in Figure 4. The Ni–La2O3/SiO2 catalyst exhibits two reduction peaks. The first peak at 340 °C corresponds to the reduction of free NiO [21,39,46]. The second peak at 385 °C corresponds to the reduction of NiO with weak interaction with La2O3 or SiO2 [40,42,47,48]. These results indicate that NiO is weakly interacting or not interacting at all with La2O3 or SiO2 in the Ni–La2O3/SiO2 catalyst.
The Ni@La2O3/SiO2 catalyst displays a broad reduction peak at 615 °C, suggesting that Ni-based species have a strong interaction with the support [42,49]. Li et al. [42] found that the small NiO particle confined into mesoporous La2O3 strongly interacts with La2O3 support, resulting in a high reduction temperature for NiO. Also, the Ni@La2O3/SiO2 catalyst, which possesses the encapsulated structure of metal Ni by La2O3 layer on SiO2 (in Figure 3a), exhibits high reduction temperatures. Thus, the high reduction temperature at 615 °C is due to the reduction of NiO, which presents strong interactions with the encapsulated La2O3 layer.

2.2. Catalytic Performance Studies

Figure 5 shows results of the catalytic performance tests conducted over the Ni@La2O3/SiO2 and Ni–La2O3/SiO2 catalysts in the DRM at 700 °C with a weight hourly space velocity (WHSV) of 120,000 mL g−1 h−1 and after 50 h of reaction. At 700 °C, the thermodynamic equilibrium conversion of CO2 and CH4 were 90.2% and 84.4%, respectively. It should be noted that this thermodynamic equilibrium consists of the DRM reaction and reverse water-gas shift reaction (RWGS: CO2 + H2 ↔ CO + H2O). As shown in Figure 5a, the CO2 conversion on the Ni@La2O3/SiO2 catalyst is 90% and reached the equilibrium conversion. The CO2 conversion in the two catalysts are higher than the CH4 conversion, a result which is mainly due to the RWGS reaction [31]. The Ni@La2O3/SiO2 catalyst exhibits stable CO2 and CH4 conversions during the DRM reaction period of 50 h. In contrast, the CO2 and CH4 conversions of the Ni–La2O3/SiO2 catalyst decrease from 75% to 62% and from 63% to 49%, within 50 h, respectively. Therefore, the Ni@La2O3/SiO2 catalyst has better activity and stability behavior than the Ni–La2O3/SiO2 catalyst during DRM.
As shown in Figure 5b, the H2/CO ratio is lower than one for both catalysts. This result is mainly due to the RWGS reaction and to a less degree to other side reactions [31], such as the reverse Boudouard reaction (C + CO2 ↔ 2 CO) [49]. The CO and H2 yields are shown in Figure 5c,d, respectively. For each catalyst, the CO yield is higher than the H2 yield. Comparing the two catalysts, the Ni@La2O3/SiO2 catalyst exhibits higher H2 and CO yields than the Ni–La2O3/SiO2 catalyst.
No sign of deactivation was observed for the Ni@La2O3/SiO2 catalyst with a WHSV of 120,000 mL g−1 h−1, as shown in Figure 5, and Figure 6 illustrates the effect of WHSV on the catalytic performance of the Ni@La2O3/SiO2 catalyst in the range of 120,000 to 1,200,000 mL g−1 h−1.
As shown in Figure 6a, in the 120,000–300,000 mL g−1 h−1 range, the CO2 and CH4 conversions obviously remain constant. When the WHSV increased to 600,000 mL g−1 h−1, the conversions start to decrease slightly, but the CH4 conversion is still higher than 70%, suggesting the very good activity of the catalyst. In particular, when the WHSV is increased to 1,200,000 mL g−1 h−1, the conversions of CH4 and CO2 are decreased to 51.7% and 65.7%. Accordingly, the CO and H2 yields show the same trend, which are decreased significantly when the WHSV becomes larger than 600,000 mL g−1 h−1. This behavior with WHSV is largely related to external mass transport effects established within the catalytic bed.
Figure 7 shows results of the stability test of Ni@La2O3/SiO2 catalyst conducted at a high WHSV of 600,000 mL g−1 h−1. Although the CH4 and CO2 conversions are slightly decreased over 80 h of reaction, the CH4 conversion is still as high as 65% after reaction, indicating the very good activity and stability of the Ni@La2O3/SiO2 catalyst.
Table 2 lists the methane conversion rates obtained over Ni@La2O3/SiO2 and some representative Ni-based catalysts reported in the literature. As listed in Table 2, the CH4 conversion rate of Ni@La2O3/SiO2 catalyst is five times higher than that of Ni–La2O3/SiO2 catalyst. Although the literature in Table 2 is limited, it appears that the Ni@La2O3/SiO2 catalyst has a significantly better methane conversion rate and coke resistance in the DRM reaction at the conditions applied.

