Next Article in Journal
Appraisal of Sulphonation Processes to Synthesize Palm Waste Biochar Catalysts for the Esterification of Palm Fatty Acid Distillate
Next Article in Special Issue
Combined Magnesia, Ceria and Nickel catalyst supported over γ-Alumina Doped with Titania for Dry Reforming of Methane
Previous Article in Journal
Experimental Studies on Co-Combustion of Sludge and Wheat Straw
Previous Article in Special Issue
In Situ Regeneration of Alumina-Supported Cobalt–Iron Catalysts for Hydrogen Production by Catalytic Methane Decomposition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Facile Fabrication of Supported Ni/SiO2 Catalysts for Dry Reforming of Methane with Remarkably Enhanced Catalytic Performance

1
Department of Chemical Engineering, School of Chemical and Material Engineering, Jiangnan University, Wuxi 214122, China
2
School of Chemistry and Chemical Engineering, Xuzhou University of Technology, Xuzhou 221018, China
*
Author to whom correspondence should be addressed.
Catalysts 2019, 9(2), 183; https://doi.org/10.3390/catal9020183
Submission received: 12 January 2019 / Revised: 9 February 2019 / Accepted: 13 February 2019 / Published: 15 February 2019
(This article belongs to the Special Issue Catalysts for Syngas Production)

Abstract

:
Ni catalysts supported on SiO2 are prepared via a facile combustion method. Both glycine fuel and ammonium nitrate combustion improver facilitate the formation of much smaller Ni nanoparticles, which give excellent activity and stability, as well as a syngas with a molar ratio of H2/CO of about 1:1 due to the minimal side reaction toward revserse water gas shift (RWGS) in CH4 dry reforming.

Graphical Abstract

1. Introduction

The availability of natural gas (or shale gas) in large reserves makes CH4 serve as a suitable feedstock used in C1 chemistry to produce desired fuels and chemicals [1]. Unfortunately, the chemical inertness of CH4 results in direct conversion, which constitutes a great challenge for highly efficient utilization [2]. Ideally, the best use of CH4 occurs when it is converted into syngas, which can facilitate further downstream conversion [3] by means of the methanol route [4] and Fischer–Tropsch synthesis (FTS) [5,6,7,8,9,10,11,12] due to good reactivity, unlike the CH4 which has a high dissociation energy C–H bond [1]. Among the most widely investigated technologies, there are comparable advantages associated with the dry reforming of CH4 (DRM) with CO2 for producing syngas [13]. On the one hand, compared to the other reforming processes, there is a 20% lower operating cost for DRM [14]; on the other hand, the reforming of CH4 using CO2 not only produces high purity syngas [15,16] but also reduces the emissions of two abundantly available greenhouse gases to alleviate global climate change [17,18,19,20].
In spite of the above-mentioned merits, DRM suffers from serious carbon deposits on the surface of Ni nanoparticles, which leads to a remarkable loss of active sites [21,22,23,24,25]. Recently, DRM research efforts have resulted in strategies to improve the stability of the catalyst [26]. Based on the fact that smaller Ni nanoparticles efficiently improve catalytic performance by avoiding carbon accumulation [27,28,29,30,31,32], the general concept is to develop the catalyst preparation protocol to obtain small Ni nanoparticles encapsulated in the support or confined by the stable porous oxide layer to prevent sintering [33,34]. For example, Tomishige et al. reported that the solid solution catalyst of nickel–magnesia, which was prepared by the co-precipitation method, showed high and stable activity without carbon deposits for 100 days [35,36]. Kawi et al. synthesized a Ni-yolk@Ni@SiO2 nanocomposite with a yolk-satellite shell structure to efficiently inhibit the sintering of Ni, which resulted in negligible carbon deposition, and the CH4 conversion was 10% after the first 2 hours of reaction under the conditions of 800 °C, a gas hourly space velocity (GHSV) of 1440 L·g−1cat·h−1, a Wcat of 0.01 g, and a CO2:CH4:N2 ratio of 1:1:1 [37]. Similarly, Wang et al. pointed out that the Ni nanoparticle cores encapsulated by the mesoporous Al2O3 shells show superior coke resistance because of the confinement effects which prevent the Ni nanoparticles from agglomeration at high temperatures, and the CH4 and CO2 conversions under the reaction conditions of 800 °C, CO2/CH4 of 1/1, and a weight hourly space velocity (WHSV) of 36 L·h−1·gcat−1 were about 88% and 92%, respectively [38].
Herein, different from the above-mentioned encapsulated Ni catalysts with relatively complicated preparation procedures, we propose a facile one-step strategy to prepare the SiO2 supported Ni catalysts toward the controlled formation of nanoparticle size and Ni-support interaction, which could lead to high activity and stability. Following the conventional impregnation method, glycine (C2H5NO2) and ammonium nitrate (NH4NO3) were introduced into the impregnated solution of nickel precursor (Ni(NO3)2·6H2O), as shown in Scheme 1. It was expected that the mixed materials with C2H5NO2 as fuel and NH4NO3 as combustion improver reacted exothermically after ignition which finished within a short time-frame with a very high temperature and release of a large quantity of gases, such as CO2, water, and N2. We thought this process might facilitate the formation of smaller crystalline materials and regulate the metal-support interaction, resulting in improved catalytic performance in the DRM reaction. To demonstrate the effects of the above combustion process on the catalytic performance, several characterizations, such as Brunauer-Emmett-Teller (BET), transmission electron microscopy (TEM), X-ray diffraction (XRD), H2 temperature-programmed reduction (H2-TPR) and thermogravimetric (TG), were employed to characterize the catalyst.