2.3. Characterization of Used Catalysts

To measure the amount of deposited carbon on the used catalysts, TG and DTA tests were conducted, and the obtained results are shown in Figure 8a,b. For the used Ni–La2O3/SiO2 catalyst, the weight loss is 11.5 wt% in the range 500–700 °C, and the DTA exhibits an obvious exothermic peak due to the oxidation of deposited carbon. The weight loss of the used Ni@La2O3/SiO2 catalyst was only 1.6 wt% over 50 h in DRM, which is significantly lower than that obtained in the used Ni–La2O3/SiO2 catalyst (11.5 wt% carbon deposition). These results indicate that Ni@La2O3/SiO2 catalyst has a very good resistance to carbon deposition.
The reason for the very good resistance to carbon deposition for the aforementioned catalyst might be partly due to the formed La2O3 layer, within which the Ni metallic active sites are well dispersed and less prone to carbon accumulation. As recently reported [52], the Ce0.8Pr0.2O2-δ- supported Ni catalyst prepared by the citrate sol-gel method, due to the presence of mobile active oxygen species in the Ce0.8Pr0.2O2-δ support, largely participates in the carbon removal via gasification to CO(g). Moreover, Ni particles smaller in size can reduce carbon accumulation [53]. Therefore, the good carbon resistance exhibited by Ni@La2O3/SiO2 catalyst seems to be largely related to the smaller nickel particle size and the presence of La2O3 coating layer on the Ni particles.
Based on the weight loss range of temperatures ca. 480–730 °C along the exothermic peak, the carbon deposited is composed of whisker carbon and encapsulated graphitic carbon. It should be noted that the encapsulated graphitic carbon is usually responsible for catalyst deactivation [54]. However, the whisker carbon, which possesses hollow structure, has little effect on the active sites of metallic Ni, and therefore is not the main reason for catalyst deactivation.
Figure 9 shows the XRD patterns of the used catalysts. After a 50 h DRM reaction, the used Ni–La2O3/SiO2 catalyst shows an obvious graphitic peak (2θ = 26.5°) [31] and Ni peaks (2θ = 44.5° and 51.8°) as well. As shown in Table 1, the crystallite size of Ni in the used catalyst increased to 44.6 nm, indicating that Ni particles in the Ni–La2O3/SiO2 catalyst are not stable during the DRM reaction, and large Ni particles and carbon deposition are formed during reaction. The weak Ni diffraction peaks found in the used Ni@La2O3/SiO2 catalyst indicate that Ni particles become smaller after DRM for 50 h. These results indicate that Ni particles in the Ni@La2O3/SiO2 catalyst are more stable than in Ni@La2O3/SiO2 catalyst. In addition, no obvious diffraction peak of graphitic carbon is observed in the used Ni@La2O3/SiO2 catalyst. As proved by the TG-DTA analysis, the amount of carbon deposition (1.6 wt%) is rather small to be detected by powder XRD.
TEM images of the used catalysts are shown in Figure 10. No whisker carbon was found in the used Ni@La2O3/SiO2 (Figure 10a,b). The Ni mean particle size is about 5 nm, which is slightly larger than that in the reduced Ni@La2O3/SiO2, indicating that the small Ni particles encapsulated into the amorphous La2O3 layer are thermally stable and largely contribute to the inhibition of carbon deposition.
As shown in Figure 10, accessible nickel particles (without encapsulation) favor the formation of carbon. Carbon nanofibers and encapsulated graphitic carbon were formed over the used Ni–La2O3/SiO2 catalyst. Methane decomposes on the nickel surface forming atomic hydrogen and carbon, the latter diffusing to free surface sites on the nickel particle to form graphitic carbon (the graphitic carbon peak is precisely seen in Figure 9) [55,56]. As shown in Figure 10c, Ni particles in the used Ni–La2O3/SiO2 catalyst are in the 10–50 nm range, which is much wider than that found in the used Ni@La2O3/SiO2 catalyst. These results are consistent with the TG and XRD results of the used catalysts. It is mentioned here that deposited carbon can plug the reactor and reduce the lifetime of the catalyst as well. However, in general, the formation of carbon nanofibers does not decrease the exposed Ni surface area of the catalyst, thereby maintaining stable catalytic activity. Ni sintering and the formation of encapsulated carbon can reduce the exposed Ni surface area and thus result in catalyst deactivation.
Based on the Ni@La2O3/SiO2 catalyst structure features, it is reasonable to propose that its excellent activity, stability, and high resistance to carbon deposition are much related to the small Ni particles encapsulated within the amorphous La2O3 layer deposited on SiO2. Chen et al. [38] has prepared Ni–La2O3/SiO2 via one-pot sol-gel method with large specific surface area (190 m2 g−1), which exhibited high activity and excellent stability for DRM at 700 ℃. However, the rate of deposited carbon on Ni–La2O3/SiO2 was 5.9 mg C gcat−1 h−1, which is much higher than that found in the present Ni@La2O3/SiO2 catalyst (0.32 mg C gcat−1 h−1). Although the BET surface area of the Ni@La2O3/SiO2 catalyst is relatively low, the amorphous La2O3 layer can encapsulate and stabilize the small nickel particles formed, thus resulting in an active and stable Ni@La2O3/SiO2 catalyst. Based on the XRD, TEM, and TPR results, it can be concluded that the formed Ni@La2O3/SiO2 catalyst structure cannot only stabilize small nickel particles and reduce carbon accumulation, but also provides more interface between Ni and La2O3. The latter can promote CO2 activation on oxygen vacant sites and on highly basic nature oxygen sites (lanthana oxycarbonates), which was found to be beneficial for inhibiting carbon deposition and enhancing catalytic performance [42,57].
Even though the encapsulation of nanometal particles in core-shell or yolk-shell structures for stabilizing nanometal particles and inhibiting carbon deposition for high-temperature reactions have been reported in the literature, encapsulated metal catalyst using an inert shell, such as SiO2, can always result in lower activity due to the blockage of active sites [58]. There is still research demand to develop a simple method for preparing encapsulated metal catalysts with high activity. In this work, a simple colloidal solution combustion method was used to prepare a Ni@La2O3/SiO2 catalyst with small Ni particles encapsulated within amorphous La2O3 layer supported on SiO2. The Ni@La2O3/SiO2 catalyst obtained exhibited high activity and low carbon deposition rate for the DRM reaction conducted at 700 °C and using 15% CH4, CH4/CO2 = 1. This method is a simple approach that can be widely applied for the preparation of encapsulated metal catalysts.