2. Results and Discussion

2.1. Characterization of the Catalyst Sample

As shown in Figure S1, all the fresh Ni/SiO2 catalysts exhibit apparent diffraction peaks at 2θ values of 37.3°, 43.2°, 63.0°, 75.4°, and 79.4° assigned to the NiO (JCPDS 22-1189). For the reduced catalysts (Figure 1a), Ni/SiO2-0/0 prepared by the conventional wetness impregnation method displayed the most intensive diffraction peaks at 2θ values of 44.5°, 52.2°, and 77.0°, which are the characteristic peaks of metallic Ni (JCPDS 1-1206). According to Figure S1, the peak at 37.3° should be assigned to NiO. As the NH4NO3 was introduced into the impregnated solution with nickel nitrate, the resulting catalyst (Ni/SiO2-0/1) exhibited almost the same diffraction peak intensity at 44.5°. However, for the case of C2H5NO2, Ni/SiO2-2/0 displays a much weaker diffraction peak. Interestingly, the addition of both C2H5NO2 and NH4NO3 results in almost no detectable diffraction peaks for Ni nanoparticles (Ni/SiO2-2/1), suggesting that smaller Ni nanoparticles can be obtained by synergistic effects of fuel and combustion improver in the combustion process, as presented in Scheme 1.
TEM images of the reduced catalysts are depicted in Figure 1b,c. The Ni/SiO2-2/1 displays an average Ni nanoparticle size of only 6.1 ± 2.7 nm which is significantly smaller than that for Ni/SiO2-0/0 (31.3 ± 13.5 nm). The significant difference in the Ni nanoparticle size further confirms the synergistic effects of C2H5NO2 and NH4NO3 in reducing the Ni nanoparticle size. The combustion process between N2O and NH3 is highly exothermic. The decomposition of nickel nitrate produces N2O gas at 250 °C, while the decomposition of C2H5NO2 gives NH3 along with CO2 and H2O. The combustion process is triggered by the reaction between N2O and NH3 to form N2 and H2O [39]. When NH4NO3 is further added, NH3 and N2O can be formed via its decomposition at a low temperature of about 200 °C, thereby promoting combustion. The high-temperature stage in a short-duration favors the formation of ultra-small nanoparticles in a short time which may be in the order of seconds [40].
Figure 1d exhibits the reduction behavior of fresh Ni/SiO2 catalysts with different molar ratios of C2H5NO2 to NH4NO3. As expected, NH4NO3 does not obviously change the H2-TPR profile compared to the case of Ni/SiO2-0/0, as both catalysts show a strong reduction peak at 300–450 °C with a small right shoulder peak at 450–510 °C. However, C2H5NO2 only (Ni/SiO2-2/0) notably weakens the peak at lower temperatures, accompanied by a shift in the right shoulder peak to the higher reduction temperature with enhanced intensity. For Ni/SiO2-2/1, the high temperature reduction peak is further intensified and shifts to a higher reduction temperature range. This result suggests that the smaller Ni nanoparticle size results in a more difficult reduction owing to a stronger metal-support interaction [41]. The reduction profiles correspond to the XRD and TEM results.

2.2. Activity Evaluation

Figure 2 shows the catalytic performance in the CH4 dry reforming reaction with CO2 over the as-prepared catalysts. In the case of catalytic activity, the CH4 conversion over Ni/SiO2-0/0 exhibits a rapid drop from 78.3% to 53.0% in the early ten hours and then gradually becomes stable. In contrast, Ni/SiO2-0/1 gives a milder and continuous decrease in CH4 conversion until the end of the reaction. Surprisingly, Ni/SiO2-2/0 exhibits a stable and higher CH4 conversion over the whole reaction period of 50 hours. Furthermore, Ni/SiO2-2/1 displays a more stable and even higher CH4 conversion. The CO2 and CH4 conversions are similar for all Ni/SiO2 catalysts. However, in the corresponding reaction period, the CO2 conversion is always slightly higher compared to the CH4 conversion. When the DRM reaction over Ni/SiO2-2/1 is stable, the conversion rates of CH4 and CO2 are 83.6% and 90.6%, respectively, which are slightly lower than their equilibrium conversion rates at 91% and 95% calculated by HSC chemistry 6.0 (Table S1). Also, in the case of the H2/CO molar ratio, it follows the same trend as that for CH4 conversion over all the Ni/SiO2 catalysts. Specifically, for the Ni/SiO2-2/1 catalyst, the desired H2/CO molar ratio at the value of 1/1 is obtained, which results from the efficiently suppressed reverse water gas shift (RWGS) reaction. How the Ni/SiO2 morphology affects the catalytic performance is discussed briefly in the following part.
The DRM reaction is extremely endothermic. Equation (1) shows that the DRM process can produce a syngas with an H2/CO ratio of 1:1. During the DRM process, several reactions simultaneously occur, like CH4 dissociation (Equation (2)), reduction of CO2 to CO (Equation (3)), and the RWGS reaction (Equation (4)).
CH4 + CO2 = 2CO + 2H2 (ΔH298K = +247 kJ mol−1)
CH4 = C(s) + 2H2 (ΔH298K = +75 kJ mol−1)
C(s) + CO2 = 2CO (ΔH298K = +171 kJ mol−1)
CO2 + H2 = CO + H2O (ΔH298K = +41.2 kJ mol−1)
The driving force for Equations (2)–(4) strongly depends on the temperature, reactant partial pressure and catalyst structures. In the investigated Ni/SiO2 catalysts, both activation of CH4 and CO2 can occur on the active Ni surface since SiO2 support is inert material. It is believed that CH4 activation tends to form an intermediate, like CHx or a formyl group, but dissociates directly to C species and H2 at high temperature. Essentially, the DRM reaction of Ni catalysts might follow a dynamic redox type mechanism as the CO2 oxidizes Ni0 to Ni to give CO, and the oxidative state Ni is reduced to Ni0 by C species as a result of CH4 dissociation. As seen from the above reaction cycle, it is clear that the presence of O from CO2 helps the dissociation of CH4. To avoid the catalyst deactivation resulting from carbon accumulation, the C species from CH4 dissociation must react timely with CO2 to give CO. The reaction rate of this step is closely related to the Ni nanoparticle size, as the larger Ni surface favors the formation of multicarbon Cn species, which are potential precursors of carbon deposits such as coke. The smaller Ni nanoparticles allow a smaller amount of carbon species on the Ni nanoparticle surface. Thus, it is easier to keep the monoatomic C species isolated, and in time, they are oxidized by CO2 to CO. By minimizing the rate of C species combination, the carbon accumulation could be effectively suppressed. Indeed, as shown in Figure 3, Ni/SiO2-0/0, with an average nanoparticle size of 31.3 ± 13.5 nm, gives the highest amount of carbon deposits with 2.7 mg carbon deposits gCH4−1 as the BET surface area is decreased to the largest extent (Table S2). In contrast, the Ni/SiO2-2/1 with a smaller nanoparticle size of 6.1 ± 2.7 nm is significantly coke-resistant, as the amount of carbon deposits decreases to 0.9 mg carbon deposits gCH4−1. The above experimental results reflect that the smaller Ni nanoparticle size is favorable to lower carbon deposits and thereby improve the catalyst stability, as shown in Figure S2. It should be noted that, in spite of the significant decrease in carbon deposits over Ni/SiO2-2/1 catalyst, a considerable amount of coke is still formed during the DRM reaction of 50 hours. It can be deduced that most of the carbon deposits might not locate on the Ni nanoparticle surface but are located on the SiO2 support since the catalytic activity is quite stable. It is reasonable for us to imagine that the Ni nanoparticles are lying on the SiO2 support and not confined by porous layer material, which provides a chance for the carbon species to grow continuously along the SiO2 support surface initiated by the Ni nanoparticle and finally form strips of nanofiber.
As seen from Figure 2, the H2/CO molar ratio is highly dependent on the CO2 conversion. A lower CO2 conversion can cause a decrease in the molar ratio of H2/CO to a large extent as a result of the RWGS reaction, as the higher concentration of CO2 drives the reaction to the right side (Equation 4). At 800 °C, the standard free energy for the RWGS reaction (ΔG0 = –8545 + 7.84T) and the reduction of CO2 to CO (ΔG0 = 39810 − 40.87T) [13] is −132.68 kJ mol−1 and −4043.51 kJ mol−1, respectively. It can be speculated that the reduction of CO2 to CO, C(s) + CO2 = 2CO, occurs more easily as a result of the lower ΔG. Comparing the value of ΔG in the RWGS reaction, the CO2 that oxidizes the C species to CO is more thermodynamically favored than its RWGS reaction. As the lower CO2 conversion corresponds to lower CH4 conversion, the C(s) species dissociated from CH4 is not sufficient for its reaction with CO2. Therefore, the CO2 reacting with H2 toward the RWGS reaction is promoted. In order to minimize the side reaction toward the RWGS, it is necessary to operate the DRM reaction with a high CO2 conversion rate.