3. Materials and Methods

3.1. Synthesis of Catalysts

A Ni@La2O3/SiO2 catalyst with Ni encapsulated within amorphous La2O3 layer on SiO2 support was prepared via a one-pot colloidal solution combustion method. La(NO3)3·6H2O (2.39 g), Ni(NO3)2·6H2O (0.50 g), and glycine (0.60 g) were added in deionized water (6.30 mL). After 20 min of ultrasonic stirring, 1.26 mL of aqueous colloidal SiO2 LUDOX TMA (34 wt%, diameter of 22 nm; Sigma-Aldrich, St. Louis, MO, USA) was added to the solution. After 30 min of ultrasonic stirring, the solution was heated to 250 °C. After a few minutes of heating, glycine and nitrate began to react to form metal oxides and release a large amount of gas. In the combustion reaction, glycine and nitrate were used as the fuel and oxidizer, respectively. After the solid was formed, it was calcined at 700 °C for 4 h, and the Ni@La2O3/SiO2 catalyst was obtained. The weight contents of Ni and La2O3 in this catalyst were 6.7% and 60.0%, respectively. For comparison, a Ni–La2O3/SiO2 catalyst with the same Ni and La2O3 contents was prepared using the same method without adding glycine.

3.2. Characterization of Catalysts

N2 adsorption/desorption curves were obtained using an Autosorb-iQ analyzer (Quantachrome Instruments, Boynton Beach, FL, USA) at −196 °C, to quantify the specific surface area, pore size distribution/mean pore size, and pore volume of the catalysts. The crystal structure of the catalysts was determined by powder X-ray diffraction (XRD). The spectra were collected using a Rigaku-Miniflex 6 (Rigaku Corporation, Tokyo, Japan) powder X-ray diffractometer equipped with CuKα (λ = 0.15406 nm), between 20° and 80° (2θ) at a scanning speed of 10° min−1.
H2-temperature programmed reduction (H2-TPR) was applied on a TP-5080 multifunctional adsorption apparatus (Xianquan, Tianjin, China) in 5% H2/Ar gas mixture with a heating rate of 10 °C min−1. Transmission electron microscopy (TEM) was performed on a Tecnai G2 F20 microscope (FEI Company, Hillsboro, OR, USA), to directly observe the morphology and size of Ni particles and of deposited carbon after DRM. To determine the amount of carbon accumulation of the used catalysts, thermogravimetric (TG) and differential thermal analysis (DTA) were conducted on an HCT-1 TG thermal analyzer (Henven Scientific Instruments, Beijing, China).

3.3. Catalysts Performance Evaluation for the DRM Reaction

The catalytic performance of the Ni-based solids was evaluated at atmospheric pressure in a fixed-bed tubular quartz reactor (internal diameter 8 mm, length 300 mm). The total flow rate of reaction gases was 100 mL min−1 and the amount of catalyst used was in the 5–50 mg range. The corresponding WHSV was in the 120,000–1,200,000 mL g−1 h−1 range. CH4, CO2, H2, and Ar (99.999% purity) were purchased from Shanghai Maytor special Gas Co. Ltd. These gases contained less than 1 ppm of water vapor and were used without further purification. Before reaction, the catalyst was in-situ reduced in 20% H2/Ar gas mixture at 700 °C for 1.5 h. The reaction gases of CO2, CH4, and Ar at a molar ratio of 15/15/70 (vol%), were introduced into the reactor at 700 °C, and Ar gas was used as the internal standard. The gas effluent was analyzed using two on-line gas chromatographs of G5 (Beijing Purkinje General Instrument Co., Ltd., Beijing, China). One chromatograph used hydrogen as a carrier gas to detect Ar, CO, CH4, and CO2. Another chromatograph used N2 as a carrier gas to detect H2.
The CO2 conversion (XCO2), CH4 conversion (XCH4) and the H2/CO gas product ratio were calculated based on the following Equations (1)–(3):
X C O 2 = [ C O 2 ] i n [ C O 2 ] o u t [ C O 2 ] i n × 100 %  
X C H 4 = [ C H 4 ] i n [ C H 4 ] o u t [ C H 4 ] i n × 100 %  
H 2 / C O = [ H 2 ] o u t [ C O ] o u t
where [x]in and [x]out represent the mole fraction of x gaseous species in the inlet feed and outlet from reactor gas mixture, respectively.
The H2 and CO product yields were calculated based on the following Equations (4) and (5):
Y H 2 = [ H 2 ] o u t 2 [ C H 4 ] i n × 100 %  
Y C O = [ C O ] o u t [ C H 4 ] i n + [ C O 2 ] i n × 100 %  

4. Conclusions

In this study, we prepared a Ni@La2O3/SiO2 catalyst with encapsulated Ni nanoparticles via the colloidal solution combustion method tested for the DRM reaction at 700 °C. In the Ni@La2O3/SiO2 catalyst, small Ni particles were encapsulated within an amorphous La2O3 layer, where this was coated on SiO2. Due to the encapsulated Ni micro-structure, more interface between Ni and La2O3 was formed, and the Ni@La2O3/SiO2 catalyst exhibited excellent activity and stability and strong resistance to carbon deposition during DRM reaction. The catalytic performance results indicated that the CH4 conversion rate of the Ni@La2O3/SiO2 catalyst was five times higher than that of Ni–La2O3/SiO2 catalyst. More importantly, the Ni@La2O3/SiO2 catalyst exhibited excellent catalytic stability and only a slight deactivation for 80 h on reaction stream. TG-DTA studies revealed that 1.6 wt% carbon was deposited on the Ni@La2O3/SiO2 catalyst after 50 h of DRM, which was much lower than that of 11.5 wt% obtained on the Ni–La2O3/SiO2 catalyst.