3. Materials and Methods

3.1. Catalyst Preparation

The supported Ni catalysts were prepared with SiO2 support (Tosoh Kabushiki-gaisha, Tokyo, Japan) by the combustion method, and the combustible materials contained hydrate glycine (C2H5NO2), ammonium nitrate (NH4NO3) and nickel nitrate (Ni(NO3)2·6H2O) with different C2H5NO2/Ni(NO3)2·6H2O and NH4NO3/Ni(NO3)2·6H2O molar ratios. Briefly, the aqueous solution of the desired amounts of C2H5NO2, NH4NO3 and Ni(NO3)2·6H2O was added into the SiO2 support at room temperature by incipient wetness impregnation, followed by drying with a rotary evaporator for 2 hours at 80 °C, and then overnight at 120 °C. Afterwards, the dried solid materials were calcined in air for 1 hour at 300 °C with a heating rate of 1 °C/min and another 3 hours at 550 °C with a heating rate of 2 °C/min. The calcined samples were denoted as Ni/SiO2-x/y, where x and y indicate the molar ratio of C2H5NO2/Ni(NO3)2·6H2O and NH4NO3/Ni(NO3)2·6H2O, respectively. The metallic Ni loading was 10 wt%. The samples were then crushed and sieved into a 40–60 mesh size for subsequent catalytic tests.

3.2. Catalyst Characterization

Fresh, reduced and spent samples were characterized by several techniques to identify and infer the effects of combustible materials such as C2H5NO2 and NH4NO3 on the catalyst morphology and the resulting catalytic performance. N2 adsorption-desorption isotherms for each sample were collected on a Micromeritics ASAP 2020 system (Micromeritics, Norcross, GA, USA). The surface area, pore size and pore volume were calculated with the N2 adsorption-desorption isotherms via the conventional Barrett–Joyner–Halenda (BJH) and Brunauer–Emmett–Teller (BET) methods. Prior to the measurements, the samples were outgassed under vacuum for 5 hours at 200 °C. The X-ray diffraction (XRD) patterns of each reduced sample were obtained with a Bruker AXS D8 Advance diffractometer (Bruker AXS, Karlsruhe, Germany) with Cu Kα radiation (λ=1.5406 Å) at a scanning rate of 6°/min with the 2θ range of 10–90°. The reducibility of the catalyst was studied by the H2 temperature-programmed reduction (H2-TPR) in an auto-controlled flow reactor system of TP-5076, which is equipped with a thermal conductivity detector (TCD, Tianjin Xianquan Co., China). The sample of 50 mg was pretreated in N2 stream at 200 °C for 1 hour. Additionally, when the temperature cooled down to 30 °C, the sample was heated to 950 °C at a heating rate of 10 °C/min in the H2/N2 flow (5 vol.% H2 in N2) of 30 mL/min. The H2-TPR spectra were obtained at the temperature range of 50–950 °C. The carbon accumulation in spent samples after reaction for 50 hours was determined by thermogravimetric (TG) analysis on a Mettler–Toledo TGA-1100SF thermogravimetric analyzer (Mettler-Toledo, Greifensee, Switzerland).

3.3. Catalytic Test

The dry reforming of CH4 with CO2 was performed at atmospheric pressure in a continuous-flow fixed bed quartz tube reactor with an inner diameter of 9 mm. For the typical experiment, 200 mg of shaped catalyst was filled into the center of the reactor. Before starting the reforming reaction, the catalyst was pre-reduced to 750 °C and atmospheric pressure for 2 hours in an H2 flow of 60 mL/min with a heating rate of 10 °C/min. After that, the reactor temperature was elevated to 800 °C, and then a flow of gas mixture with a molar ratio of CH4/CO2/N2 = 9/9/2 was fed with a flow rate of 160 mL/min. The products were analyzed by online gas chromatography (Agilent GC 7820A, Agilent, USA). CH4, CO2, H2, N2 and CO were measured by a TCD detector with a 5A molecular sieve column and a Porapak Q column. Additionally, 10% of N2 was employed as an internal standard. The conversions of CH4 and CO2 were calculated with the following formulas:
XCH4 = (FCH4-inFCH4-out)/FCH4-in × 100%
XCO2 = (FCO2-inFCO2-out)/FCO2-in × 100%
where X and F indicate the conversion and flow rate of i gas in the feed or the effluent, respectively.