Author Contributions

Conceptualization, L.W.; data curation, L.W. and R.H.; formal analysis, L.W., H.L., D.G. and L.M.; funding acquisition, L.W. and H.L.; investigation, L.W., R.H. and Z.Z.; writing—original draft, L.W. and R.H.; writing—review and editing, H.L., Q.W., D.G. and H.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Fundamental Research Funds for Zhejiang Provincial Universities and Research Institutes (No. 2019JZ00003), the Science and Technology Foundation of Zhoushan (No. 2018C21013), the Scientific Research Project of Zhejiang Education Department (No. Y201636192), and the university students’ innovation program of Zhejiang province (No. 2018R411009).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Khodakov, A.Y.; Chu, W.; Fongarland, P. Advances in the Development of Novel Cobalt Fischer–Tropsch Catalysts for Synthesis of Long-Chain Hydrocarbons and Clean Fuels. Chem. Rev. 2007, 107, 1692–1744. [Google Scholar] [CrossRef]
  2. Abdulrasheed, A.; Jalil, A.A.; Gambo, Y.; Ibrahim, M.; Hambali, H.U.; Shahul Hamid, M.Y. A review on catalyst development for dry reforming of methane to syngas: Recent advances. Renew. Sustain. Energy Rev. 2019, 108, 175–193. [Google Scholar] [CrossRef]
  3. Jang, W.-J.; Shim, J.-O.; Kim, H.-M.; Yoo, S.-Y.; Roh, H.-S. A review on dry reforming of methane in aspect of catalytic properties. Catal. Today 2019, 324, 15–26. [Google Scholar] [CrossRef]
  4. Aramouni, N.A.K.; Touma, J.G.; Tarboush, B.A.; Zeaiter, J.; Ahmad, M.N. Catalyst design for dry reforming of methane: Analysis review. Renew. Sustain. Energy Rev. 2018, 82, 2570–2585. [Google Scholar] [CrossRef]
  5. Ghelamallah, M.; Granger, P. Impact of barium and lanthanum incorporation to supported Pt and Rh on α-Al2O3 in the dry reforming of methane. Fuel 2012, 97, 269–276. [Google Scholar] [CrossRef]
  6. De Miguel, S.R.; Vilella, I.M.J.; Maina, S.P.; San José-Alonso, D.; Román-Martínez, M.C.; Illán-Gómez, M.J. Influence of Pt addition to Ni catalysts on the catalytic performance for long term dry reforming of methane. Appl. Catal. A Gen. 2012, 435, 10–18. [Google Scholar] [CrossRef]
  7. Crisafulli, C.; Scirè, S.; Maggiore, R.; Minicò, S.; Galvagno, S. CO2 reforming of methane over Ni–Ru and Ni–Pd bimetallic catalysts. Catal. Lett. 1999, 59, 21–26. [Google Scholar] [CrossRef]
  8. Wang, Y.; Yao, L.; Wang, S.; Mao, D.; Hu, C. Low-temperature catalytic CO2 dry reforming of methane on Ni-based catalysts: A review. Fuel Process. Technol. 2018, 169, 199–206. [Google Scholar] [CrossRef]
  9. Abdullah, B.; Abd Ghani, N.A.; Vo, D.-V.N. Recent advances in dry reforming of methane over Ni-based catalysts. J. Clean. Prod. 2017, 162, 170–185. [Google Scholar] [CrossRef] [Green Version]
  10. Zhang, G.; Liu, J.; Xu, Y.; Sun, Y. A review of CH4–CO2 reforming to synthesis gas over Ni-based catalysts in recent years (2010–2017). Int. J. Hydrog. Energy 2018, 43, 15030–15054. [Google Scholar] [CrossRef]
  11. Xie, T.; Shi, L.; Zhang, J.; Zhang, D. Immobilizing Ni nanoparticles to mesoporous silica with size and location control via a polyol-assisted route for coking- and sintering-resistant dry reforming of methane. Chem. Commun. 2014, 50, 7250–7253. [Google Scholar] [CrossRef] [PubMed]
  12. Zhang, Q.; Zhang, T.; Shi, Y.; Zhao, B.; Wang, M.; Liu, Q.; Wang, J.; Long, K.; Duan, Y.; Ning, P. A sintering and carbon-resistant Ni-SBA-15 catalyst prepared by solid-state grinding method for dry reforming of methane. J. CO2 Util. 2017, 17, 10–19. [Google Scholar] [CrossRef]
  13. Djinović, P.; Osojnik Črnivec, I.G.; Erjavec, B.; Pintar, A. Influence of active metal loading and oxygen mobility on coke-free dry reforming of Ni–Co bimetallic catalysts. Appl. Catal. B Environ. 2012, 125, 259–270. [Google Scholar] [CrossRef]
  14. Yan, X.; Hu, T.; Liu, P.; Li, S.; Zhao, B.; Zhang, Q.; Jiao, W.; Chen, S.; Wang, P.; Lu, J.; et al. Highly efficient and stable Ni/CeO2-SiO2 catalyst for dry reforming of methane: Effect of interfacial structure of Ni/CeO2 on SiO2. Appl. Catal. B Environ. 2019, 246, 221–231. [Google Scholar] [CrossRef]