4. Conclusions

In summary, the combustion method was applied to prepare SiO2 supported Ni catalysts which showed remarkably smaller Ni nanoparticle sizes due to the synergistic effects of C2H5NO2 and NH4NO3 in the combustion process. This kind of Ni/SiO2 catalyst exhibits excellent coke-resistance performance and effectively suppresses the side reaction toward RWGS compared to that prepared with the conventional wetness impregnation method. As a result, there is almost no loss of activity with the H2/CO molar ratio close to the theoretical value at 1/1 after a 50-hour stability test over the Ni/SiO2-2/1 catalyst.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/9/2/183/s1, Figure S1: XRD patterns of fresh Ni/SiO2 catalysts prepared with the combustion method by using different ratios of C2H5NO2 to NH4NO3, Figure S2: TEM images of spent Ni/SiO2-0/0 catalyst ((a) and (b)) and Ni/SiO2-2/1 catalyst ((c) and (d)) after 50-hours of reaction, Table S1: The equilibrium conversions of CH4 and CO2, H2/CO molar ratio, and selectivity to H2O calculated by HSC chemistry 6.0, Table S2: BET surface area of as-prepared Ni/SiO2 catalysts.

Author Contributions

Conceptualization, X.L. and Y.X. (Yan Xu); methodology, X.L., Y.X. (Yan Xu) and Q.L.; formal analysis, Y.X. (Yan Xu) and Q.L.; investigation, Y.X. (Yan Xu) and Q.L.; writing—original draft preparation, Y.X. (Yan Xu) and X.L.; writing—review and editing, B.L., F.J. and Y.X. (Yuebing Xu); funding acquisition, X.L.