  15. Luisetto, I.; Tuti, S.; Battocchio, C.; Lo Mastro, S.; Sodo, A. Ni/CeO2–Al2O3 catalysts for the dry reforming of methane: The effect of CeAlO3 content and nickel crystallite size on catalytic activity and coke resistance. Appl. Catal. A Gen. 2015, 500, 12–22. [Google Scholar] [CrossRef]
  16. Ay, H.; Üner, D. Dry reforming of methane over CeO2 supported Ni, Co and Ni–Co catalysts. Appl. Catal. B Environ. 2015, 179, 128–138. [Google Scholar] [CrossRef]
  17. Gould, T.D.; Izar, A.; Weimer, A.W.; Falconer, J.L.; Medlin, J.W. Stabilizing Ni Catalysts by Molecular Layer Deposition for Harsh, Dry Reforming Conditions. ACS Catal. 2014, 4, 2714–2717. [Google Scholar] [CrossRef]
  18. Kim, J.-H.; Suh, D.J.; Park, T.-J.; Kim, K.-L. Effect of metal particle size on coking during CO2 reforming of CH4 over Ni–alumina aerogel catalysts. Appl. Catal. A Gen. 2000, 197, 191–200. [Google Scholar] [CrossRef]
  19. Abdel Karim Aramouni, N.; Zeaiter, J.; Kwapinski, W.; Ahmad, M.N. Thermodynamic analysis of methane dry reforming: Effect of the catalyst particle size on carbon formation. Energy Convers. Manag. 2017, 150, 614–622. [Google Scholar] [CrossRef]
  20. Han, J.W.; Park, J.S.; Choi, M.S.; Lee, H. Uncoupling the size and support effects of Ni catalysts for dry reforming of methane. Appl. Catal. B Environ. 2017, 203, 625–632. [Google Scholar] [CrossRef]
  21. Zhang, Q.; Tang, T.; Wang, J.; Sun, M.; Wang, H.; Sun, H.; Ning, P. Facile template-free synthesis of Ni-SiO2 catalyst with excellent sintering- and coking-resistance for dry reforming of methane. Catal. Commun. 2019, 131, 105782. [Google Scholar] [CrossRef]
  22. Kawi, S.; Kathiraser, Y.; Ni, J.; Oemar, U.; Li, Z.; Saw, E.T. Progress in Synthesis of Highly Active and Stable Nickel-Based Catalysts for Carbon Dioxide Reforming of Methane. ChemSusChem 2015, 8, 3556–3575. [Google Scholar] [CrossRef] [PubMed]
  23. Tian, H.; Li, X.; Zeng, L.; Gong, J. Recent Advances on the Design of Group VIII Base-Metal Catalysts with Encapsulated Structures. ACS Catal. 2015, 5, 4959–4977. [Google Scholar] [CrossRef]
  24. Liu, W.; Li, L.; Zhang, X.; Wang, Z.; Wang, X.; Peng, H. Design of Ni-ZrO2@SiO2 catalyst with ultra-high sintering and coking resistance for dry reforming of methane to prepare syngas. J. CO2 Util. 2018, 27, 297–307. [Google Scholar] [CrossRef]
  25. Li, Z.; Jiang, B.; Wang, Z.; Kawi, S. High carbon resistant Ni@Ni phyllosilicate@SiO2 core shell hollow sphere catalysts for low temperature CH4 dry reforming. J. CO2 Util. 2018, 27, 238–246. [Google Scholar] [CrossRef]
  26. Zhang, L.; Wang, F.; Zhu, J.; Han, B.; Fan, W.; Zhao, L.; Cai, W.; Li, Z.; Xu, L.; Yu, H.; et al. CO2 reforming with methane reaction over Ni@SiO2 catalysts coupled by size effect and metal-support interaction. Fuel 2019, 256, 115954. [Google Scholar] [CrossRef]
  27. Zhao, X.; Li, H.; Zhang, J.; Shi, L.; Zhang, D. Design and synthesis of NiCe@m-SiO2 yolk-shell framework catalysts with improved coke- and sintering-resistance in dry reforming of methane. Int. J. Hydrog. Energy 2016, 41, 2447–2456. [Google Scholar] [CrossRef]
  28. Wang, F.; Han, B.; Zhang, L.; Xu, L.; Yu, H.; Shi, W. CO2 reforming with methane over small-sized Ni@SiO2 catalysts with unique features of sintering-free and low carbon. Appl. Catal. B Environ. 2018, 235, 26–35. [Google Scholar] [CrossRef]
  29. Wang, C.; Jie, X.; Qiu, Y.; Zhao, Y.; Al-Megren, H.A.; Alshihri, S.; Edwards, P.P.; Xiao, T. The importance of inner cavity space within Ni@SiO2 nanocapsule catalysts for excellent coking resistance in the high-space-velocity dry reforming of methane. Appl. Catal. B Environ. 2019, 259, 118019. [Google Scholar] [CrossRef]
  30. Zhang, J.; Li, F. Coke-resistant Ni@SiO2 catalyst for dry reforming of methane. Appl. Catal. B Environ. 2015, 176, 513–521. [Google Scholar] [CrossRef]
  31. Das, S.; Ashok, J.; Bian, Z.; Dewangan, N.; Wai, M.H.; Du, Y.; Borgna, A.; Hidajat, K.; Kawi, S. Silica–Ceria sandwiched Ni core–shell catalyst for low temperature dry reforming of biogas: Coke resistance and mechanistic insights. Appl. Catal. B Environ. 2018, 230, 220–236. [Google Scholar] [CrossRef]
  32. Dou, J.; Zhang, R.; Hao, X.; Bao, Z.; Wu, T.; Wang, B.; Yu, F. Sandwiched SiO2@Ni@ZrO2 as a coke resistant nanocatalyst for dry reforming of methane. Appl. Catal. B Environ. 2019, 254, 612–623. [Google Scholar] [CrossRef]
  33. Li, K.; He, F.; Yu, H.; Wang, Y.; Wu, Z. Theoretical study on the reaction mechanism of carbon dioxide reforming of methane on La and La2O3 modified Ni (1 1 1) surface. J. Catal. 2018, 364, 248–261. [Google Scholar] [CrossRef]
  34. Sierra Gallego, G.; Batiot-Dupeyrat, C.; Barrault, J.; Mondragón, F. Dual Active-Site Mechanism for Dry Methane Reforming over Ni/La2O3 Produced from LaNiO3 Perovskite. Ind. Eng. Chem. Res. 2008, 47, 9272–9278. [Google Scholar] [CrossRef]
  35. Zhang, Z.; Verykios, X.E.; MacDonald, S.M.; Affrossman, S. Comparative Study of Carbon Dioxide Reforming of Methane to Synthesis Gas over Ni/La2O3 and Conventional Nickel-Based Catalysts. J. Phys. Chem. 1996, 100, 744–754. [Google Scholar] [CrossRef]
  36. Charisiou, N.D.; Tzounis, L.; Sebastian, V.; Hinder, S.J.; Baker, M.A.; Polychronopoulou, K.; Goula, M.A. Investigating the correlation between deactivation and the carbon deposited on the surface of Ni/Al2O3 and Ni/La2O3-Al2O3 catalysts during the biogas reforming reaction. Appl. Surf. Sci. 2019, 474, 42–56. [Google Scholar] [CrossRef] [Green Version]
  37. Mo, W.; Ma, F.; Ma, Y.; Fan, X. The optimization of Ni–Al2O3 catalyst with the addition of La2O3 for CO2–CH4 reforming to produce syngas. Int. J. Hydrog. Energy 2019, 44, 24510–24524. [Google Scholar] [CrossRef]
  38. Chen, C.; Wang, X.; Zhang, L.; Zou, X.; Ding, W.; Lu, X. Synthesis of mesoporous Ni–La2O3/SiO2 by ploy(ethylene glycol)-assisted sol-gel route as highly efficient catalysts for dry reforming of methane with a H2/CO ratio of unity. Catal. Commun. 2017, 94, 38–41. [Google Scholar] [CrossRef]
  39. Cheng, H.; Li, G.; Zhao, H.; Lu, X.; Xu, Q.; Tao, W. Effects of preparation technique and lanthana doping on Ni/La2O3-ZrO2 catalysts for hydrogen production by CO2 reforming of coke oven gas. Catal. Today 2018, 318, 23–31. [Google Scholar] [CrossRef]
  40. Li, X.; Li, D.; Tian, H.; Zeng, L.; Zhao, Z.-J.; Gong, J. Dry reforming of methane over Ni/La2O3 nanorod catalysts with stabilized Ni nanoparticles. Appl. Catal. B Environ. 2017, 202, 683–694. [Google Scholar] [CrossRef]
  41. Xu, L.; Liu, W.; Zhang, X.; Tao, L.; Xia, L.; Xu, X.; Song, J.; Zhou, W.; Fang, X.; Wang, X. Ni/La2O3 Catalysts for Dry Reforming of Methane: Insights into the Factors Improving the Catalytic Performance. ChemCatChem 2019, 11, 2887–2899. [Google Scholar] [CrossRef]
  42. Li, K.; Chang, X.; Pei, C.; Li, X.; Chen, S.; Zhang, X.; Assabumrungrat, S.; Zhao, Z.-J.; Zeng, L.; Gong, J. Ordered mesoporous Ni/La2O3 catalysts with interfacial synergism towards CO2 activation in dry reforming of methane. Appl. Catal. B Environ. 2019, 259, 118092. [Google Scholar] [CrossRef]
  43. Voskanyan, A.A.; Chan, K.-Y.; Li, C.-Y.V. Colloidal Solution Combustion Synthesis: Toward Mass Production of a Crystalline Uniform Mesoporous CeO2 Catalyst with Tunable Porosity. Chem. Mater. 2016, 28, 2768–2775. [Google Scholar]
  44. Tang, G.; Gong, D.; Liu, H.; Wang, L. Highly Loaded Mesoporous Ni–La2O3 Catalyst Prepared by Colloidal Solution Combustion Method for CO2 Methanation. Catalysts 2019, 9, 442. [Google Scholar] [CrossRef] [Green Version]
  45. 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] [Green Version]
  46. Zhao, B.; Yao, Y.; Shi, H.; Yang, F.; Jia, X.; Liu, P.; Ma, X. Preparation of Ni/SiO2 catalyst via novel plasma-induced micro-combustion method. Catal. Today 2019, 337, 28–36. [Google Scholar] [CrossRef]
  47. Wang, M.; Zhang, Q.; Zhang, T.; Wang, Y.; Wang, J.; Long, K.; Song, Z.; Liu, X.; Ning, P. Facile one-pot synthesis of highly dispersed Ni nanoparticles embedded in HMS for dry reforming of methane. Chem. Eng. J. 2017, 313, 1370–1381. [Google Scholar] [CrossRef]
  48. Garbarino, G.; Wang, C.; Valsamakis, I.; Chitsazan, S.; Riani, P.; Finocchio, E.; Flytzani-Stephanopoulos, M.; Busca, G. A study of Ni/Al2O3 and Ni–La/Al2O3 catalysts for the steam reforming of ethanol and phenol. Appl. Catal. B Environ. 2015, 174–175, 21–34. [Google Scholar] [CrossRef] [Green Version]
  49. Xu, Y.; Lin, Q.; Liu, B.; Jiang, F.; Xu, Y.; Liu, X. A Facile Fabrication of Supported Ni/SiO2 Catalysts for Dry Reforming of Methane with Remarkably Enhanced Catalytic Performance. Catalysts 2019, 9, 183. [Google Scholar] [CrossRef] [Green Version]