Funding

This work is supported by the National Natural Science Foundation of China (21576119, 21878127), the Fundamental Research Funds for the Central Universities (JUSRP51720B, JUSRP11813), and the Program of Introducing Talents of Discipline to Universities (111 Project B13025).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Caballero, A.; Pérez, P.J. Methane as Raw Material in Synthetic Chemistry: The Final Frontier. Chem. Soc. Rev. 2013, 42, 8809–8820. [Google Scholar] [CrossRef] [PubMed]
  2. Schwach, P.; Pan, X.L.; Bao, X.H. Direct Conversion of Methane to Value-Added Chemicals over Heterogeneous Catalysts: Challenges and Prospects. Chem. Rev. 2017, 117, 8497–8520. [Google Scholar] [CrossRef] [PubMed]
  3. Pruett, R.L. Synthesis Gas: A Raw Material for Industrial Chemicals. Science 1981, 211, 11–16. [Google Scholar] [CrossRef] [PubMed]
  4. Reubroycharoen, P.; Yamagami, T.; Vitidsant, T.; Yoneyama, Y.; Ito, M.; Tsubaki, N. Continuous Low-Temperature Methanol Synthesis from Syngas Using Alcohol Promoters. Energy Fuels 2003, 17, 817–821. [Google Scholar] [CrossRef]
  5. Xiong, H.F.; Jewell, L.L.; Coville, N.J. Shaped Carbons as Supports for the Catalytic Conversion of Syngas to Clean Fuels. ACS Catal. 2015, 5, 2640–2658. [Google Scholar] [CrossRef]
  6. Zhang, Q.H.; Kang, J.C.; Wang, Y. Development of Novel Catalysts for Fischer-Tropsch Synthesis: Tuning the Product Selectivity. ChemCatChem 2010, 2, 1030–1058. [Google Scholar] [CrossRef]
  7. Jiang, F.; Liu, B.; Li, W.P.; Zhang, M.; Li, Z.J.; Liu, X.H. Two-Dimensional Graphene-Directed Formation of Cylindrical Iron Carbide Nanocapsules for Fischer–Tropsch Synthesis. Catal. Sci. Technol. 2017, 7, 4609–4621. [Google Scholar] [CrossRef]
  8. Zheng, J.; Cai, J.; Jiang, F.; Xu, Y.B.; Liu, X.H. Investigation of the Highly Tunable Selectivity to Linear α-Olefins in Fischer-Tropsch Synthesis over Silica-Supported Co and CoMn Catalysts by Carburization–Reduction Pretreatment. Catal. Sci. Technol. 2017, 7, 4736–4755. [Google Scholar] [CrossRef]
  9. Jiang, F.; Zhang, M.; Liu, B.; Xu, Y.B.; Liu, X.H. Insights into the Influence of Support and Potassium or Sulfur Promoter on Iron-based Fischer-Tropsch Synthesis: Understanding the Control of Catalytic Activity, Selectivity to Lower Olefins, and Catalyst Deactivation. Catal. Sci. Technol. 2017, 7, 1245–1265. [Google Scholar] [CrossRef]
  10. Cai, J.; Jiang, F.; Liu, X.H. Exploring Pretreatment Effects in Co/SiO2 Fischer-Tropsch Catalysts: Different Oxidizing Gases Applied to Oxidation-Reduction Process. Appl. Catal. B Environ. 2017, 210, 1–13. [Google Scholar] [CrossRef]
  11. Liu, X.H.; Tokunaga, M. Controllable Fischer-Tropsch Synthesis by in Situ-Produced 1-Olefins. ChemCatChem 2010, 2, 1569–1572. [Google Scholar] [CrossRef]
  12. Liu, X.H.; Linghu, W.S.; Li, X.H.; Asami, K.; Fujimoto, K. Effects of Solvent on Fischer-Tropsch Synthesis. Appl. Catal. A Gen. 2006, 303, 251–257. [Google Scholar] [CrossRef]
  13. Pakhare, D.; Spivey, J. A review of Dry (CO2) Reforming of Methane over Noble Metal Catalysts. Chem. Soc. Rev. 2014, 43, 7813–7837. [Google Scholar] [CrossRef] [PubMed]
  14. Ross, J.R.H. Natural Gas Reforming and CO2 Mitigation. Catal. Today 2005, 100, 151–158. [Google Scholar] [CrossRef]
  15. Bitter, J.H.; Seshan, K.; Lercher, J.A. On the Contribution of X-ray Absorption Spectroscopy to Explore Structure and Activity Relations of Pt/ZrO2 Catalysts for CO2/CH4 Reforming. Top. Catal. 2000, 10, 295–305. [Google Scholar] [CrossRef]
  16. Galuszka, J.; Pandey, R.N.; Ahmed, S. Methane Conversion to Syngas in a Palladium Membrane Reactor. Catal. Today 1998, 46, 83–89. [Google Scholar] [CrossRef]
  17. Vooradi, R.; Bertran, M.O.; Frauzem, R.; Anne, S.B.; Gani, R. Sustainable Chemical Processing and Energy-Carbon Dioxide Management: Review of Challenges and Opportunities. Chem. Eng. Res. Des. 2018, 131, 440–464. [Google Scholar] [CrossRef]
  18. Peter, S.C. Reduction of CO2 to Chemicals and Fuels: A Solution to Global Warming and Energy Crisis. ACS Energy Lett. 2018, 3, 1557–1561. [Google Scholar] [CrossRef]
  19. Ha, K.S.; Bae, J.W.; Woo, K.J.; Jun, K.W. Efficient Utilization of Greenhouse Gas in a Gas-to-Liquids Process Combined with Carbon Dioxide Reforming of Methane. Environ. Sci. Technol. 2010, 44, 1412–1417. [Google Scholar] [CrossRef]
  20. De Vasconcelos, B.R.; Minh, D.P.; Lyczko, N.T.; Phan, S.; Sharrock, P.; Nzihou, A. Upgrading Greenhouse Gases (Methane and Carbon Dioxide) into Syngas using Nickel-Based Catalysts. Fuel 2018, 226, 195–203. [Google Scholar] [CrossRef]
  21. Ginsburg, J.M.; Pina, J.; El Solh, T.; De Lasa, H.I. Coke Formation over a Nickel Catalyst under Methane Dry Reforming Conditions:  Thermodynamic and Kinetic Models. Ind. Eng. Chem. Res. 2005, 44, 4846–4854. [Google Scholar] [CrossRef]
  22. Schulz, L.A.; Kahle, L.C.S.; Delgado, K.H.; Schunk, S.A.; Jentys, A.; Deutschmann, O.; Lercher, J.A. On the Coke Deposition in Dry Reforming of Methane at Elevated Pressures. Appl. Catal. A Gen. 2015, 504, 599–607. [Google Scholar] [CrossRef]
  23. Bitter, J.H.; Seshan, K.; Lercher, J.A. Deactivation and Coke Accumulation during CO2/CH4 Reforming over Pt Catalysts. J. Catal. 1999, 183, 336–343. [Google Scholar] [CrossRef]
  24. Mette, K.; Kühl, S.; Tarasov, A.; Willinger, M.G.; Kröhnert, J.; Wrabetz, S.; Trunschke, A.; Scherzer, M.; Girgsdies, F.; Düdder, H.; et al. High-Temperature Stable Ni Nanoparticles for the Dry Reforming of Methane. ACS Catal. 2016, 6, 7238–7248. [Google Scholar] [CrossRef]
  25. Pawar, V.; Ray, D.; Subrahmanyam, C.; Janardhanan, V.M. Study of Short-Term Catalyst Deactivation due to Carbon Deposition during Biogas Dry Reforming on Supported Ni Catalyst. Energy Fuels 2015, 29, 8047–8052. [Google Scholar] [CrossRef]
  26. Li, S.R.; Gong, J.L. Strategies for Improving the Performance and Stability of Ni-Based Catalysts for Reforming Reactions. Chem. Soc. Rev. 2014, 43, 7245–7256. [Google Scholar] [CrossRef] [PubMed]
  27. Aleksandrov, H.A.; Pegios, N.; Palkovits, R.; Simeonov, K.; Vayssilov, G.N. Elucidation of the Higher Coking Resistance of Small Versus Large Nickel Nanoparticles in Methane Dry Reforming via Computational Modeling. Catal. Sci. Technol. 2017, 7, 3339–3347. [Google Scholar] [CrossRef]
  28. Christensen, K.O.; Chen, D.; Lødeng, R.; Holmen, A. Effect of Supports and Ni Crystal Size on Carbon Formation and Sintering during Steam Methane Reforming. Appl. Catal. A Gen. 2006, 314, 9–22. [Google Scholar] [CrossRef]
  29. Gonzalez-Delacruz, V.M.; Pereniguez, R.; Ternero, F.; Holgado, J.P.; Caballero, A. Modifying the Size of Nickel Metallic Particles by H2/CO Treatment in Ni/ZrO2 Methane Dry Reforming Catalysts. ACS Catal. 2011, 1, 82–88. [Google Scholar] [CrossRef]
  30. Baudouin, D.; Rodemerck, U.; Krumeich, F.; de Mallmann, A.; Szeto, K.C.; Ménard, H.; Veyre, L.; Candy, J.P.; Webb, P.B.; Thieuleux, C.; et al. Particle size Effect in the Low Temperature Reforming of Methane by Carbon Dioxide on Silica-Supported Ni Nanoparticles. J. Catal. 2013, 297, 27–34. [Google Scholar] [CrossRef]
  31. Margossian, T.; Larmier, K.; Kim, S.M.; Krumeich, F.; Fedorov, A.; Chen, P.; Müller, C.R.; Copéret, C. Molecularly-Tailored Nickel Precursor and Support Yield a Stable Methane Dry Reforming Catalyst with Superior Metal Utilization. J. Am. Chem. Soc. 2017, 139, 6919–6927. [Google Scholar] [CrossRef] [PubMed]
  32. Abba, M.O.; Gonzalez-DelaCruz, V.M.; Colón, G.; Sebti, S.; Caballero, A. In Situ XAS Study of an Improved Natural Phosphate Catalyst for Hydrogen Production by Reforming of Methane. Appl. Catal. B Environ. 2014, 150–151, 459–465. [Google Scholar] [CrossRef]
  33. Tian, H.; Li, X.Y.; Zeng, L.; Gong, J.L. Recent Advances on the Design of Group VIII Base-Metal Catalysts with Encapsulated Structures. ACS Catal. 2015, 5, 4959–4977. [Google Scholar] [CrossRef]
  34. 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]
  35. Tomishige, K.; Yamazaki, O.; Chen, Y.G.; Yokoyama, K.; Li, X.H.; Fujimoto, K. Development of Ultra-Stable Ni Catalysts for CO2 Reforming of Methane. Catal. Today 1998, 45, 35–39. [Google Scholar] [CrossRef]
  36. Chen, Y.G.; Tomishige, K.; Yokoyama, K.; Fujimoto, K. Promoting Effect of Pt, Pd and Rh Noble Metals to the Ni0.03Mg0.97O Solid Solution Catalysts for the Reforming of CH4 with CO2. Appl. Catal. A Gen. 1997, 165, 335–347. [Google Scholar] [CrossRef]