  50. Osaki, T.; Mori, T. Role of Potassium in Carbon-Free CO2 Reforming of Methane on K-Promoted Ni/Al2O3 Catalysts. J. Catal. 2001, 204, 89–97. [Google Scholar] [CrossRef]
  51. Le Saché, E.; Pastor-Pérez, L.; Watson, D.; Sepúlveda-Escribano, A.; Reina, T.R. Ni stabilised on inorganic complex structures: Superior catalysts for chemical CO2 recycling via dry reforming of methane. Appl. Catal. B Environ. 2018, 236, 458–465. [Google Scholar] [CrossRef]
  52. Vasiliades, M.A.; Damaskinos, C.M.; Kyprianou, K.K.; Kollia, M.; Efstathiou, A.M. The effect of Pt on the carbon pathways in the dry reforming of methane over Ni-Pt/Ce0.8Pr0.2O2-δ catalyst. Catal. Today 2019, in press. [Google Scholar] [CrossRef]
  53. Chai, Y.; Fu, Y.; Feng, H.; Kong, W.; Yuan, C.; Pan, B.; Zhang, J.; Sun, Y. A Nickel-Based Perovskite Catalyst with a Bimodal Size Distribution of Nickel Particles for Dry Reforming of Methane. ChemCatChem 2018, 10, 2078–2086. [Google Scholar] [CrossRef]
  54. Vasiliades, M.A.; Djinović, P.; Davlyatova, L.F.; Pintar, A.; Efstathiou, A.M. Origin and reactivity of active and inactive carbon formed during DRM over Ni/Ce0.38Zr0.62O2-δ studied by transient isotopic techniques. Catal. Today 2018, 299, 201–211. [Google Scholar] [CrossRef]
  55. Torres, D.; Pinilla, J.L.; Suelves, I. Co-, Cu- and Fe-Doped Ni/Al2O3 Catalysts for the Catalytic Decomposition of Methane into Hydrogen and Carbon Nanofibers. Catalysts 2018, 8, 300. [Google Scholar] [CrossRef] [Green Version]
  56. Torres, D.; Pinilla, J.L.; Suelves, I. Screening of Ni-Cu bimetallic catalysts for hydrogen and carbon nanofilaments production via catalytic decomposition of methane. Appl. Catal. A Gen. 2018, 559, 10–19. [Google Scholar] [CrossRef]
  57. Vasiliades, M.A.; Makri, M.M.; Djinović, P.; Erjavec, B.; Pintar, A.; Efstathiou, A.M. Dry reforming of methane over 5 wt% Ni/Ce1-xPrxO2-δ catalysts: Performance and characterisation of active and inactive carbon by transient isotopic techniques. Appl. Catal. B Environ. 2016, 197, 168–183. [Google Scholar] [CrossRef]
  58. Cao, A.; Lu, R.; Veser, G. Stabilizing metal nanoparticles for heterogeneous catalysis. Phys. Chem. Chem. Phys. 2010, 12, 13499–13510. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of Ni@La2O3/SiO2 and Ni–La2O3/SiO2 solid catalysts. Reduced catalysts were treated with 20% H2/Ar at 700 °C for 1.5 h before XRD analysis.
Figure 1. XRD patterns of Ni@La2O3/SiO2 and Ni–La2O3/SiO2 solid catalysts. Reduced catalysts were treated with 20% H2/Ar at 700 °C for 1.5 h before XRD analysis.
Catalysts 10 00038 g001
Figure 2. TEM images of Ni@La2O3/SiO2 and Ni–La2O3/SiO2 catalysts: (a,b) fresh Ni–La2O3/SiO2, (c,d) fresh Ni@La2O3/SiO2, and (e,f) reduced Ni@La2O3/SiO2.
Figure 2. TEM images of Ni@La2O3/SiO2 and Ni–La2O3/SiO2 catalysts: (a,b) fresh Ni–La2O3/SiO2, (c,d) fresh Ni@La2O3/SiO2, and (e,f) reduced Ni@La2O3/SiO2.
Catalysts 10 00038 g002
Figure 3. STEM -EDS mapping images of Si, La, and Ni images of Ni@La2O3/SiO2 after reduction.
Figure 3. STEM -EDS mapping images of Si, La, and Ni images of Ni@La2O3/SiO2 after reduction.
Catalysts 10 00038 g003
Figure 4. H2-TPR traces of Ni@La2O3/SiO2 and Ni–La2O3/SiO2 catalysts.
Figure 4. H2-TPR traces of Ni@La2O3/SiO2 and Ni–La2O3/SiO2 catalysts.
Catalysts 10 00038 g004
Figure 5. Stability tests performed over the Ni@La2O3/SiO2 and Ni–La2O3/SiO2 catalysts in the dry reforming of methane (DRM). (a) CO2 and CH4 conversions, (b) H2/CO ratio, (c) CO yield, and (d) H2 yield. Reaction conditions: Total P = 1 atm, CH4/CO2/Ar = 15/15/70 (vol%), T = 700 °C, weight hourly space velocity (WHSV) = 120,000 mL g−1 h−1.
Figure 5. Stability tests performed over the Ni@La2O3/SiO2 and Ni–La2O3/SiO2 catalysts in the dry reforming of methane (DRM). (a) CO2 and CH4 conversions, (b) H2/CO ratio, (c) CO yield, and (d) H2 yield. Reaction conditions: Total P = 1 atm, CH4/CO2/Ar = 15/15/70 (vol%), T = 700 °C, weight hourly space velocity (WHSV) = 120,000 mL g−1 h−1.
Catalysts 10 00038 g005
Figure 6. Effect of space velocity on the catalytic performance of Ni@La2O3/SiO2 catalyst. Reaction conditions: P = 1 atm, CH4/CO2/Ar = 15/15/70 (vol%), T = 700 °C.