  37. Li, Z.W.; Mo, L.Y.; 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]
  38. Huang, Q.; Fang, X.Z.; Cheng, Q.Z.; Li, Q.; Xu, L.J.; Xu, X.L.; Liu, W.M.; Gao, Z.X.; Zhou, W.F.; Wang, X. Synthesis of Highly Active and Stable Ni@Al2O3 Embedded Catalyst for Methane Dry Reforming: On the Confinement Effects of Al2O3 Shells for Ni Nanoparticles. ChemCatChem 2017, 9, 3563–3571. [Google Scholar] [CrossRef]
  39. Varma, A.; Mukasyan, A.S.; Rogachev, A.S.; Manukyan, K.V. Solution Combustion Synthesis of Nanoscale Materials. Chem. Rev. 2016, 116, 14493–14586. [Google Scholar] [CrossRef]
  40. Manukyan, K.V.; Chen, Y.S.; Rouvimov, S.; Li, P.; Li, X.; Dong, S.; Liu, X.Y.; Furdyna, J.K.; Orlov, A.; Bernstein, G.H.; et al. Ultrasmall α-Fe2O3 Superparamagnetic Nanoparticles with High Magnetization Prepared by Template-Assisted Combustion Process. J. Phys. Chem. C 2014, 118, 16264–16271. [Google Scholar] [CrossRef]
  41. Gao, X.Y.; Hidajat, K.; Sawi, S. Facile synthesis of Ni/SiO2 catalyst by sequential hydrogen/air treatment: A superior anti-coking catalyst for dry reforming of methane. J. CO2 Util. 2016, 15, 146–153. [Google Scholar] [CrossRef]
Scheme 1. One-step facile synthesis of Ni catalysts supported on silica (SiO2) prepared by the combustion of Ni(NO3)2–C2H5NO2–NH4NO3 impregnated in the porous SiO2.
Scheme 1. One-step facile synthesis of Ni catalysts supported on silica (SiO2) prepared by the combustion of Ni(NO3)2–C2H5NO2–NH4NO3 impregnated in the porous SiO2.
Catalysts 09 00183 sch001
Figure 1. (a) XRD patterns of reduced Ni/SiO2 catalysts prepared with the combustion method by using different ratios of C2H5NO2 to NH4NO3. (b,c) TEM images and Ni size distribution of the reduced Ni/SiO2-0/0 and Ni/SiO2-2/1 catalysts, respectively. (d) H2-TPR profiles of the fresh Ni/SiO2 catalysts prepared with the combustion method.
Figure 1. (a) XRD patterns of reduced Ni/SiO2 catalysts prepared with the combustion method by using different ratios of C2H5NO2 to NH4NO3. (b,c) TEM images and Ni size distribution of the reduced Ni/SiO2-0/0 and Ni/SiO2-2/1 catalysts, respectively. (d) H2-TPR profiles of the fresh Ni/SiO2 catalysts prepared with the combustion method.
Catalysts 09 00183 g001
Figure 2. CH4 conversion (a), CO2 conversion (b), and H2/CO molar ratio (c) as a function of time on stream for Ni/SiO2 catalysts prepared with the combustion method at different molar ratios of C2H5NO2 to NH4NO3 (black line: conventional wetness impregnation method; purple line: NH4NO3 only; blue line: C2H5NO2 only; red line: 2/1 ratio of C2H5NO2 to NH4NO3). The reaction was carried at 800 °C with 200 mg of catalyst and a molar ratio of CH4/CO2/N2 = 9/9/2 with 160 mL/min.
Figure 2. CH4 conversion (a), CO2 conversion (b), and H2/CO molar ratio (c) as a function of time on stream for Ni/SiO2 catalysts prepared with the combustion method at different molar ratios of C2H5NO2 to NH4NO3 (black line: conventional wetness impregnation method; purple line: NH4NO3 only; blue line: C2H5NO2 only; red line: 2/1 ratio of C2H5NO2 to NH4NO3). The reaction was carried at 800 °C with 200 mg of catalyst and a molar ratio of CH4/CO2/N2 = 9/9/2 with 160 mL/min.
Catalysts 09 00183 g002
Figure 3. TG patterns of spent Ni/SiO2 catalysts after the dry reforming (DRM) reaction of 50 hours. The catalytic results are shown in Figure 2.
Figure 3. TG patterns of spent Ni/SiO2 catalysts after the dry reforming (DRM) reaction of 50 hours. The catalytic results are shown in Figure 2.
Catalysts 09 00183 g003

Share and Cite

MDPI and ACS Style

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. https://doi.org/10.3390/catal9020183

AMA Style

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(2):183. https://doi.org/10.3390/catal9020183

Chicago/Turabian Style

Xu, Yan, Qiang Lin, Bing Liu, Feng Jiang, Yuebing Xu, and Xiaohao Liu. 2019. "A Facile Fabrication of Supported Ni/SiO2 Catalysts for Dry Reforming of Methane with Remarkably Enhanced Catalytic Performance" Catalysts 9, no. 2: 183. https://doi.org/10.3390/catal9020183

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