Figure 6. Effect of space velocity on the catalytic performance of Ni@La2O3/SiO2 catalyst. Reaction conditions: P = 1 atm, CH4/CO2/Ar = 15/15/70 (vol%), T = 700 °C.
Catalysts 10 00038 g006
Figure 7. Stability test conducted over the Ni@La2O3/SiO2 catalyst for 80 h at high space velocity. Reaction conditions: P = 1 atm, CH4/CO2/Ar = 15/15/70 (vol%), T = 700 °C, WHSV = 600,000 mL g−1 h−1.
Figure 7. Stability test conducted over the Ni@La2O3/SiO2 catalyst for 80 h at high space velocity. Reaction conditions: P = 1 atm, CH4/CO2/Ar = 15/15/70 (vol%), T = 700 °C, WHSV = 600,000 mL g−1 h−1.
Catalysts 10 00038 g007
Figure 8. (a) thermogravimetric (TG) and (b) differential thermal analysis (DTA) curves of the used Ni@La2O3/SiO2 and Ni–La2O3/SiO2 catalysts after 50 h of DRM at 700 °C with a WHSV of 120,000 mL g−1 h−1.
Figure 8. (a) thermogravimetric (TG) and (b) differential thermal analysis (DTA) curves of the used Ni@La2O3/SiO2 and Ni–La2O3/SiO2 catalysts after 50 h of DRM at 700 °C with a WHSV of 120,000 mL g−1 h−1.
Catalysts 10 00038 g008
Figure 9. XRD patterns of the used Ni@La2O3/SiO2 and Ni–La2O3/SiO2 catalysts after 50 h of DRM at 700 °C with a WHSV of 120,000 mL g−1 h−1.
Figure 9. XRD patterns of the used Ni@La2O3/SiO2 and Ni–La2O3/SiO2 catalysts after 50 h of DRM at 700 °C with a WHSV of 120,000 mL g−1 h−1.
Catalysts 10 00038 g009
Figure 10. TEM images of the used (a,b) Ni@La2O3/SiO2 and (c,d) Ni–La2O3/SiO2 catalysts after 50 h of DRM at 700 °C with a WHSV of 120,000 mL g−1 h−1.
Figure 10. TEM images of the used (a,b) Ni@La2O3/SiO2 and (c,d) Ni–La2O3/SiO2 catalysts after 50 h of DRM at 700 °C with a WHSV of 120,000 mL g−1 h−1.
Catalysts 10 00038 g010
Table 1. Physical properties of Ni@La2O3/SiO2 and Ni–La2O3/SiO2.
Table 1. Physical properties of Ni@La2O3/SiO2 and Ni–La2O3/SiO2.
SamplesSBET
(m2 g−1)
Pore Volume (cm3 g−1)Average Pore Size (nm)dNiO a (nm)dNi b (nm)
By XRD c By TEM dBy XRD c By TEM d
Ni@La2O3/SiO219.00.2143.9N. D. eN. D. eN. D. e3.5 (5)
Ni–La2O3/SiO224.10.1728.68.67.816.4 (44.6)(26.7)
a NiO size in the fresh catalyst. b Ni size in the reduced and used catalysts. Data in brackets correspond to the Ni size of used catalyst. c Crystallite size determined by XRD using the Scherrer equation. d Mean particle size determined by TEM images analysis. e The particle in TEM image was too small to be observed, or the peak in XRD pattern was too weak to be used for calculations.
Table 2. Methane conversion and carbon deposition rates of Ni@La2O3/SiO2, Ni–La2O3/SiO2, and recently reported catalysts for the DRM reaction at 700 °C.
Table 2. Methane conversion and carbon deposition rates of Ni@La2O3/SiO2, Ni–La2O3/SiO2, and recently reported catalysts for the DRM reaction at 700 °C.
Ni (wt %)WHSV
(mL gcat−1 h−1)
XCH4 (%)CH4 Conversion Rate
mmol CH4 gNi−1 s−1
Carbon Deposition Rates
mg C gNi−1 h−1
Ref.
Ni@La2O3/SiO26.31,200,0003512.40 a /This Work
Ni@La2O3/SiO26.3600,0006110.10 b 0.33 dThis Work
Ni–La2O3/SiO26.3120,000491.74 c 2.60This Work
Ni/La2O3-LOC5.7300,00023.62.31 2.30[40]
SiO2@Ni@ZrO28.9180,00043.13.60 Not detected[32]
Ni/CeO2-SiO2548,00078.52.34 /[14]
Ni@SiO23.618,000752.33 /[28]
Ni/LaZrxOy12.760,000550.640.56[51]
a The rate was calculated after reaction for 30 h. b The rate was calculated after reaction for 80 h. c The rate was calculated after reaction for 50 h. d WHSV = 120,000 mL gcat−1 h−1.

Share and Cite

MDPI and ACS Style

Wang, L.; Hu, R.; Liu, H.; Wei, Q.; Gong, D.; Mo, L.; Tao, H.; Zhang, Z. Encapsulated Ni@La2O3/SiO2 Catalyst with a One-Pot Method for the Dry Reforming of Methane. Catalysts 2020, 10, 38. https://doi.org/10.3390/catal10010038

AMA Style

Wang L, Hu R, Liu H, Wei Q, Gong D, Mo L, Tao H, Zhang Z. Encapsulated Ni@La2O3/SiO2 Catalyst with a One-Pot Method for the Dry Reforming of Methane. Catalysts. 2020; 10(1):38. https://doi.org/10.3390/catal10010038

Chicago/Turabian Style

Wang, Luhui, Rong Hu, Hui Liu, Qinhong Wei, Dandan Gong, Liuye Mo, Hengcong Tao, and Zhonghuai Zhang. 2020. "Encapsulated Ni@La2O3/SiO2 Catalyst with a One-Pot Method for the Dry Reforming of Methane" Catalysts 10, no. 1: 38. https://doi.org/10.3390/catal10010038

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